ABSTRACT

MEUNIER, CARL JOSEPH. Advancing and Analysis Strategies for Enhanced Throughput, Quantification, Chemical Diversity, and Selectivity. (Under the direction of Dr. Leslie A. Sombers).

The central nervous system (CNS) is a heterogeneous, complex, and plastic network of cells that regulates a myriad of autonomous bodily functions, perception and cognition, as well as emotion and behavior. CNS function is largely controlled by regulating neurotransmission, in which chemicals released from neurons bind with protein receptors eliciting cellular change.

Electrochemical techniques, especially when combined with carbon-fiber microelectrodes

(CFMEs), are suited for measuring at functional time and spatial resolutions.

Namely, background-subtracted fast-scan cyclic voltammetry (FSCV) has been invaluable to exciting advances in neuroscience. Yet, significant improvements in the application of FSCV would expand the utility and adoption of the technique. The work presented herein focuses largely on developing methods to improve selectivity against complex interfering signals, improve quantification of in vivo recordings, and enhance FSCV for monitoring neuropeptides.

Advances in engineered electrodes and the drive to explore other brain regions have expanded the detectable pool of analytes and interferents. In many cases, multiple species can be detected simultaneously at a single recording site. Herein, improvements were made to a custom modified-sawhorse waveform (MSW) previously developed for peptide detection. By thoroughly characterzing MSW parameter impact on peptide response, an optimized waveform was developed and used to record the first known neuropeptide dynamics in vivo. Further, the standard analysis strategy for these complex recordings, principal component regression, is unreliable in situations when signals share sources of variance and requires careful construction of a training set containing all components of a complex signal. Furthermore, the need for background-subtraction limits the analysis window to <90 seconds, as slow changes to the background signal result in subtraction artifacts called electrochemical drift. Herein, a combined voltammetry and regression-based predictive modeling approach was developed. By using a double-triangular waveform in conjunction with partial least-squares regression (DW-PLSR), signals with significant shared variance could be elucidated. Furthermore, DW-PLSR was capable of subtracting electrochemical drift extending the analysis window to >10 minutes, was demonstrated to be tailorable for recordings of other analytes, and simplified training set contruction; only requiring signals from interfering species.

Additional research focused on developing methods to enhance quantification of measurements in live tissue. Tissue exposure reduces electrode senstivity and can result in shifts to cyclic voltammogram position and shape vital for effective multivariate calibration.

Traditionally, FSCV is performed at acutely implanted electrodes that were post-calibrated in buffer, offering some measure of altered electrode performance. Now it is common for week to month long experiments using permantely implanted electrodes. However, the inability to recover electrodes led to the use of standard data sets for analysis (data recorded in vitro or from other electrodes/animals), which have been demonstrated to be unreliable. Herein, the effect of impedance changes on recordings was explored; changes in the ionic compoisiton, adsorption, electrode fouling, etc., result in impedance changes. By modulating impedance in vitro, performance changes observed in live tissue could be mimicked and used to predict changes in electrode performance. Furthermore, electrochemical impedance spectroscopy (EIS) was used to develop equivalent circuit models for CFMEs, which were used to evaluate changes at the electrode/solution interface. It was determined that electrode behavior deviates further from an ideal capacitor and surface area decreases in tissue, presumably due to fouling. Then, a multiplexed FSCV and EIS paradigm was developed that allowed real-time tracking of sensor impedance/capacitance without sacrificing temporal resolution.

Overall, the experiments and advances described herein provide tools that enable researchers to target new molecules, enahance selectivity, increase throughput, and inform regarding changing electrode performance. Ultimately, these works will expand the utility of

FSCV, facilitate advances in neuroscience, and prove useful for real-time electrochemical monitoring in other applications.

© Copyright 2020 by Carl Joseph Meunier

All Rights Reserved Advancing Voltammetry and Analysis Strategies for Enhanced Throughput, Quantification, Chemical Diversity, and Selectivity

by Carl Joseph Meunier

A dissertation submitted to the Graduate Faculty of North Carolina State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy

Chemistry

Raleigh, North Carolina 2020

APPROVED BY:

______Leslie A. Sombers David C. Muddiman Committee Chair

______Erin S. Baker Douglas Call

DEDICATION

This work is dedicated to family and friends, past and present, who have all been integral in my progression that has led me to closing this chapter in my life.

My parents, Arthur and Rosemary Meunier, have raised three independent, adaptable, and kindhearted children who have spread their wings and bring positive energy to all situations. They have always encouraged critical thinking, supported me in my pursuits, and provided guidance when I would fall off course. This work is dedicated to them.

This work is dedicated to my late grandparents John and Rosalee Balbach, my late grandfather Arthur Meunier, and my grandmother Judith Meunier.

Finally, this work is dedicated to the next generations of my family. I hope that one day this accomplishment inspires them to question the world, and work to uncover its mysteries.

You’re only as old as your ability to process new information.

-Phonte Coleman, 2018

ii

BIOGRAPHY

Carl Joseph Meunier was born in Madison, WI, on September 8th, 1992, to parents Arthur and Rosemary Meunier; his sister, Marie Meunier, was born three years earlier. His family moved to Dyersville, IA where he spent his formative years; his brother, Allen Meunier, was born there.

His family then relocated, spending time in both Virginia, IL and Jacksonville, IL, where he attended high school.

Carl grew up in ‘small town’ America, where he developed a love for the outdoors, sports, and independence. A good portion of his youth was spent riding bikes, camping, fishing, and playing sports (basketball, baseball, football, golf). He acquired a diverse skill and was introduction to critical thinking through his involvement in Boy Scouts as well as his involvement in remodeling multiple homes with his parents. During high school, Carl played varsity basketball, golf, and baseball, was a talented trombonist, became an Eagle Scout, and led a group to design and construct a record setting pumpkin-throwing trebuchet, which certainly did not hurt his interest in the sciences. He did various jobs such as baling hay on a farm, pressing shirts at a dry-cleaner, working as a handyman for a local business, and as a jack-of-all-trades at a grocery store. Through these experiences Carl developed an appreciation for hard work, a strong social awareness, and a drive to acquire knowledge.

In the fall of 2011, Carl began undergraduate studies at Bradley University in Peoria, IL, where he graduated magna cum laude with a Bachelor of Science in both Chemistry and Spanish, graduating magna cum laude. During his undergraduate years, he joined the Illinois Delta chapter of Sigma Phi Epsilon fraternity, the Chemistry Club, and Phi Sigma Iota, an honor society for foreign languages and literature. Carl spent a summer working in the Zoppe Family Circus as a crewmember/stagehand with a close friend, performing hard manual labor, living in filth, making

iii

new friends, and enjoying every moment. He also spent a summer studying abroad in Spain further expanding his horizons. Having traveled alone, he quickly learned to adapt to new environments, expanded his social skills, and became introspective. He also became heavily involved in teaching by becoming a teaching assistant for general and organic chemistry labs, as well as in research under the direction of Dr. Luke M. Haverhals. He received numerous awards for his academic performance, service to the Department of Chemistry and , and demonstrated potential as a researcher. At Bradley, Carl also met his partner, Nicole Smiddy, who was also studying Chemistry.

Nearing his final year at Bradley, a number of faculty members suggested Carl apply to graduate school. In the summer of 2015, he moved to Raleigh, NC to begin his doctoral studies in the Department of Chemistry at North Carolina State University under the direction of Dr. Leslie

A. Sombers. Here, has worked on advancing electrochemical sensing technology for applications in neuroscience, gained a strong network of friends, and was in solidarity with Nicole, who was working on her Ph.D. in Chemistry at UNC-Chapel Hill.

iv

ACKNOWLEDGMENTS

To my parents, Arthur and Rosemary Meunier, I would like to express my gratitude in all you have done for me and for guiding me on my journey to become a man. You instilled in me an appreciation for persistence and hard work, allowed me the freedom to form my own opinions, granted me the independence that gave me courage to tackle the unknown, and provided me a moral framework that helped form my affable disposition. To my siblings, Marie and Allen

Meunier, although are interactions are too few nowadays, I thank you for the many late nights spent catching up and providing counsel for each other as we continue to mature as adults. To my beautiful partner, Nicole Smiddy, I am fortunate to have shared the graduate school experience with you. Together we have made it through some stressful times, provided each other comfort and support, and formed a deeper connection with each other. I look forward to continuing life’s journey with you. To all my friends that have gave their time, energy, and resources during my doctoral studies, I sincerely thank you for providing support, and sharing in the good and bad times. I owe each one of you a beer.

To educators that helped spark my interest in science and broadened my exposure to the world of science. As a young man from rural Illinois, I never would have thought that I would be in my current position. The entire chemistry faculty at Bradley provided me solid fundamental knowledge, made teaching and research enjoyable, and instilled confidence in me pursuing graduate studies. I would like to thank Dr. Reza Ghiladi (NC State) for providing guidance that helped me obtain a National Science Foundation Graduate Research Fellowship enabling me to support myself during my graduate studies.

Importantly, I would like to thank Dr. Leslie Sombers for her continual support and guidance during my graduate career. Leslie afforded me independence to develop and pursue some

v

of my own ideas, serve as a leader within the lab, and opportunities to gain experience communicating our work at conferences or through preparation of manuscripts and grants that all have made me a better scientist. To all lab members, past and current that I have overlapped with,

I thank you for the support, ideas, and discussions. I want to thank Dr. Gregory McCarty and Dr.

James Roberts for being catalysts in advancing my work, impactful mentors, fundamental in helping me establish my footing in the lab, and friends. I want to thank Christie Lee for always being willing to assist me in animal studies, sharing stupid jokes, and for the friendship we developed. Furthermore, I would like to thank Dr. David Muddiman, Dr. Erin Baker, Dr. Douglas

Call, and Dr. Gufeng Wang for serving as members of my doctoral committee.

Finally, none of this work would have been possible without funding from the Department of Chemistry at North Carolina State University, the National Science Foundation, and the

National Institute of Health.

vi

TABLE OF CONTENTS

LIST OF FIGURES ...... xii

CHAPTER 1: Introduction to Neuron Function, Approaches for Monitoring Neurotransmission, and Dissertation Overview...... 1 1.1 Neurons, Neurotransmission, and Brain Plasticity ...... 1 1.2 Methods for Monitoring Neurotransmission in Live Biological Preparations ...... 5 1.2.1 Electrophysiology ...... 5 1.2.2 Microdialysis and Solid-Phase Extraction Sampling with Mass Spectrometric Detection ...... 6 1.2.3 Electrochemical Methods at Microelectrodes ...... 8 1.3 Critical Issues Associated with In Vivo Electrochemistry ...... 10 1.3.1 Chemical Selectivity ...... 10 1.3.2 Complex Data Analysis...... 11 1.3.3 Quantification of Measurements...... 12 1.4 Electrochemical Impedance Spectroscopy ...... 13 1.5 Dissertation Overview ...... 14 1.6 References ...... 16

CHAPTER 2: Fast-Scan Voltammetry for In Vivo Measurements of Neurochemical Dynamics ...... 23 2.1 Introduction ...... 23 2.1.1 Use of FSCV in Analyte Detection...... 24 2.2 Materials ...... 26 2.2.1 Electrode Fabrication ...... 26 2.2.2 Surgery ...... 27 2.2.3 Electrochemistry ...... 28 2.2.4 Electrode Calibration ...... 29 2.3 Methods ...... 29 2.3.1 Electrochemistry, Instrumentation, and Software ...... 29 2.3.2 Electrode Fabrication ...... 34 2.3.3 Surgery for Electrode Implantation in Rat Brain Tissue ...... 39 2.3.4 Voltammetric Measurements...... 43 2.3.5 Post-Experiment ...... 45 2.3.6 Data Analysis ...... 47 2.3.7 Additional Experimental Options ...... 48 2.3.8 Critical Considerations and Alternative Methods...... 51 2.4 Conclusions ...... 58 2.5 References ...... 59

CHAPTER 3: Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin...... 68 3.1 Introduction ...... 68 3.2 Methods ...... 71 3.2.1 Chemicals ...... 71

vii

3.2.2 Microelectrode Fabrication ...... 71 3.2.3 Electrochemical Data Acquisition In Vitro ...... 72 3.2.4 Animal Subjects and Care ...... 73 3.2.5 Ex Vivo Adrenal Slice Preparation ...... 74 3.2.6 In Vivo Experiments ...... 74 3.2.7 Statistics and Graphics ...... 76 3.3 Results and Discussion ...... 76 3.3.1 An Introduction to the Modified Sawhorse Waveform ...... 76 3.3.2 Waveform Application Frequency...... 78 3.3.3 Accumulation Potential and Scan Rate ...... 79 3.3.4 Amperometric Potential and Transition Potential ...... 81 3.3.5 MSW 1.0 vs MSW 2.0: A Direct Comparison ...... 84 3.3.6 Simultaneous Measurements of CA and M-ENK Fluctuations in Living Adrenal Tissue ...... 85 3.3.7 Simultaneous Measurements of CA and M-ENK in the Dorsal Striatum ...... 87 3.4 Conclusions ...... 91 3.5 References ...... 92

CHAPTER 4: Electrochemical Selectivity Achieved Using a Double Voltammetric Waveform and Partial Least Squares Regression: Differentiating Endogenous Hydrogen Peroxide Fluctuations from Shifts in pH ...... 98 4.1 Introduction ...... 98 4.2 Experimental Section ...... 101 4.2.1 Chemicals ...... 101 4.2.2 Microelectrode Fabrication ...... 101 4.2.3 Flow-Injection System ...... 102 4.2.4 Data Acquisition ...... 102 4.2.5 Animal Experiments ...... 103 4.2.6 Data Processing and Analysis ...... 104 4.3 Results and Discussion ...... 104 4.3.1. FSCV Detection of H2O2 and ΔpH ...... 104 4.3.2 Principal Component Regression: Complications in Distinguishing H2O2 and ΔpH ...... 107 4.3.3 Double Waveform-Partial Least Squares Regression Model ...... 109 4.3.4 Construction of an In Vivo Training Set ...... 113 4.3.5 Applying the DW-PLSR Model to In Vivo Data ...... 115 4.4 Conclusion ...... 120 4.5 References ...... 122

CHAPTER 5: Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double Waveform Partial Least Squares Regression ...... 128 5.1 Introduction ...... 128 5.2 Materials and Methods ...... 131 5.2.1 Data Acquisition ...... 131 5.2.2 Animal Experiments ...... 131 5.2.3 Data Processing and Analysis ...... 132

viii

5.3 Results and Discussion ...... 132 5.3.1 Electrochemical Drift in FSCV Recordings ...... 132 5.3.2 The Double-Waveform Partial-Least-Squares Regression Model ...... 133 5.3.3 Subtraction of Drift from Extended Dopamine Recordings ...... 137 5.3.4 Drift Subtraction from Extended In Vivo Recordings...... 140 5.3.5 Integrating Drift Signals from Multiple Electrodes for Drift Subtraction ...... 144 5.4 Conclusions ...... 147 5.5 References ...... 149

CHAPTER 6: The Background Signal as an In Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data ...... 155 6.1 Introduction ...... 155 6.2 Methods ...... 158 6.2.1 Chemicals ...... 158 6.2.2 Microelectrode Fabrication ...... 159 6.2.3 Model Circuit Fabrication and Characterization ...... 159 6.2.4 Flow-Injection System ...... 159 6.2.5 Data Acquisition ...... 160 6.2.6 Animal Experiments ...... 160 6.2.7 Statistics ...... 162 6.3 Results and Discussion ...... 162 6.3.1 Combining Fast-Scan Cyclic Voltammetry and Principal Component Regression ...... 162 6.3.2 The Carbon-Fiber Microelectrode Surface Influences Oxidation Potential...... 164 6.3.3 Model Circuits to Simulate Impedance Contributions Due to Electrode Fouling ...... 166 6.3.4 Employing Model Circuits to Investigate How Impedance Shifts Potential ...... 169 6.3.5 Validation of the Impedance Model ...... 172 6.4 Conclusion ...... 175 6.5 References ...... 176

CHAPTER 7: Interpreting Dynamic Interfacial Changes at Carbon Fiber Microelectrodes Using Electrochemical Impedance Spectroscopy ...... 181 7.1 Introduction ...... 181 7.2 Experimental Section ...... 183 7.2.1 Chemicals ...... 183 7.2.2 Electrode Fabrication ...... 184 7.2.3 Data Acquisition ...... 184 7.2.4 Animal Experiments ...... 185 7.2.5 Data Processing and Analysis ...... 186 7.3 Results and Discussion ...... 186 7.3.1 Overview of the Electrochemical System and EIS ...... 186 7.3.2 Determining Equivalent Circuits for CFMEs ...... 188 7.3.3 CFME Equivalent Circuits Inform on the Nature of the Electrode/Solution Interface ...... 194

ix

7.3.4 Tissue Exposure Shifts Impedance ...... 196 7.4 Conclusion ...... 199 7.5 References ...... 201

CHAPTER 8: Tracking Carbon Microelectrode Impedance Time-Locked with Fast-Scan Cyclic Voltammetry ...... 206 8.1 Introduction ...... 206 8.2 Experimental Section ...... 209 8.2.1 Data Acquisition ...... 209 8.2.2 Animal Experiments ...... 210 8.2.3 Data Processing and Statistics ...... 212 8.3 Results and Discussion ...... 213 8.3.1 Surface State/Microenvironment Impact FSCV Performance ...... 213 8.3.2 Performing Impedance Measurements Using FSCV Specific Hardware ...... 215 8.3.3 CFME Impedance Properties are Paired with FSCV Performance ...... 218 8.3.4 Combined Impedimetric and Voltammetric Measurement with High Temporal Resolution ...... 222 8.3.5 Critical Considerations Using the Multiplexed Paradigm ...... 226 8.4 Conclusions ...... 228 8.5 References ...... 229

APPENDICES ...... 234

APPENDIX A: Supplemental Information to Chapter 4 ...... 235 A.1 Summary ...... 235 A.2 Table of Contents ...... 235 A.3 Supplemental Figures ...... 236

APPENDIX B: Supplemental Information to Chapter 5 ...... 238 B.1 Summary...... 238 B.2 Table of Contents ...... 238 B.3 Supplemental Figures ...... 239 B.4 Additional Methods ...... 242 B.4.1 Chemicals ...... 242 B.4.2 Microelectrode Fabrication ...... 243 B.4.3 Flow-Injection System ...... 243

APPENDIX C: Supplemental Information to Chapter 8 ...... 244 C.1 Summary...... 244 C.2 Table of Contents ...... 244 C.3 Relevant Equations...... 245 C.4 Additional Data Processing ...... 246 C.4.1 Notes ...... 247 C.5 Supplemental Figures ...... 248 C.6 Additional Methods ...... 250 C.6.1 Chemicals ...... 250

x

C.6.2 Microelectrode Fabrication ...... 250 C.6.3 Model Circuit Fabrication and Characterization ...... 250 C.6.4 Flow-Injection System ...... 251 C.7 References ...... 252

APPENDIX D: Poly-Dioxythiophene and Nafion Copolymer Membranes for Enhanced Voltammetric Peptide Selectivity ...... 253 D.1 Introduction ...... 253 D.2 Methods ...... 254 D.2.1 Chemicals ...... 254 D.2.2 Data Acquisition ...... 255 D.2.3 Electropolymerization ...... 255 D.3 Results and Discussion ...... 256 D.3.1 Peptide FSCV Before and After Electrodeposition of PEDOT:Nafion Coating ...... 256 D.3.2 Tailoring the Copolymer Coating to Alter Peptide Selectivity...... 258 D.4 Conclusion ...... 261 D.5 References ...... 263

APPENDIX E: Fourier Analysis of FSCV Recordings to Identify Time-Points of Molecular Fluctuations ...... 264 E.1 Introduction ...... 264 E.2 Methods ...... 265 E.2.1 Chemicals ...... 265 E.2.2 Data Acquisition ...... 265 E.2.3 Data Analysis ...... 266 E.3 Results and Discussion ...... 266 E.3.1 Fourier Transform Fast-Scan Cyclic Voltammetry (FT-FSCV) ...... 266 E.3.2 Identification of Important Time-Points in Extended Recordings Using FT-FSCV ...... 269 E.4 Conclusion ...... 270 E.5 References ...... 272

xi

LIST OF FIGURES

Figure 1.1. Diagram summarizing neurotransmission ...... 3

Figure 2.1. Introduction to FSCV ...... 24

Figure 2.2. The FSCV potential waveform can be optimized for detection of different molecules...... 32

Figure 2.3. Representative FSCV recordings with different levels of noise ...... 39

Figure 2.4. A graphical representation of the exposed rat skull during surgery...... 42

Figure 2.5. Example of a dopamine release event ...... 45

Figure 2.6. FSCV flow cell and headstage ...... 46

Figure 2.7. Principal component regression ...... 48

Figure 2.8. Analog and digital filtering in FSCV ...... 56

Figure 2.9. Stimulation pulses are gated to prevent overlap with application of the voltammetric waveform ...... 57

Figure 3.1. Introduction to the MSW ...... 78

Figure 3.2. Waveform application frequency impacts sensitivity to MENK ...... 79

Figure 3.3. Characterization of the voltammetric signal for M-ENK when varying accumulation potential and scan rate ...... 81

Figure 3.4. Potentials selected at both nodes of the second segment of the forward scan influence the voltammetric response of M-ENK ...... 82

Figure 3.5. Transition potential that distinguishes the first segment ofthe forward scan from the second can be tailored to maximize sensitivity or selectivity ...... 83

Figure 3.6. Improved detection of M-ENK with MSW 2.0 ...... 85

Figure 3.7. Simultaneous monitoring of M-ENK and CA dynamics in an adrenal slice preparation with MSW 2.0 ...... 87

Figure 3.8. M-ENK recorded in the dorsal striatum of an anesthetized rat...... 88

Figure 3.9. Neurochemical fluctuations recorded in the dorsomedial striatum of awake, freely behaving rats ...... 90

Figure 4.1. Representative color plots and CVs for H2O2, + 0.25 basic ΔpH and a mixture of both ...... 106

xii

Figure 4.2. Principal component regression is insufficient to resolve individual contributions to the voltammetric signal from ΔpH and H2O2 ...... 108

Figure 4.3. A double waveform-partial least-squares regression (DW-PLSR) model ...... 112

Figure 4.4. Training set selection tool ...... 115

Figure 4.5. In vivo training set for PLSR model ...... 116

Figure 4.6. Application of DW-PLSR model to in vivo data collected in the striatum of a freely moving, 6-OHDA lesioned rat ...... 118

Figure 4.7. Application of the DW-PLSR model to data collected in the striatum of an intact rat ...... 120

Figure 5.1. Introduction to the double-waveform partial-least-squares regression (DW-PLSR) paradigm ...... 134

Figure 5.2. DW-PLSR effectively removes drift from a 15 min recording containing a rapid fluctuation of adenosine ...... 136

Figure 5.3. DW-PLSR can be used to effectively remove drift from 30 min recordings containing DA and H2O2 calibrations ...... 140

Figure 5.4. DW-PLSR model removes drift from data collected in the dorsal striatum of an anesthetized rat, improving visualization of endogenous H2O2 transients ...... 142

Figure 5.5. DW-PLSR model removes drift from voltammetric data collected in the ventral striatum of an anesthetized rat during a 10 min ischemic event ...... 144

Figure 5.6. DW-PLSR model constructed using drift signals from a collection of electrodes revealing both transient fluctuations and gradual increases in H2O2 concentration ...... 145

Figure 6.1. Number of references in PubMed using “fast-scan cyclic voltammetry” per year, since 1984 ...... 157

Figure 6.2. Overview of fast-scan cyclic voltammetry ...... 163

Figure 6.3. Electrode surface influences oxidation potential ...... 165

Figure 6.4. Equivalent circuit diagram depicting an in vivo electrochemical system ...... 167

Figure 6.5. Impedance characterization of model Randles circuits ...... 168

Figure 6.6. Impedance impacts the nature of the electrochemical system ...... 171

Figure 6.7. Establishing a relationship between Ep.QH and Ep,DA ...... 172

xiii

Figure 6.8. Model validation in live tissue ...... 174

Figure 7.1. Introduction to EIS and FSCV ...... 187

Figure 7.2. EIS performed from 1 Hz to 1 MHz at CFMEs insulated in glass and silica ...... 190

Figure 7.3. Equivalent model circuits used to fit EIS data ...... 193

Figure 7.4. The equivalent circuit model can be used to accurately report on controlled manipulations to the electrochemical system ...... 196

Figure 7.5. Tissue exposure impacts impedance properties ...... 199

Figure 8.1. Fast-Scan Cyclic Voltammetry and Surface State ...... 215

Figure 8.2. Validation of impedance spectroscopic measurements on FSCV hardware ...... 217

Figure 8.3. EIS measurements inform about FSCV performance ...... 219

Figure 8.4. Time-division multiplexing of FSCV and EIS measurements ...... 223

Figure 8.5. FSCV:EIS100Hz measurements in the medulla of adrenal tissue slices ...... 225

Figure 8.6. 100, 500, and 1000 Hz frequency-division multiplexing of EIS portion of FSCV:EIS waveform ...... 227

Figure A.1. Additional PLSR information for data in Figure 4.6 ...... 236

Figure A.2. Additional PLSR information for data in Figure 4.7 ...... 237

Figure B.1. In vitro drift removal for other calibrations not displayed in Figure 5.3 ...... 240

Figure B.2. Additional PLSR information for data in Figures 5.4 and 5.5 ...... 241

Figure B.3. Additonal information for data in Figure 5.6 ...... 242

Figure C.1. Examples of time-delay introduced into the EIS current response using the multiplexed setup ...... 248

Figure C.2. Comparison of 100 Hz, 500 Hz, and 1000 Hz single- and multi-frequency FSCV:EIS measurements ...... 249

Figure D.1. Schematic of the copolymerization of PEDOT and Nation that coats the CFME surface ...... 254

Figure D.2. Peptide FSCV before and after application of PEDOT:Nafion coating ...... 257

Figure D.3. Dopamine and ascorbic acid selectivity for different copolymer coatings ...... 259

xiv

Figure D.4. Copolymer coating effect on peptide FSCV peptide selectivity ...... 261

Figure E.1. Visualization of representative FSCV data in both the potential and frequency domains ...... 268

Figure E.2. Frequency domain analyses help discern electrochemical drift from analyte signals ...... 270

xv

CHAPTER 1

Introduction to Neuron Function, Approaches for Monitoring Neurotransmission, and

Dissertation Overview

1.1 Neurons, Neurotransmission, and Brain Plasticity

The central nervous system (CNS) is an unfathomably heterogeneous, complex, and continuously changing network of cells that regulates a myriad of autonomous bodily functions, perception and cognition, as well as emotion and behavior.1,2 Remarkably, the brain accomplishes all this through modulations of electrical activity in networks of neurons through release of chemical messengers called neurotransmitters. Neurons (Figure 1.1) are electrically excitable cells consisting of a cell body, axon(s), and dendrite(s). Neurotransmitters released from axon terminals can interact with membrane bound protein receptors on a dendrite of another neuron, resulting in a change in cellular response. Astrocytes and glial cells serve as the foundation and metabolic support network for neurons but have been shown to participate in modulating neuronal communication and function.3–7 For example, one astrocyte function is regulate microcirculation of blood in the brain.8 Axons project to other regions in the CNS and form 5-50 nm wide interfaces with dendrites called synapses. The connectivity of these networks is complex; >1 neuron may interface with an individual dendrite. When a neuron becomes sufficiently depolarized, vesicles containing molecules fuse with the cellular membrane and are released from axon projections, or the presynaptic neuron, into the synaptic cleft. From here, these molecules can diffuse and interact with membrane bound protein receptors, thereby triggering changes in chemical/electrical properties of the postsynaptic neuron through the dendrite.

1

Neurotransmitter release events begin with the firing and propagation of an action potential. The resting potential across a cell membrane, typically around -70 mV, arises due to differences intracellular and extracellular ionic composition; Na+ and K+ intracellular concentrations are lower and higher than extracellular, respectively.9 Transmembrane proteins called voltage-gated channels exist for these , which have pores that open at a particular cell membrane voltage threshold allowing ions to migrate down their concentration gradient across the cell membrane; other voltage-gated ion channels exist, however Na+ and K+ are predominately responsible. When the threshold potential is reached, a series of events are triggered that result in an action potential. First, voltage-gated Na+ channels open resulting in a large influx of Na+ that further depolarizes the cell to a more positive potential. Once a certain limit is reached, Na+ channels will rapidly close and K+ ion channels will open allowing K+ to flow into the extracellular space. This repolarizes the cell, at which point K+ channels will close slowly, leading to a brief refractory period in which the cell membrane is hyperpolarized before the action potential is complete. The action potential propagates through the cell in this fashion.

Once the action potential reaches the axon terminal, it can trigger vesicular release of neurotransmitters. Neurotransmitters are packaged in large dense core (LDCV) or small synaptic vesicles (SSV) which are intracellular membrane cavities. When a neuron is depolarized, voltage- gated Ca2+ channels open allow influx of Ca2+ which binds to synaptotagmin, a presynaptic membrane bound protein, initiating the fusion of vesicles to the cell membrane.10 SSVs are typically located close to the presynaptic cell membrane, contain only small molecule neurotransmitters, and make up the readily releasable pool. LDCVs contain small molecule neurotransmitters co-packaged with a chromogranin core composed of peptides and proteins.

There is evidence that multiple modes of vesicular release can be initiated through different

2

isoforms of synaptotagmin resulting in either partial or full fusion of vesicle and cell membranes, leading to partial of complete release of contents.11–15

Figure 1.1. Diagram summarizing neurotransmission. Main physical features of a neuron (top) and a synapse formed between two neurons (bottom).

The most common neurotransmitter are acetylcholine, gamma-aminobutyric acid, and glutamate, as well as monoamines such as dopamine, norepinephrine and .10 Other species such as hydrogen peroxide and neuropeptides have been demonstrated to act as neurotransmitters too.16,17 Once released, they can bind to protein receptors on the postsynaptic membrane and act as agonists (drives a particular function) or antagonists (blocks a particular function). Receptors are membrane bound G-coupled proteins that initiate or inhibit some physiological response when bound. Receptor-ligand binding can be highly specific or chemically promiscuous, and the same neurotransmitter can have opposing effects at different receptor types

3

or based on the dose/magnitude. Alternatively, neurotransmitters are eventually enzymatically broken down in the extracellular space or undergo reuptake by the presynaptic neuron through autoreceptors. During reuptake, specialized protein transports on the presynaptic membrane can transport neurotransmitters back inside the neuron.10 These processes result in tight regulation of extracellular lifetime and diffusion distance. Importantly, the expression of receptors, autoreceptors, ion-channels, and neurotransmitters can be regulated by neurons based on previous events to alter chemical communication. This results in a large degree of plasticity in the brain, and fortuitously allows for diversity among subjects.

This complexity also presents many instances where malfunctions can occur that can lead to neurological disease states that result from genetic or environmental influences. Dopaminergic and serotonergic dysfunction is implicated in neurological disorders such as schizophrenia,18–20

Parkinson’s disease,21–23 substance abuse,24–26 and depression.27,28 Hydrogen peroxide is a reactive species generated as a byproduct of many biological processes, and its dysregulation has been associated with signaling apoptosis and the progression of Parkinson’s disease.29,30 Opioid neuropeptides have been shown to be directly linked to functions regulating feeding behavior,31 perception of pain,32 reward and reinforcement,33 and sedation.34 Further, these opioid neuropeptides, as well as opiates (semi-synthetic opioids) such as hydrocodone and heroin, are linked to substance abuse due to influences on the mesolimbic dopaminergic pathways that are associated with reward-based learning. Therefore, research focused on elucidating the diverse roles of neurotransmitters have a broad impact covering many scientific disciplines and society.

4

1.2 Methods for Monitoring Neurotransmission in Live Biological Preparations

1.2.1 Electrophysiology

Electrophysiology is the study of the electrical properties of biological cells and is particularly useful at electrically excitable cells such as neurons or muscle fibers.

Electrophysiology involves measuring voltage or current changes, generally associated with action potentials, on a variety of scales ranging from single cells to a local region of cells.35 With respect to neuroscience, electrophysiology measurements are either ‘field’ or ‘patch clamp’ type, and broadly speaking are used to study neuronal activity by monitoring cell membrane potentials, current flow across cell membranes, or action potential firing.10

Patch clamp measurements use a micropipette filled with electrolyte and an electrode to interface with a single cell or a portion of a cell membrane. In whole cell measurements, the micropipette is used to create an opening in the cell membrane to access the cytosol, enabling the measurement of cell membrane potentials or currents by either controlling the cell current or potential, respectively. These measurements are particularly useful in elucidating the effect of neurotransmitter-receptor binding on potentiating or inhibiting firing of action potentials responsible for neurotransmitter release.36 For example, ionotropic glutamate receptors are ligand- gated Na+ channels that allow influx of Na+ when bound. Whole-cell measurements can be used to determine glutamate induced changes to resting membrane potentials, as well as the frequency of action potential firing. Alternatively, the micropipette can be used to form a seal around (cell- attached patch) or extract a specific portion (inside-out patch) of a cell membrane for a more specific studies of individual ion channel or receptor function.

In field-type measurements, a microelectrode is placed in tissue and rests on a floating ground.37 Electrical activity generated by adjacent neurons can be monitored through changes in

5

the electrodes potential that can resemble action potentials; the number of neurons sampled is dependent on microelectrode dimensions. Field-type measurements are particularly useful for monitoring electrical activity of cells in behaving animals, and due to the relative simplicity of application (i.e. not having to interface with individual cells) it is commonly performed at microelectrode arrays to map spatial brain activity. For example, field electrophysiology measurements have been key to understanding how neurons in the primary visual cortex respond with high specificity for a particular stimulus.38,39

1.2.2 Microdialysis and Solid-Phase Extraction Sampling with Mass Spectrometric Detection

A large amount of information regarding chemical composition of various brain regions has been gained by extracting and processing, or sampling fluid from the extracellular space of specific brain regions. Microdialysis is common method used to sample fluid from the extracellular space of live animals. Using a microprobe and stereotaxic coordinates, precise targeting of specific brain regions can be achieved. The tip of the probe contains a semipermeable membrane, usually a molecular weight cutoff membrane, allowing transfer of molecules in the extracellular space into the perfusing dialysate.40,41 Dialysate can then be collected (generally >1min/sample) over time in order to measure molecular changes. The samples be subjected to further processing or can be used directly in analyses. Microdialysis has been fundamental to a large body of research associated with various behavior and pharmacology. For example, microdialysis has been used to monitor neuropeptide species fluctuations revealing a link with feeding behavior.42 Solid-phase extraction

(SPE) is another sampling method that extracts components of a sample through interactions with the stationary phase. SPE has been used in direct sampling of biological preparations. For example, a number of works utilize a SPE microprobe to sample the brain producing similar results to microdialysis sampling,43 and in others, individual solid-phase beads or capillaries were placed

6

above highly specific locations of a live brain slice to collect low-level neuropeptides.44,45

Microdialysis and SPE sampling has been useful in a variety of studies. However, the timescale of sample collection is not appropriate for monitoring subsecond neurotransmission. Furthermore, common microprobe dimensions result in tissue damage which can impact the function of surrounding tissue.46 Also, probes may not be appropriate for specific sampling of smaller brain regions.

Liquid chromatography is likely the most common methods used to further isolate specific molecules for detection. Mass spectrometry provides low-detection limits and unmatched chemical selectivity for powerful analyses. Matrix-assisted laser desorption/ionization (MALDI) and electrospray ionization (ESI), coupled with a number of mass analyzers, are most commonly employed for biological samples. MALDI utilizes laser pulses at a dried matrix of sample and photoadsorber, to vaporize and ionize the sample components which can then be accelerated into the mass analyzer. ESI utilizes electrodes to apply a high voltage to a liquid sample in a small capillary to drive formation of an aerosol of charged droplets. As solvent evaporates from the droplets, coulombic fissions result in ionized molecules that are directed towards the mass analyzer. Mass analyzers separate ions based on their mass-to-charge ratio using controlled electric and magnetic fields. For larger species, namely peptides and proteins, it is common to perform tandem MS (MS/MS) experiments. In MS/MS experiments a specific mass-to-charge species is isolated by the first mass analyzer and loaded into a special chamber where various chemical reactions can be performed that lead to fragmentation or dissociation. These reaction products are then directed towards a second mass analyzer to study the fragments. Reaction products are dependent on the starting reagent, thus MS/MS provides a means to identify compounds with similar or identical mass-to-charge ratios. In this way, distinct mass-to-charge ratios, isotopic

7

patterns, and fragmentation patterns can be utilized to identify a large diversity of distinct species in a given sample with high confidence.

1.2.3 Electrochemical Methods at Microelectrodes

The use of microelectrodes for biological measurements was largely pioneered by Adams and colleagues,47–49 and significantly advanced by Wightman and colleagues.37,50,51 Most biological electrochemistry measurements are performed at carbon-fiber microelectrodes, as they provide a number of advantages for in vivo measurements. First, dimensions of microelectrodes cause significantly less tissue damage than microdialysis probes and allow precise measurement of neurochemicals in discrete brain regions for high spatial resolution. The small dimensions also enhance mass transport to the sensor surface through both planar and radial diffusion, resulting in lower detection limits.52 Second, carbon has renewable surfaces leading to biofouling resistance and are more cost efficient than other common electrode materials (, , etc.). 53 Third, carbon surfaces can contain and be populated with oxygen-containing functional groups that benefit adsorption and transfer of certain neurochemicals.54 Fourth, the use of microelectrodes generates little current relative to measurements at macroelectrodes. This permits measurements to be made in a two-electrode setup without significant ohmic drop, as current passed does not polarize the reference/counter electrode. Another byproduct is that less electrodes need to be implanted for in vivo studies. Recent advances in materials fabrication have led to the popularization of carbon nanostructures, such as nanotubes, as standalone sensors.55,56 However, it is more common for these materials to be used to modify the surface of carbon fiber to enhance sensitivity and electron transfer kinetics.57,58

Two main classes of electrochemical measurements are utilized in biological experiments, in which the potential at an electrode surface is controlled to drive oxidation or reduction of many

8

important neurochemicals, namely dopamine, norepinephrine and serotonin. Electron transfer is monitored as current which provides information regarding the magnitude of the event(s).

Amperometry is a method where the electrode potential is set at a level to oxidize or reduce an analyte of interest. These measurements have high temporal resolution but have limited chemical selectivity, as any molecule that is redox-active at the applied potential may contribute to monitored current. Size or charge exclusion membranes can be utilized to enhance selectivity, but at the expense of response time.59 Amperometry measurements are particularly well suited for measurements of vesicular release at single cells in culture using microelectrodes. During a vesicle release event, neurotransmitters are released that are oxidized or reduced at the electrode surface.

The frequency, magnitude, and rise/decay characteristics of current spikes over time provides information regarding function and biophysics of neurochemical release.11,60,61 For example, amperometry was used to investigate the mechanism by which zinc modulates vesicular release of dopamine from PC12 cells.62

Background-subtracted fast-scan cyclic voltammetry (FSCV) uses a controlled potential waveform to drive oxidation and reduction at microelectrodes. This dissertation is focused on advancing applications of FSCV, and in depth methods and applications are provided in Chapter

2. Briefly, a typical measurement employs a triangular waveform ranging from -0.4 V to +1.3 V

(vs Ag/AgCl) applied at 400 V s-1, resulting in an 8.5 ms measurement that is generally applied with a frequency of 10-60 Hz; electrode held at -0.4 V between measurements to allow preconcentration of certain charged species.63 The use of rapid scan rates generates a significant background charging current (nonfaradaic) that is stable and can be subtracted from measurements in the presence of an analyte. This reveals potential vs current relationships, or cyclic voltammograms, for specific analytes which contain peaks associated with redox events. Analytes

9

can oxidize and reduce at different potentials with different electron transfer kinetics. Thus, cyclic voltammograms can have different peak position and shapes for different analytes giving FSCV an additional dimension of selectivity. FSCV is well suited for measurements of neurochemical fluctuations in live tissue or subjects and has provided invaluable insight regarding various molecular mechanisms that regulate behavior and CNS function.64,65 Thus, FSCV at microelectrodes offers spatiotemporal resolution and sensitivity appropriate for studying subsecond neurotransmission, while affording adequate chemical selectivity.

1.3 Critical Issues Associated with In Vivo Electrochemistry

1.3.1 Chemical Selectivity

While FSCV offers an additional dimension of selectivity compared to amperometry, selectivity is a major concern due to the chemical complexity of in vivo systems (pH 7.4). Three main strategies are employed to improve confidence in chemical identification: 1) tailoring the waveform for a specific analyte, 2) application of an electrode coating, and 3) knowledge regarding the neurotransmitter localization within certain brain regions.

For example, epinephrine, norepinephrine, dopamine, DOPAC (3,4- dihydroxyphenylacetic acid; a dopamine metabolite), and ascorbic acid have similar voltammetric signatures using a triangular waveform ranging from 0.0 V to +1.3 V.66 By extending the waveform to +1.45 V, epinephrine can easily be distinguished as it undergoes an additional oxidation at ~+1.4 V unobservable for the other molecules.67 Adjusting the lower waveform limit to -0.4 V reduces the sensitivity to DOPAC and ascorbic acid, which exist as conjugate bases at pH 7.4, and enhances sensitivity to the , which exist as conjugate acids at pH 7.4, through electrostatic forces. Additional charge-based manipulations can be achieved through the

10

use of ion-exchange coatings such as Nafion, a sulfonated tetrafluoroethylene, to enhance selectivity for the catecholamines.68 There are also applications of size-exclusion membranes that abolish or slow diffusion of larger species towards the electrode surface.69 Finally, a large body of literature demonstrate specific dopamine localization within the striatum and norepinephrine within the locus coeruleus. Thus, measurements in the striatum made with an appropriate waveform or coating are almost certainly attributed to dopamine.

1.3.2 Complex Data Analysis

Analysis of in vivo FSCV recordings is time-intensive. First, many in vivo recordings contain contributions from more than one analyte or interferent. Thus, statistical tools are employed, namely principal component analysis with inverse least-squares regression (PCR), to develop multivariate models enabling deconvolution of signal contributions.51,70 In PCR, a training set composed of analyte/interferent specific cyclic voltammograms covering relevant magnitudes for all contributors to the mixed signal must be constructed. These data undergo an orthogonal linear transformation to a new coordinate system, and principal component vectors are assigned to the maximal sources of variance in the training set, where each component generally describes an analyte or interferent. The mixed signal data can then be projected onto the principal components, and the score on each component can be used to inform about individual contributions. PCR performance is determined by the quality of the training set. However, construction of training sets using in vivo data is complicated by difficulties associated with obtaining independent cyclic voltammograms for each analyte or interferent. Thus, recordings must be manually inspected to ensure proper selection of training set cyclic voltammograms.

Second, the background signal undergoes changes over time which results in background subtraction artifacts called electrochemical drift. Drift is associated with continual renewal and

11

etching of the carbon surface, changes in impedance, as well as fouling and changes to an electrodes capacitance.71,72 Thus, its signal features are highly diverse, can span the entire potential range, and can change in each recording or by changing the background position. Therefore, PCR is generally ineffective in deconvolution of signals containing drift. This has limited the use of

FSCV to monitor molecular fluctuations within 30-90 seconds of the background position. For extended recordings the background position must be placed throughout the recording which can result in missed neurochemical events.

1.3.3 Quantification of Measurements

An advantage of FSCV at CFME is that electrodes can be permanently implanted to make measurements over weeks to months in behaving animals. It has been established that acute implantation (minutes to hours) of CFMEs change the electrode surface, resulting in decreased sensitivity and shifts of cyclic voltammogram features.73–76 Acutely implanted sensors can be removed and post-calibrated in a flow system providing some control for performance changes that result from tissue exposure. However, permanently implanted sensors are extremely difficult to extract from tissue without damage. Paired with the difficulty in selecting in vivo cyclic voltammograms for constructing training sets, it has become common for these recordings to be analyzed using sensitivities and cyclic voltammograms recorded in vitro.76 Mismatch between the shape of in vivo and in vitro voltammogram features result in multivariate model failure and/or incorrect assignment of signal variance to analytes or interferents. Furthermore, the use of in vitro calibration factors for in vivo recordings likely result in misestimation of analyte concentration.

Although these issues persist, advances in methods to validate the effectiveness of PCR models have increased the confidence in results drawn from multivariate analysis.51,77

12

1.4 Electrochemical Impedance Spectroscopy

Electrochemical impedance spectroscopy (EIS) is a powerful and non-destructive electrochemical method that utilizes a small applied potential or current perturbation to characterize an electrochemical system or interface. The potential or current perturbation is small to minimize significant chemical or structural changes to a system, and to reduce relevant mathematical relationships to linear forms.78 This perturbation is performed across a range of frequencies (typically Hz to MHz) to characterize the frequency dependent impedance of a system.

The quotient of the voltage and current amplitudes provides the total impedance of a system at a given frequency, based on Ohm’s law. The time by which the current precedes or lags the voltage is measured as a phase angle which is reported in radians or degrees; for a pure resistor the voltage and current are in phase. Total impedance and phase angle are easily determined, and can be used to calculate the real and imaginary components of impedance. Data are generally displayed as

Bode plots, which display an impedance property as a function of frequency, and Nyquist plots, which displays the relative magnitude of real and imaginary components. The real component is indicative of the resistance, whereas the imaginary component describes the reactance of a system, which is related to inductance or capacitance. Reactance as a function of frequency can then be used to determine either the capacitance and/or inductance of a system.

A common type of analysis used in EIS studies involves equivalent circuit modeling.

Equivalent circuit modeling uses combinations of circuit elements to model the impedance response of real systems, in which each element in the circuit model represents a physical property of the electrochemical system.52,78 Algorithms are used to optimize circuit element parameters

(magnitudes) to best fit the observed impedance data. Ideal models are then selected based on the ability to assign circuit elements to physical properties and the error between the fit and real data,

13

where Occam’s razor is employed; i.e., if adding a circuit element does not significantly improve the fit, then it is discarded. The equivalent model can then be used to study the effect of system perturbations on specific physical properties. EIS and equivalent circuit modeling are highly utilized to study processes implicated in science,79 coatings,80 generation of renewable energy,81 energy storage,82 surface science, and sensing.83

1.5 Dissertation Overview

This dissertation consists of seven additional chapters that focus on further developing the application of FSCV at CFMEs for in vivo measurements. However, the developed methodologies should be useful in many electrochemical molecular sensing applications. Chapter 2 provides a detailed description of FSCV application for in vivo molecular monitoring in animals. It also provides a brief review of current methodologies and advances to enhance chemical selectivity and sensitivity, data analysis, and throughput, as well as a discussion of some limitations and caveats associated with FSCV use. Chapter 3 chronicles the characterization of a modified sawhorse waveform (MSW) developed for reliable and selective detection of neuropeptides and amino acids. The MSW parameters’ impact on resulting measurements are investigated in order to optimize the approach. Further, the first known measurements of subsecond neuropeptide measurements in live tissue are reported. Chapter 4 focuses on FSCV method development for enhanced chemical selectivity for endogenous hydrogen peroxide in vivo. By coupling a ‘double’ voltammetric waveform with partial least squares regression (DW-PLSR), a predictive model can be constructed to characterize and remove interfering signals that traditional FSCV analyses struggle to handle. Chapter 5 describes the application of DW-PLSR to remove electrochemical drift. The FSCV background signal slowly changes over time, leading to background-subtraction

14

artifacts termed ‘electrode drift’ that limits analysis to 30-90 second segments before a new background position must be selected. Using DW-PLSR this time window can be increased to >10 minutes, increasing throughput and enabling evaluation of more gradual changes in analyte concentration(s). Chapter 6 describes the effect of differential impedance on FSCV measurements and analyses. In short, in vivo measurements are made in chemically and physically diverse environments that result in matrix effects and fouling of the sensor. This leads to changes in sensor impedance, distortion of cyclic voltammogram shape/position, and changes to sensitivity, which if unaccounted for, can lead to significant error in analyses. Chapter 7 covers the use of electrochemical impedance spectroscopy (EIS) to develop equivalent circuit models for CFMEs.

EIS and the models were used to investigate the effect of electrode fabrication method, carbon fiber material, electrochemical carbon activation, and tissue exposure on the electrode/solution interface, impedance, and FSCV performance. Chapter 8 features the development of a multiplexed FSCV and EIS measurement paradigm. This method enables real time tracking of

CFME impedance/capacitance that are directly correlated with electrochemical performance, while maintaining the advantageous temporal resolution of FSCV.

15

1.6 References

1. Mai, J. K.; Paxinos, G. The Human Nervous System; Academic Press, 2011.

2. Brodal, P. The Central Nervous System: Structure and Function; Oxford University Press, 2004.

3. Khakh, B. S.; Sofroniew, M. V. Diversity of Astrocyte Functions and Phenotypes in Neural Circuits. Nat. Neurosci. 2015, 18 (7), 942–952.

4. Nedergaard, M. Direct Signaling from Astrocytes to Neurons in Cultures of Mammalian Brain Cells. Science 1994, 263 (5154), 1768.

5. Henn, F. A.; Haljamae, H.; Hamberger, A. Glial Cell Function: Active Control of Extracellular K+ Concentration. Brain Res. 1972, 43 (2), 437–443.

6. Araque, A.; Navarrete, M. Glial Cells in Neuronal Network Function. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365 (1551), 2375–2381.

7. Jackson, F. R. Glial Cell Modulation of Circadian Rhythms. Glia 2011, 59 (9), 1341–1350.

8. Zonta, M.; Angulo, M. C.; Gobbo, S.; Rosengarten, B.; Hossmann, K.-A.; Pozzan, T.; Carmignoto, G. Neuron-to-Astrocyte Signaling Is Central to the Dynamic Control of Brain Microcirculation. Nat. Neurosci. 2003, 6 (1), 43–50.

9. Wright, S. H. Generation of Resting Membrane Potential. Adv. Physiol. Educ. 2004, 28 (4), 139–142.

10. Bear, M. F.; Connors, B. W.; Paradiso, M. A. Neuroscience; Lippincott Williams & Wilkins, 2007; Vol. 2.

11. Wightman, R. M.; Haynes, C. L. Synaptic Vesicles Really Do Kiss and Run. Nat. Neurosci. 2004, 7 (4), 321–322.

12. Rao, T. C.; Passmore, D. R.; Peleman, A. R.; Das, M.; Chapman, E. R.; Anantharam, A. Distinct Fusion Properties of Synaptotagmin-1 and Synaptotagmin-7 Bearing Dense Core Granules. Mol. Biol. Cell 2014, 25 (16), 2416–2427.

13. Rao, T. C.; Santana Rodriguez, Z.; Bradberry, M. M.; Ranski, A. H.; Dahl, P. J.; Schmidtke, M. W.; Jenkins, P. M.; Axelrod, D.; Chapman, E. R.; Giovannucci, D. R.; et al. Synaptotagmin Isoforms Confer Distinct Activation Kinetics and Dynamics to Chromaffin Cell Granules. J. Gen. Physiol. 2017, 149 (8), 763–780.

14. Hu, R.; Ren, B.; Lin, C.-J.; Oleinick, A.; Svir, I.; Tian, Z.-Q.; Amatore, C. How “Full” Is “Full Fusion” during Exocytosis from Dense Core Vesicles? Effect of SDS on “Quantal” Release and Final Fusion Pore Size. J. Electrochem. Soc. 2016, 163 (9), H853–H865.

16

15. Cárdenas, A. M.; Marengo, F. D. How the Stimulus Defines the Dynamics of Vesicle Pool Recruitment, Fusion Mode, and Vesicle Recycling in Neuroendocrine Cells. J. Neurochem. 2016, 137 (6), 867–879.

16. Lee, C. R.; Patel, J. C.; O’Neill, B.; Rice, M. E. Inhibitory and Excitatory Neuromodulation by Hydrogen Peroxide: Translating Energetics to Information. J. Physiol. 2015, 593 (16), 3431–3446.

17. Salio, C.; Lossi, L.; Ferrini, F.; Merighi, A. Neuropeptides as Synaptic Transmitters. Cell Tissue Res. 2006, 326 (2), 583–598.

18. Remington, G. Alterations of Dopamine and Serotonin Transmission in Schizophrenia. In Progress in Brain Research; Di Giovann, G., Di Matteo, V., Esposito, E., Eds.; Elsevier, 2008; Vol. 172, pp 117–140.

19. Mita, T.; Hanada, S.; Nishino, N.; Kuno, T.; Nakai, H.; Yamadori, T.; Mizoi, Y.; Tanaka, C. Decreased Serotonin S2 and Increased Dopamine D2 Receptors in Chronic Schizophrenics. Biol. Psychiatry 1986, 21 (14), 1407–1414.

20. Meltzer, H. Y. Clinical Studies on the Mechanism of Action of Clozapine: The Dopamine- Serotonin Hypothesis of Schizophrenia. (Berl.) 1989, 99 (1), S18–S27.

21. Eskow Jaunarajs, K. L.; George, J. A.; Bishop, C. L-DOPA-Induced Dysregulation of Extrastriatal Dopamine and Serotonin and Affective Symptoms in a Bilateral Rat Model of Parkinson’s Disease. Neuroscience 2012, 218, 243–256.

22. Scatton, B.; Javoy-Agid, F.; Rouquier, L.; Dubois, B.; Agid, Y. Reduction of Cortical Dopamine, Noradrenaline, Serotonin and Their Metabolites in Parkinson’s Disease. Brain Res. 1983, 275 (2), 321–328.

23. Barone, P. Neurotransmission in Parkinson’s Disease: Beyond Dopamine. Eur. J. Neurol. 2010, 17 (3), 364–376.

24. Rocha, B. A.; Fumagalli, F.; Gainetdinov, R. R.; Jones, S. R.; Ator, R.; Giros, B.; Miller, G. W.; Caron, M. G. Cocaine Self-Administration in Dopamine-Transporter Knockout Mice. Nat. Neurosci. 1998, 1 (2), 132–137.

25. Ritz, M.; Lamb, R.; Goldberg; Kuhar, M. Cocaine Receptors on Dopamine Transporters Are Related to Self-Administration of Cocaine. Science 1987, 237 (4819), 1219.

26. Parsons, L. H.; Justice Jr., J. B. Serotonin and Dopamine Sensitization in the Nucleus Accumbens, Ventral Tegmental Area, and Dorsal Raphe Nucleus Following Repeated Cocaine Administration. J. Neurochem. 1993, 61 (5), 1611–1619.

27. Kalia, M. Neurobiological Basis of Depression: An Update. Depress. Genet. Pathophysiol. Clin. Manif. 2005, 54 (5, Supplement), 24–27.

17

28. Nestler, E. J.; Carlezon, W. A. The Mesolimbic Dopamine Reward Circuit in Depression. Biol. Psychiatry 2006, 59 (12), 1151–1159.

29. Halliwell, B. Oxidative Stress and Neurodegeneration: Where Are We Now? J. Neurochem. 2006, 97 (6), 1634–1658.

30. Anantharam, V.; Lehrmann, E.; Kanthasamy, A.; Yang, Y.; Banerjee, P.; Becker, K. G.; Freed, W. J.; Kanthasamy, A. G. Microarray Analysis of Oxidative Stress Regulated Genes in Mesencephalic Dopaminergic Neuronal Cells: Relevance to Oxidative Damage in Parkinson’s Disease. Neurochem. Int. 2007, 50 (6), 834–847.

31. Simone, D. A.; Bodnar, R. J.; Goldman, E. J.; Pasternak, G. W. Involvement of Opioid Receptor Subtypes in Rat Feeding Behavior. Life Sci. 1985, 36 (9), 829–833.

32. Inturrisi, C. E. Clinical Pharmacology of Opioids for Pain. Clin. J. Pain 2002, 18 (4), S3–S13.

33. Le Merrer, J.; Becker, J. A. J.; Befort, K.; Kieffer, B. L. Reward Processing by the Opioid System in the Brain. Physiol. Rev. 2009, 89 (4), 1379–1412.

34. Gross, J. B.; Bailey, P. L.; Connis, R. T.; Coté, C. J.; Davis, F.; Epstein, B.; Gilbertson, L.; Nickinovich, D.; Zerwas, J. Practice Guidelines for Sedation and Analgesia by Non- Anesthesiologists. Anesthesiology 2002, 96 (4), 1004–1017.

35. Hodgkin, A. L.; Huxley, A. F. Action Potentials Recorded from Inside a Nerve Fibre. Nature 1939, 144 (3651), 710–711.

36. Borgland, S. L.; Malenka, R. C.; Bonci, A. Acute and Chronic Cocaine-Induced Potentiation of Synaptic Strength in the Ventral Tegmental Area: Electrophysiological and Behavioral Correlates in Individual Rats. J. Neurosci. 2004, 24 (34), 7482.

37. Takmakov, P.; McKinney, C. J.; Carelli, R. M.; Wightman, R. M. Instrumentation for Fast- Scan Cyclic Voltammetry Combined with Electrophysiology for Behavioral Experiments in Freely Moving Animals. Rev. Sci. Instrum. 2011, 82 (7), 074302.

38. Allison, T.; Begleiter, A.; McCarthy, G.; Roessler, E.; Nobre, A. C.; Spencer, D. D. Electrophysiological Studies of Color Processing in Human Visual Cortex. Electroencephalogr. Clin. Neurophysiol. Potentials Sect. 1993, 88 (5), 343–355.

39. Levi, D. M.; Manny, R. E.; Klein, S. A.; Steinman, S. B. Electrophysiological Correlates of Hyperacuity in the Human Visual Cortex. Nature 1983, 306 (5942), 468–470.

40. Schultz, K. N.; Kennedy, R. T. Time-Resolved Microdialysis for In Vivo Neurochemical Measurements and Other Applications. Annu. Rev. Anal. Chem. 2008, 1 (1), 627–661.

41. Perry, M.; Li, Q.; Kennedy, R. T. Review of Recent Advances in Analytical Techniques for the Determination of Neurotransmitters. Anal. Chim. Acta 2009, 653 (1), 1–22.

18

42. DiFeliceantonio, A. G.; Mabrouk, O. S.; Kennedy, R. T.; Berridge, K. C. Enkephalin Surges in Dorsal Neostriatum as a Signal to Eat. Curr. Biol. 2012, 22 (20), 1918–1924.

43. Cudjoe, E.; Bojko, B.; de Lannoy, I.; Saldivia, V.; Pawliszyn, J. Solid-Phase Microextraction: A Complementary In Vivo Sampling Method to Microdialysis. Angew. Chem. Int. Ed. 2013, 52 (46), 12124–12126.

44. Hatcher, N. G.; Richmond, T. A.; Rubakhin, S. S.; Sweedler, J. V. Monitoring Activity- Dependent Peptide Release from the CNS Using Single-Bead Solid-Phase Extraction and MALDI TOF MS Detection. Anal. Chem. 2005, 77 (6), 1580–1587.

45. Hatcher, N. G.; Atkins, N.; Annangudi, S. P.; Forbes, A. J.; Kelleher, N. L.; Gillette, M. U.; Sweedler, J. V. Mass Spectrometry-Based Discovery of Circadian Peptides. Proc. Natl. Acad. Sci. 2008, 105 (34), 12527.

46. Jaquins-Gerstl, A.; Michael, A. C. Comparison of the Brain Penetration Injury Associated with Microdialysis and Voltammetry. J. Neurosci. Methods 2009, 183 (2), 127–135.

47. Gerhardt, G. A.; Oke, A. F.; Nagy, G.; Moghaddam, B.; Adams, R. N. Nafion-Coated Electrodes with High Selectivity for CNS Electrochemistry. Brain Res. 1984, 290 (2), 390– 395.

48. Kissinger, P. T.; Hart, J. B.; Adams, R. N. Voltammetry in Brain Tissue — a New Neurophysiological Measurement. Brain Res. 1973, 55 (1), 209–213.

49. Adams, R. N. In Vivo Electrochemical Measurements in the CNS. Prog. Neurobiol. 1990, 35 (4), 297–311.

50. Heien, M. L.; Phillips, P. E.; Stuber, G. D.; Seipel, A. T.; Wightman, R. M. Overoxidation of Carbon-Fiber Microelectrodes Enhances Dopamine Adsorption and Increases Sensitivity. Analyst 2003, 128 (12), 1413–1419.

51. Keithley, R. B.; Mark Wightman, R.; Heien, M. L. Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis. TrAC Trends Anal. Chem. 2009, 28 (9), 1127–1136.

52. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

53. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal. Chem. 2010, 82 (5), 2020–2028.

54. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir 2010, 26 (11), 9116– 9122.

19

55. Schmidt, A. C.; Wang, X.; Zhu, Y.; Sombers, L. A. Carbon Nanotube Yarn Electrodes for Enhanced Detection of Neurotransmitter Dynamics in Live Brain Tissue. ACS Nano 2013, 7 (9), 7864–7873.

56. Wang, J.; Deo, R. P.; Poulin, P.; Mangey, M. Carbon Nanotube Fiber Microelectrodes. J. Am. Chem. Soc. 2003, 125 (48), 14706–14707.

57. Swamy, B. K.; Venton, B. J. Carbon Nanotube-Modified Microelectrodes for Simultaneous Detection of Dopamine and Serotonin in Vivo. Analyst 2007, 132 (9), 876–884.

58. Puthongkham, P.; Yang, C.; Venton, B. J. Carbon Nanohorn-Modified Carbon Fiber Microelectrodes for Dopamine Detection. Electroanalysis 2018, 30 (6), 1073–1081.

59. Wang, J. Permselective Coatings for Amperometric Biosensing. In Biosensors and Chemical Sensors; ACS Symposium Series; American Chemical Society, 1992; Vol. 487, pp 125–132.

60. Mellander, L. J.; Trouillon, R.; Svensson, M. I.; Ewing, A. G. Amperometric Post Spike Feet Reveal Most Exocytosis Is via Extended Kiss-and-Run Fusion. Sci. Rep. 2012, 2 (1), 907.

61. Amatore, C.; Arbault, S.; Bonifas, I.; Guille, M. Quantitative Investigations of Amperometric Spike Feet Suggest Different Controlling Factors of the Fusion Pore in Exocytosis at Chromaffin Cells. Biophys. Chem. 2009, 143 (3), 124–131.

62. Ren, L.; Pour, M. D.; Majdi, S.; Li, X.; Malmberg, P.; Ewing, A. G. Zinc Regulates Chemical- Transmitter Storage in Nanometer Vesicles and Exocytosis Dynamics as Measured by Amperometry. Angew. Chem. Int. Ed. 2017, 56 (18), 4970–4975.

63. Bath, B. D.; Michael, D. J.; Trafton, B. J.; Joseph, J. D.; Runnels, P. L.; Wightman, R. M. Subsecond Adsorption and Desorption of Dopamine at Carbon-Fiber Microelectrodes. Anal. Chem. 2000, 72 (24), 5994–6002.

64. Flagel, S. B.; Clark, J. J.; Robinson, T. E.; Mayo, L.; Czuj, A.; Willuhn, I.; Akers, C. A.; Clinton, S. M.; Phillips, P. E. M.; Akil, H. A Selective Role for Dopamine in Stimulus–Reward Learning. Nature 2011, 469 (7328), 53–57.

65. Phillips, P. E. M.; Stuber, G. D.; Heien, M. L. A. V.; Wightman, R. M.; Carelli, R. M. Subsecond Dopamine Release Promotes Cocaine Seeking. Nature 2003, 422 (6932), 614–618.

66. Schmidt, A. C.; Dunaway, L. E.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Multiple Scan Rate Voltammetry for Selective Quantification of Real-Time Enkephalin Dynamics. Anal. Chem. 2014, 86 (15), 7806–7812.

67. Pihel, Karin.; Schroeder, T. J.; Wightman, R. Mark. Rapid and Selective Cyclic Voltammetric Measurements of Epinephrine and Norepinephrine as a Method To Measure Secretion from Single Bovine Adrenal Medullary Cells. Anal. Chem. 1994, 66 (24), 4532–4537.

20

68. Vreeland, R. F.; Atcherley, C. W.; Russell, W. S.; Xie, J. Y.; Lu, D.; Laude, N. D.; Porreca, F.; Heien, M. L. Biocompatible PEDOT: Nafion Composite Electrode Coatings for Selective Detection of Neurotransmitters in Vivo. Anal. Chem. 2015, 87 (5), 2600–2607.

69. Wilson, L. R.; Panda, S.; Schmidt, A. C.; Sombers, L. A. Selective and Mechanically Robust Sensors for Electrochemical Measurements of Real-Time Hydrogen Peroxide Dynamics in Vivo. Anal. Chem. 2018, 90 (1), 888–895.

70. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E. M.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234.

71. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal. Chem. 2017, 89 (11), 6166–6174.

72. Meunier, C. J.; McCarty, G. S.; Sombers, L. A. Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double-Waveform Partial-Least-Squares Regression. Anal. Chem. 2019, 91 (11), 7319–7327.

73. Meunier, C. J.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Background Signal as an in Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2017, 8 (2), 411–419.

74. Roberts, J. G.; Toups, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode Calibration Strategy for Voltammetric Measurements In Vivo. Anal. Chem. 2013, 85 (23), 11568–11575.

75. Logman, M. J.; Budygin, E. A.; Gainetdinov, R. R.; Wightman, R. M. Quantitation of in Vivo Measurements with Carbon Fiber Microelectrodes. J. Neurosci. Methods 2000, 95 (2), 95–102.

76. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2016, 7 (3), 349–359.

77. Keithley, R. B.; Wightman, R. M. Assessing Principal Component Regression Prediction of Neurochemicals Detected with Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2011, 2 (9), 514–525.

78. Orazem, M. E.; Tribollet, B. Electrochemical Impedance Spectroscopy; John Wiley & Sons, 2017.

79. Jüttner, K. Electrochemical Impedance Spectroscopy (EIS) of Corrosion Processes on Inhomogeneous Surfaces. Electrochimica Acta 1990, 35 (10), 1501–1508.

80. Mansfeld, F. Use of Electrochemical Impedance Spectroscopy for the Study of Corrosion Protection by Polymer Coatings. J. Appl. Electrochem. 1995, 25 (3), 187–202.

21

81. Adachi, M.; Sakamoto, M.; Jiu, J.; Ogata, Y.; Isoda, S. Determination of Parameters of Electron Transport in Dye-Sensitized Solar Cells Using Electrochemical Impedance Spectroscopy. J. Phys. Chem. B 2006, 110 (28), 13872–13880.

82. Cañas, N. A.; Hirose, K.; Pascucci, B.; Wagner, N.; Friedrich, K. A.; Hiesgen, R. Investigations of Lithium–Sulfur Batteries Using Electrochemical Impedance Spectroscopy. Electrochimica Acta 2013, 97, 42–51.

83. Park, J.-Y.; Park, S.-M. DNA Hybridization Sensors Based on Electrochemical Impedance Spectroscopy as a Detection Tool. Sensors 2009, 9 (12).

22

CHAPTER 2

Fast-Scan Voltammetry for In vivo Measurements of Neurochemical Dynamics

The work was completed in collaboration with Leslie A. Sombers, and has been submitted as a chapter in The Brain Reward System published in the Neuromethods series by Springer Nature.

2.1 Introduction

Electrochemical measurements permit in situ detection of neurochemicals with high sampling rates (µs-ms) when coupled with microelectrodes.1,2 These probes offer high spatial resolution and exact minimal tissue damage.3 Electrochemical methods used in vivo generally involve the measurement of current flow in response to application of a controlled potential, and can be divided into two groups: voltammetric and amperometric methods. Background-subtracted fast-scan cyclic voltammetry (FSCV) is well suited for in vivo studies, as it affords the best balance of temporal resolution, sensitivity, and chemical selectivity. For example, FSCV has been used to monitor naturally occurring dopamine transients associated with reinforcement learning,4–10 as well as in the initiation and modulation of self-paced movements.11,12 FSCV has also been used to monitor serotonin dynamics following treatment selective serotonin reuptake inhibitors

(SSRIs).13,14 One study revealed that SSRIs potentiated serotonin release in pair-housed animals, but not those that were socially isolated.15 Furthermore, FSCV can paired with other techniques such as microinjection, iontophoresis, and electrophysiology that can provide additional information regarding specific neurotransmitter function(s).5,16–18

23

Figure 2.1. Introduction to FSCV. A) Typical FSCV potential waveform applied to a carbon fiber working electrode. B) This generates a predominately non-faradaic background current (gray) that grows with repeated waveform application (black). The red trace is an example of the signal in the presence of dopamine. C) The background can be subtracted to reveal a cyclic voltammogram for the oxidation of dopamine (DA) to dopamine-σ-quinone (DOQ), and reduction of DOQ to DA. D) The amplitude of the current is directly related to dopamine concentration. E) Cyclic voltammograms can be converted to linear voltammograms, and plotted in three dimensions. F) This information can be plotted in 2-D to quickly visualize large amounts of data where current is plotted in false color. A vertical line through the color plot can be used to reconstruct the cyclic voltammogram for analyte identification. G) A horizontal line through the color plot can be used to report on current at an analyte’s redox potential over time; i.e., chemical dynamics. H) Representative scanning electron micrographs for carbon-fiber microelectrodes insulated in borosilicate glass (top) and fused silica (right).

2.1.1 Use of FSCV in Analyte Detection

Carbon-fiber microelectrodes (CFMEs) are most commonly used for neurochemical measurements, as this carbon material offers an inherent resistance to biofouling relative to other common electrode materials, such as gold and platinum.19–21 The carbon fiber (5-10 µm diameter) is typically insulated using a borosilicate-glass or fused-silica capillary, and the protruding carbon fiber is cut to a defined length (typically 75-125 µm).1,22 These cylindrically shaped carbon-fiber microelectrodes (CFMEs) are >40x smaller in each dimension than typical microdialysis probes,

24

permitting them to be stereotaxically implanted in a discrete brain region with significantly less tissue damage. In fact, CFMEs are particularly well suited to probe brain regions that have gradations in the density of neuronal terminals over these dimensions.23,24 The high aspect ratio and surface roughness provide a large surface area for sensing. Currents generated during in vivo measurements are typically on the order of 10s-100s of nanoamperes. This magnitude of current does not result in significant polarization of the (Ag/AgCl).25 Thus, reliable measurements are made without implantation of an auxiliary (counter) electrode.

In FSCV, a dynamic potential waveform is controlled at the CFME to drive into, and out of, molecules that are located in the vicinity of that electrode (Figure 1). Molecules that are capable of heterogeneous electron transfer have molecular orbitals that either contain, or are devoid of, electrons. As the potential is scanned through the waveform, the energy level of the

CFME shifts with respect to that of the molecular orbitals of the analyte and an electron has a probability to be extracted from, or injected into, that molecular orbital. This is termed ‘oxidation’ and ‘reduction’, respectively. The movement of charge is measured in the form of current

(charge/time), which is plotted versus the applied potential, resulting in a cyclic voltammogram

(CV). Importantly, the positions of these energy levels are specific to a given molecule. Thus, the kinetics associated with the heterogeneous electron transfer result in CVs that can be used as a

‘signature’ for a particular analyte, allowing for putative identification.

Dopamine is the neurotransmitter that is most often monitored using FSCV. Its characteristic CV (Figure 1) is used to distinguish it from a number of interferents, such as ascorbic acid and shifts in pH. Advances in electrode materials,26–29 voltammetric strategies,30–33 and data analysis strategies have increased the number of neurochemicals targeted using FSCV.34–37

Currently, FSCV is used to monitor catecholamines (dopamine, norepinephrine, epinephrine),38,39

25

indolamines (serotonin and ),40,41 hydrogen peroxide,42 oxygen,43 peptides,31 and purines

(adenosine and guanosine) in vivo.30,44 Although CVs serve to putatively identify a given analyte, selectivity is always a concern. For example, the CVs for dopamine and norepinephrine are virtually indistinguishable, and epinephrine can only be distinguished from these by sweeping to more positive potentials (+1.4 V).45 Therefore, positive identification of a specific species must be electrochemically, anatomically, physiologically, and pharmacologically verified.46 For example, to validate dopamine measurements using FSCV:

1. The in vivo signal should be consistent with a CV recorded using a standard collected in

vitro.

2. The signal must have been collected in an anatomical region known to contain dopamine

terminals.

3. Stimulation of cell bodies should illicit dopamine release.

4. Pharmacological manipulation should modulate release and/or reuptake in a predictable

manner, consistent with the literature.

When voltammetry is used to identify a molecule that has not been previously characterized using

FSCV, an independent strategy should be used to validate the measurement (whenever possible).46

2.2 Materials

2.2.1 Electrode Fabrication

1. Carbon fiber: T-650, polyacrylonitrile-based fibers are commonly used (Cytec Industries,

Woodland Park, NJ)

2. Borosilicate-glass capillary (1.0 mm O.D., 0.5 mm I.D.; A-M Systems, Sequim, WA),

typically used for measurements in anesthetized animals.

26

3. Fused-silica tubing (164.7 µm O.D., 98.6 µm I.D.; Polymicro Technologies, Phoenix, AZ),

typically used for experiments in awake animals.

4. Vertical electrode puller: PE-21 (Narishige, Tokyo, Japan)

5. Vacuum pump

6. Stereo microscope

7. Glass microscope slide

8. Surgical scalpel

9. Rubber/silicone putty

10. Silver paint: Silver Print II (GC Electronics, Rockford, IL)

11. Silver wire: 0.5 mm diameter for Ag/AgCl reference electrode (Sigma-Aldrich, St. Louis,

IL)

12. Insulated steel wire leads (Squires Electronics, Inc., Cornelius, OR)

13. Gold connectors: PCB socket (Newark Electronics, Chicago, IL)

14. Fast-hardening epoxy (McMaster Carr, Atlanta, GA)

15. Conductive silver epoxy (MG Chemical, Thief River Falls, MN)

16. Liquid electrical tape (GC Electronics, Rockford, IL)

2.2.2 Surgery

1. Anesthetic: ketamine/xylazine, urethane, or isoflurane (general anesthesia). Also

bupivacaine (0.25%; local anesthetic)

a. Use of isoflurane requires additional equipment including an isoflurane vaporizer,

oxygen gas, induction chamber, and an anesthesia mask that is compatible with the

stereotaxic instrument.

2. Heating pad(s)

27

3. Stereotaxic instrument: such as Model 900 Small Animal Stereotaxic Instrument (David

Kopf Instruments, Tujunga, CA)

4. Guide cannula for reference electrode (BASi, West Lafayette, IN)

5. Surgical anchor screws (Gexpro, Indianapolis, IN)

6. Cranioplastic grip cement (Dentsply International, Inc., Milford, DE)

7. Surgical tools and materials: scissors, tweezers, gauze, cotton swabs, Betadine®, Virkon,

screwdriver

8. Dremmel® rotary tool and small diameter drill bit (consistent with surgical screws)

2.2.3 Electrochemistry

1. Multifunction input/output data acquisition card: PCIe-6363 (National Instruments, Austin,

TX)

2. Software for data collection and analysis: such as High Definition Cyclic Voltammetry

(University of North Carolina at Chapel Hill, Department of Chemistry, Electronics

Facility) or Fast Scan Cyclic Voltammetry Software (Pinnacle Technology Inc., Lawrence,

KS)

3. Potentiostat. One of the following systems is appropriate: Universal Electrochemistry

Instrument (UEI, UNC-Chapel Hill, Department of Chemistry, Electronics Facility),

WaveNeuro FSCV System (Pine Research Instrumentation, Durham, NC), or Fast Scan

Cyclic Voltammetry System (Pinnacle Technology Inc., Lawrence, KS)

4. Headstage: miniaturized current-to-voltage converter. Can be purchased from UNC-

Chapel Hill (Department of Chemistry, Electronics Facility) or Pine Research

Instrumentation.

28

5. Commutator: 25 channels (Crist Instruments, Hagerstown, MD). Allows electrical

connectivity to a freely moving animal

6. Screened behavior chamber, grounded: custom made for in vivo experiments (Med

Associates, Inc., St. Albans, VT). Screening serves as a Faraday cage.

2.2.4 Electrode Calibration

1. Buffered electrolyte: phosphate-buffered saline, TRIS-buffered saline, or artificial cerebral

spinal fluid.

2. Grounded Faraday cage: custom built in house

3. Flow-injection apparatus: six-port, two-position high-performance liquid chromatography

valve, air actuator, and digital valve interface (VICI, Houston, TX)

4. Custom molded flow chamber

5. Micromanipulator, for electrode placement.

6. Analyte standards: prepared fresh or from frozen aliquots. An example for dopamine HCl:

5 mM in 0.1 M perchloric acid, stored at -20 °C. Dilutions made in buffer.

7. A commercial flow cell that uses a gravity-feed system is available from Pine Research

Instrumentation (Durham, NC).

2.3 Methods

2.3.1 Electrochemistry, Instrumentation, and Software

2.3.1.1 Introduction

When a potential waveform is applied to a CFME, two sources of current are generated:

(1) Non-faradaic current (or charging current) results from the movement of charged species in close vicinity to the electrode/solution interface, and (2) faradaic current is generated through the

29

transfer of electrons into, and out of, specific molecules at the electrode surface. The non-faradaic current results largely from rapid charging of the electrochemical double-layer at the electrode surface. It increases linearly with the scan rate, and is the dominant source of current generated using FSCV. By contrast, faradaic current is generated: (1) with each waveform application, as redox-active functional groups on the carbon surface are oxidized and reduced47, and (2) through the oxidation and reduction of solution phase analytes at the electrode surface. The overall current profile is specific to a given electrode. In the absence of a chemical change at the electrode surface, the signal is stable over tens of seconds. This ‘background’ signal can be subtracted from a signal recorded in the presence of an electroactive neurochemical to reveal a characteristic CV (Figure

1). The overall shape of the voltammogram reveals information about analyte identity, reversibility, and electron transfer kinetics,25 and the magnitude of the peaks is indicative of the concentration of the analyte at the electrode surface. It should be noted that background subtraction limits the utility of FSCV to monitoring analyte fluctuations when in vivo, as basal levels of neurotransmitter that are stable over tens of seconds are subtracted in the background signal.

CFMEs typically generate hundreds of nanoamperes to single-digit microamperes of background current with each scan, dependent on the surface area of the electrode. By contrast, the faradaic current generated in response to fluctuating neurochemicals encountered in vivo is generally on the order of hundreds of picoamperes to tens of nanoamperes. The large difference in the scale of the background signal as compared to that generated in oxidation or reduction of a species of interest requires the use of specialized instrumentation to reduce contamination of the resultant signal by noise. In addition, all measurements are performed in a grounded faraday cage.

Instrumentation used for FSCV is often custom-made; however, some commercial systems are available (as described in the Methods). High-Definition Cyclic Voltammetry (HDCV) is probably

30

the software that is most often used to control waveform output and to process the collected signals, in conjunction with data acquisition cards that are used to interface the software with the hardware.48 Briefly, the instrumentation generally consists of an analog low-pass filter and a head- amp module, and the ‘headstage’ consists of relevant electrode leads as well as an op-amp that serves as a current-voltage converter (Figure 6).49 The computer-generated potential waveform contains digitization noise that is smoothed using a low-pass filter prior to reaching the CFME.

The current generated at the CFME is converted to an output voltage via a current-to-voltage converter, which amplifies the signal with minimal low-pass filtering. The data acquisition cards are responsible for digital-to-analog and analog-to-digital conversion to and from the headstage, respectively. The gain of the current-to-voltage amplifier is set to convert the output voltage to current based on the size of the electrode.

2.3.1.2 Waveform Selection and Data Interpretation

The potential waveform used for FSCV has a direct impact on the sensitivity and chemical selectivity of the measurement. For example, dopamine measurements are commonly made using a waveform that scans from -0.4 to +1.3 V and back (vs. Ag/Ag/Cl), using a scan rate of 400 V s-

1, and a waveform application frequency of 10 Hz. The voltammetric scan is completed in ~10 ms and the time between waveform applications is held at -0.4 V, facilitating the pre-concentration

(adsorption) of positively-charged dopamine at the electrode surface, even when it is present at low concentrations.50 Application of more extreme oxidation potentials improves the detection of many analytes through the generation of oxygen-containing functionalities on the carbon surface, which can facilitate adsorption. Sufficiently positive potentials have even been shown to etch the carbon, to provide a renewed sensor surface with each voltammetric scan.51–53 Both the waveform and its application rate directly impact measurement sensitivity, chemical selectivity, temporal

31

resolution, and data density. For these reasons, the waveform can be manipulated to enable detection of a number of analytes that cannot be reliably detected using the standard triangular waveform, such as serotonin, oxygen, and neuropeptides.15,31,43 The waveform should be selected based on the chemical nature of the analyte (Figure 2), the presence of potential interferents, and the lifetime of the targeted species in the extracellular space.

Figure 2.2. The FSCV potential waveform can be optimized for detection of different molecules. Reprinted with permission from Roberts, J.G.; Sombers, L.A. Fast-Scan Cyclic Voltammetry: Chemical Sensing in the Brain and Beyond. Analytical Chemistry. 2018, 90, 490-504. Copyright 2018 American Chemical Society.

Using conventional FSCV methods, measurements are made with a high sampling rate

(tens to hundreds of kHz) and the waveform is repeated every ~100 ms. Thus, a significant amount of information is collected in real time. To quickly interpret real-time recordings, data are displayed in the form of a color plot (Figure 1) in which time (or CV number) is plotted on the x-

32

axis, potential (or data point within a CV) is plotted on the y-axis, and current is plotted in false color. Color plots allow data collected over tens of seconds to be quickly visualized. For extended recordings (>30-90 s), the time point at which the background is subtracted can be moved, to change the reference point. Taking a vertical slice of the color plot returns a CV that can be used for analyte identification, and a horizontal slice returns information about chemical dynamics in the form of current versus time. The current collected at a specific potential can be converted to concentration through sensor calibration. In this way, information regarding the fluctuations of one or more analytes can be interpreted using voltammograms as chemical identifiers.

2.3.1.3 Stimulated Neurochemical Release

A variety of strategies can be used to depolarize neurons and induce neurotransmitter release from terminals. Perhaps the most straightforward approach is to use direct electrical stimulation. For this, a calibrated biphasic stimulus isolator delivers a waveform to the stimulating electrode, resulting in the controlled injection of current. The stimulating waveform is typically a biphasic square wave with a defined frequency, amplitude, pulse width, and number of pulses

(pulse time), depending on the experimental goals. For instance, a typical electrical stimulation of the rat midbrain consists of a series of 60-120, biphasic pulses (125-150 µA) with a 2 ms width, applied at 60 Hz.54 The stimulating waveform can be computer controlled (software) or delivered using a digitally triggered waveform generator. To avoid significant disturbance of electrochemical recordings, the stimulus is only delivered in the time between potential sweeps

(for FSCV). Additional approaches for cellular depolarization include optical stimulation via optogenetics, chemogenetic stimulation via designer receptors that are exclusively activated by designer drugs (DREADDs), and microinfusion of a chemical stimulus.

33

2.3.2 Electrode Fabrication

CFMEs are typically insulated with either borosilicate glass or fused-silica capillary

(Figure 1H).55 Glass insulated CFMEs are easy to produce; however, they are relatively fragile and thus they are most commonly used for measurements in tissue slices, or in anesthetized animals. CFMEs that are insulated with silica capillary take longer to fabricate, but are more robust. Thus, these sensors can be used to reliably monitor neurotransmission during behavioral studies, sometimes lasting months.22

2.3.2.1 Borosilicate-Glass CFME

1. Connect a vacuum pump to a Büchner (vacuum) flask held in place using a ring-stand. Place a

single-holed, rubber stopper in the flask, and attach a long plastic tube with a small inner

diameter.

2. Place a single carbon fiber on a flat, clean, high-contrast surface that is well illuminated. Using

one hand, pinch one end of a borosilicate glass capillary in the small inner diameter tubing.

With the other hand, press one end of the carbon fiber against the surface. Carefully aspirate

the fiber through the glass capillary so that the fiber extends from both ends.

3. Place the filled capillary in a vertical capillary puller. Heat softens the glass and gravity pulls

the capillary to create two individual segments of glass, each with a taper that forms a seal

around the carbon fiber. Cut the fiber with a scissor to make two separate electrodes.

4. Cover a microscope slide with transparent tape, and affix a notched piece of putty at one end.

The putty serves as a platform to gently lower the taper/seal of the electrode onto the slide.

Inspect individual electrodes to ensure a tight glass seal is formed around the carbon fiber.

34

5. Cut the exposed carbon fiber to length (~100 µm is standard) with a scalpel under a

magnification of 10x. Verify the integrity of the seal and the length of the exposed carbon

under 40x magnification. Discard the electrode if defects are observed.

6. Using a soldering , solder a gold pin connector to one end of a steel wire lead. Then,

establish electrical connection with the carbon fiber. This can be accomplished by:

a. Coat a steel wire lead with conductive silver paint and carefully insert it into the back

of the glass capillary. Slight rotation of the wire ensures connectivity with the carbon

fiber.

b. Backfill the glass capillary with a solution of 4 M potassium acetate and 150 mM

potassium chloride. Then insert a wire lead to establish electrical connection.

7. Affix the lead to the glass capillary using small-diameter heat-shrink tubing.

2.3.2.2 Fused-Silica CFME

1. Cut fused-silica tubing using a pristine scalpel on a sheet of paper folded so that cut sections

do not become lost. The length of the tubing segments will be tailored to the desired depth of

the implant. To ensure clean cuts, replace the scalpel blade frequently.

2. Lower the silica tubing slowly, and at an angle, into a shallow glass dish containing 70%

isopropyl alcohol (IPA). The goal is to remove any air bubbles from the tube. This may take

~20-30 min.

3. Place pre-cut carbon fibers (~2 cm longer than tubing) into the 70% IPA bath. Using a boom

microscope for visualization, use tweezers to hold the tubing, and using a cotton swab, float

the carbon fiber into the tubing until it extends out of both ends equally.

4. Remove from the IPA bath, and allow outside of tubing to dry. Using a microscope, note the

more cleanly cut end of the tube. Affix filled tubing to a piece of folded tape (L or V-shaped)

35

placed on a small paper, in a manner that suspends the tubing in air. Fix paper to a container

to prevent movement. Allow the IPA to completely evaporate overnight.

5. Take an individual piece of paper with attached tubing and visualize under a boom microscope.

Place a drop of two-part hardening epoxy resin on the end of a small needle (26 gauge) and

slowly raise the epoxy drop from below to deposit a small bead of epoxy at the clean-cut end

of the tubing. This will serve as the electrode seal. Allow the epoxy to dry overnight.

a. An ideal seal will be dome-shaped, with no epoxy on the exposed carbon fiber, and

minimal to no epoxy left on the shaft of the silica tubing.

b. If epoxy is evident on the extended carbon fiber, slowly pull the other end of the fiber

to pull the epoxy into the dome.

c. Tip: If there are issues obtaining domed epoxy seals, try allowing the epoxy to thicken

prior to application.

6. Apply a bead of conductive silver epoxy to the opposite end of the capillary (the other end of

the carbon fiber should be evident here). Place a gold pin into the bead of silver epoxy, parallel

to the shaft of the tubing. Allow this to dry overnight, or cure in the oven at 100 °C for one

hour.

7. Apply two-part hardening epoxy resin with a toothpick to seal in the silver epoxy segment.

Allow this to dry overnight, or cure at 100 °C for 30 min.

8. Coat the silver/epoxy/gold pin connector assembly with liquid electrical tape using a toothpick.

Ensure that this coating also interfaces with the fused-silica tubing. Allow this to dry.

9. Cut the exposed carbon fiber to the desired length under a microscope as described in 3.2.1.

a. Alternatively, the electrode can be cut to the desired length between steps 5 and 6.

36

2.3.2.3 Ag/AgCl Reference Electrode

1. Cut a piece of silver wire to approximately 10 mm and solder it into a gold pin connector.

2. Cover the solder with a quick dry epoxy, so that soldering material does not ultimately come

into contact with tissue.

3. On the day of surgery, the reference electrode is chloridized using 0.1 M HCl and a 9 V battery.

Connect the positive terminal to the gold pin on the silver wire, and connect a wire or paper

clip to the negative terminal. Submerge both wires into the 0.1 M HCl for ~10 s, or until the

surface of the silver wire turns slightly white.

a. Alternatively, the reference wire can be chloridized by soaking the exposed silver in

Clorox bleach for 2-4 h.

2.3.2.4 CFME Pretreatment and Assessing Viability

Prior to use in an animal, electrode viability is generally confirmed in a buffered electrolyte. Electrodes can also be subjected to pretreatments such as flame-etching,56 laser activation,57,58 and electrochemical conditioning.51,59 Such pretreatments can substantially improve sensitivity to a number of analytes. Due to the relative ease of electrochemical conditioning, this method will be briefly described.

1. Place a chloridized silver electrode into a beaker containing a buffered electrolyte. Lower the

exposed carbon tip of the CFME into the buffered electrolyte. Connect the headstage leads

(from the potentiostat) to the appropriate electrodes.

2. All potentiometric conditioning generally involves applying potentials ≥+1.0 V. Perhaps the

most common approach is to apply an FSCV waveform scanning from -0.4 to +1.4 V at 400

V s-1 using a waveform application frequency of 60 Hz, for ~30 min. Over this time, the

37

background signal should grow substantially, indicative of increasing electrode capacitance.

The background signal should ultimately stabilize.

3. Lower the waveform application frequency to 10 Hz, and cycle on the waveform for an

additional 5-10 min.

4. Verify electrode integrity. Make a 30-60 s recording in the buffered electrolyte using a 10 Hz

waveform application frequency. Analyze the recording for excessive noise, glitches, and

electrode drift (changes in the background signal over time).

a. Assess noise by inspecting CVs around the time-point of background-subtraction, and

the current versus time traces. There may be differences in noise levels associated with

different instruments, headstages, and gain, so evaluation of acceptable noise will be

based on application (i.e. how large are the signals of interest).

b. Glitches appear as large spikes in current. These are usually indicative of poor electrical

connection to the carbon fiber.

c. Drift is indicative of electrode instability, and can be evaluated by inspecting color plots

and current versus time traces (across all potentials) for changes in current over time.

If there is excessive drift (>1-2 nA for a 100-µm CFME), further conditioning may

stabilize the electrode. Alternatively, drift may indicate an imperfect seal about the

carbon fiber. This issue is not remediated with additional conditioning.

38

Figure 2.3. Representative FSCV recordings with different levels of noise. The left color plot is from a recording made outside a faraday cage with an ungrounded potentiostat. The right color plot is from a noise optimized recording in a faraday cage with a grounded potentiostat.

2.3.3 Surgery for Electrode Implantation in Rat Brain Tissue

FSCV has been utilized in a range of animals including rats,4,5,13 mice,15 voles,60 birds,61 fruit flies (drosophila),62 pigs,63 non-human primates,64 and even in humans.65 The following section will describe implantation of a CFME in rat striatum to monitor dopamine release. These methods can serve as the framework for studies targeted at monitoring naturally occurring and stimulated neurotransmission in different brain regions and pathways. Note that all procedures should be approved by the appropriate Institutional Animal Care and Use Committee prior to adoption in the laboratory.

1. Sterilize working area and stereotaxic instrument prior to procedure. Ensure that all surgical

tools are clean and free of debris, then sterilize using an autoclave or bead sterilizer. Make sure

other materials (swabs, gauze, etc.) are also sterile.

2. Anesthetize the rat using an approved procedure. Check for reflexes, and shave the top of the

head using a set of electric trimmers. Then, secure the animal in a stereotaxic frame. Apply eye

lubricant and rest the animal on heating pad to maintain body temperature.

3. Remove any loose hairs around shaved site. Clean the shaved site by alternating application of

70% isopropyl alcohol and Betadine®, using a new gauze each time. Repeat three times.

39

4. Locally anesthetize the scalp with a subcutaneous injection of 0.25% bupivacaine (~0.3 mL).

Make an incision (~15-20 mm) on the top of the head, parallel with the animal’s spine.

Carefully retract and cut away the skin to expose the skull (~20x15 mm, Fig. 4). Removal of

conjunctive tissue and cleaning of the skull is facilitated by using 2-3% hydrogen peroxide and

a cotton swab.

5. At this point, both bregma and lambda should be clearly visible (Figure 4). Use the stereotaxic

apparatus to level the head until the difference in the dorso-ventral (DV) coordinates at bregma

and lambda is <0.2 mm. Make note of bregma coordinates. These will serve as a reference

point for all implants.

6. Mark the skull above the desired implantation sites using a rat atlas. For example, a location

that is +1.2 mm anterior-posterior (AP) and +1.4 mm medial-lateral (ML) relative to bregma)

can be used to position the working electrode above the nucleus accumbens of a rat (~300 g).

If incorporating a stimulating electrode (or optrode) into the experimental design, this position

can also be marked (e.g. -5.2 mm AP and +1.0 mm ML for the ventral tegmental area). Mark

a spot contralateral to the working electrode for placement of the reference electrode. See

Figure 4.

7. Use a rotary tool (Dremmel®) and a small-gauge drill bit to carefully drill through the skull at

the marked locations. Drill additional holes for surgical screws:

a. Anesthetized preparation: Drill a hole at ~45° near the reference electrode, and partially

insert surgical screw. See Figure 4.

b. Survival surgery: Drill four holes at ~45° at the corners of the exposed skull and insert

surgical screws. See Figure 4.

40

8. Use the stereotaxic instrument to slowly lower electrodes. Prior to each implantation, use a

sterile needle to clear dura mater. Prior to application of cement, ensure skull is clean and dry.

a. Anesthetized preparation: Lower the reference electrode into brain tissue and secure in

place using cranioplastic cement. The cement should cover both the reference electrode

and the adjacent surgical screw.

b. Survival surgery: Mount the reference electrode in a locking intracerebral guide

cannula. Lower the guide cannula/reference electrode assembly, and affix to the skull

and to the closest surgical screw using cranioplastic cement. This allows a fresh

reference electrode to be carefully inserted into tissue for electrochemical

measurements, and then removed from the guide cannula and replaced with a stylet

while the rat is in the home cage.

c. If using a stimulating electrode or optrode, it is slowly lowered to target depth and can

be cemented in place.

d. Lower the CFME to just above the hole. Connect the CFME and reference electrodes

(once cement has hardened) to the headstage and begin applying a FSCV waveform.

Continue to lower the CFME until electrical connection is established. Then slowly

lower the CFME to target depth (-7.0 mm DV for ventral striatum).

9. If this is a survival surgery, affix the working and stimulating electrodes to the skull and

complete the domed cement ‘headcap’ by filling in gaps between electrodes with cement,

covering all surgical screws. This cap should fill the exposed skull area. Allow to harden.

Remove animal from stereotaxic instrument, and place in a clean cage on top of a heating pad

until recovery. Once fully awake, offer soft food and fresh water along with a fruit-flavored

41

analgesic (acetaminophen 0.1-0.3 g/kg). Monitor the animal’s implantation/headcap site,

weight, and behavior daily for abnormalities. Generally, animals recover within 1-2 days.

10. Additional Notes

 See section 3.4.1 for process that can be incorporated into the procedure for freely

moving surgeries.

 If a stimulating electrode is not to be implanted, the stimulation wires can be cut, and

the threaded connector can be placed in the cranioplastic cement to provide a tether.

 Breathing rate (breath/s) and hydration should be monitored throughout the surgery.

Figure 2.4. A graphical representation of the exposed rat skull during surgery. Example markings for placement of electrodes and surgical screws are highlighted. Stimulating electrode/tether image: copyright Plastics One 2019.

42

2.3.4 Voltammetric Measurements

2.3.4.1 Anesthetized Preparation

1. Immediately following surgery, place the stereotaxic instrument inside a grounded faraday

cage.

2. Connect the working and reference electrodes to the headstage. Connect the stimulating

electrode, as appropriate.

3. Condition the CFME on the FSCV waveform to allow it to stabilize in the recording

environment. A typical triangular waveform (-0.4 V – 1.3 V) can be applied at 60 Hz for ~15

min, followed by another 5-10 min at 10 Hz.

4. If using a stimulating electrode, cell bodies in the midbrain can be depolarized and recordings

at the working electrode can be monitored for time-locked dopamine release at the terminals.

Stimulations are generally spaced in ~5 min increments to allow replenishment of vesicular

stores.

5. If no (or low-amplitude) dopamine release is observed, lower the working electrode (0.1 mm

increments) and repeat stimulation. Alternatively, stimulating electrode placement can be

similarly adjusted (if it was not cemented in place).

6. The overall procedure is dependent on the experimental goals. Generally, this will involve

monitoring stimulated or phasic neurochemical release before and after some pharmacological

manipulation.

7. Additional Notes

 During the freely moving surgery, a similar procedure described in steps 4-5 can be

used to position the working electrode prior to affixing in place with cranioplastic

cement.

43

 Specific anesthetics can directly influence phasic dopamine release.66

 Breathing rate (breath/s) and hydration should be continually monitored.

2.3.4.2 Freely-Moving Preparation

Three to five days after surgery, the animal can be prepared for the experiment (if fully recovered). However, many experimenters allow for up to 3 weeks of recovery time, during which the animal can begin various behavioral trainings/assessments.

1. Behavior chamber should be placed inside a grounded faraday cage, and appropriate

connections with instrumentation, lights, and cameras established.

2. Test electrical connectivity and noise through the headstage and commutator using a dummy

circuit (resistor and capacitor in series), prior to experimentation. The headstage cable should

be just long enough to allow the animal to explore the entirety of the behavior chamber, and

noise should be minimal (<~0.3 nA). It is important that test recordings are free of noise prior

to attaching an electrode, as noise will increase when an electrode is connected. There should

be no evidence of any pattern or striping in the color plots. If noise is detected, disconnect all

non-essential components to the system (lights, behavioral lines, etc) and test again.

Components can be added, one at a time, until the problem is identified.

3. On experiment day, place the animal inside the behavioral chamber, and tether. Connect the

working and reference electrodes to the headstage.

4. Condition the CFME on the desired waveform for ~15 min at 60 Hz, and then for ~10 min at

10 Hz, until the electrode is stable, as determined by inspection of the background current.

During this time, the animal acclimates to the behavioral chamber.

5. Proceed with FSCV recordings, behavioral tests, pharmacological manipulations, etc. See

Figure 5 for representative data from a recording in the nucleus accumbens. This file includes

44

an electrical stimulation (12 biphasic, 175 μA pulses at a frequency of 60 Hz with a pulse width

of 2 ms) of the ventral tegmental area.

Figure 2.5. Example of a dopamine release event (a) recorded in the nucleus accumbens of a freely moving rat in response to electrical stimulation of the ventral tegmental area. In this example, H2O2 (b) is also recorded along with a change in pH (c), due to a hemodynamic response to the stimulation.

2.3.5 Post-Experiment

2.3.5.1 Verification of Electrode Placement

The tip of the CFME is too small to leave an observable mark in the tissue. Therefore, an electrical lesion can be created at the carbon-fiber tip by applying a high current in an anesthetized animal.22 However, this process renders the electrode useless for calibration. The rat is transcardially perfused (0.9% saline and 10% formalin), then decapitated and the brain removed for histology.

2.3.5.2 Electrode Post-calibration

Whenever possible, the electrode should be carefully removed from the brain for calibration. The electrode is rinsed in water and calibrated for physiological concentrations (ideally

≥3) of the analytes of interest using a flow-injection apparatus (Figure 6). This system is composed

45

of a simple electrochemical cell and a sample loop that is used to reproducibly inject a small bolus of analyte onto the electrode surface. A syringe pump provides a constant flow (0.5-3 mL/min) of buffer through the cell. A six-port HPLC value and a digitally controlled pneumatic actuator control analyte injections. The reference electrode is submerged in the buffer, the working electrode is slowly lowered into the center of the cell using a micromanipulator, and both are connected to the headstage. The same waveform used for in vivo measurements is applied to the

CFME, and samples are introduced to the electrode surface under software control. Each concentration is sampled at least in triplicate, and the averaged peak current is plotted against concentration. A linear regression is used to relate collected current to analyte concentration.

Figure 2.6. FSCV flow cell and headstage. A) A typical flow cell used for electrode calibration. B) Headstage used to connect the potentiosat to the electrodes. The same headstage is used to tether the animal during freely moving measurements. Tethering strategies can differ. Images: copyright Pine Research Instrumentation 2019, photos used with permission.

46

2.3.6 Data Analysis

High-Definition Cyclic Voltammetry (HDCV) software can be used for data analysis and visualization, as well as to export data as ‘.txt’ files. MATLAB (Mathworks, Inc., Natick, MA) or

R (R Core Team; https://www.r-project.org/) can also be used for data visualization, processing and quantitative analysis. Principal Component Regression (PCR; Figure 7) is the most common multivariate statistical analysis approach used to assess complex FSCV data. It is used to separate intensity-based data into relevant vector components and noise. A training set is constructed that contains cyclic voltammograms for the major species observed in a recording, at ~5 physiological concentrations each. Principal components describing major sources of variance (from both analyte and interferent signals) in the data set are selected, allowing species with sufficiently unique sources of variance to be discriminated from other analytes and interferents, and for noise to be discarded. Then, unknown cyclic voltammograms recorded in vivo are projected onto the principal components to determine unknown concentrations. A well-detailed explanation of the use and validation of PCR in analysis of FSCV data can be found elsewhere.34,67

47

Figure 2.7. Principal component regression. A) A representative CV with contributions from both dopamine and hydrogen peroxide. B) Training set data for these analytes must exhibit the correct shape (and peak position), and encompass the magnitude of the signals measured in vivo. C) Projection of the training data onto factor space. Vectors, or principal components, are assigned to the largest sources of variances, with each orthogonal to the next. The scores of the principal components inherent to each training CV are regressed against analyte concentration. The mixed signal is then projected onto the principal component space. D) Mixed signal PC scores are used to determine the contribution of each analyte.

2.3.7 Additional Experimental Options

2.3.7.1 Electrophysiology

FSCV can be combined with other approaches, such as electrophysiological measurements of cell activity.16,49 For instance, a switch can be incorporated into the headstage to disconnect the microelectrode from the potential waveform between sweeps. Thus, the electrode ‘floats’, adopting the potential of the recording microenvironment in which it resides. This method allows specific molecular dynamics to be correlated with local neuronal activity.

48

2.3.7.2 Optogenetics and DREADDs

Advances in biotechnology have enabled a range of optogenetic and chemogenetic tools for investigation of brain function. Optogenetic approaches employ a virus to genetically modify a specific set of neurons to express light-sensitive ion channels (typically rhodopsins).68,69 An implanted optical fiber can then be used to selectively activate or inhibit the specific subsets of neurons. However, penetration of light in tissue is limited, and use of high intensity light can result in artifacts (photoelectric effects), or local heating of tissue (photothermal effects).70

Chemogenetic approaches are similar in that a virus can be used to selectively target specific neuronal subsets in a given brain region. The targeted cells are genetically modified to express designer G-protein coupled receptors that are selectively activated or inhibited by non-endogenous designer drugs (DREADDS).71,72 The most common designer drug is probably clozapine N-oxide

(CNO), which can be systemically or locally administered.

2.3.7.3 Microelectrode Arrays and Wireless Headcaps

Recent advances in instrumentation and microelectrode technologies have enabled microelectrode arrays to be incorporated into voltammetric experiments, providing regionally specific chemical information.73 For example, arrays of polymer-insulated CFMEs have been used to monitor dopamine dynamics at multiple recording sites in rat striatum over months, with minimal tissue damage.64

Most experiments in awake animals employ a tethered wire connection, which can significantly limit the options for behavioral challenges and can introduce electrical artifacts in response to vigorous movement of electrical connections. However, wireless integrated circuits can be utilized to create miniaturized potentiostat devices mounted directly on the animal itself.74,75

This advance can enable investigation of social interactions and more natural movement. However,

49

limitations in wireless data transfer rates, especially in studies incorporating multiple electrodes, have limited widespread use. Real-time data compression strategies prior to transfer are leading to more utilization of this technology.76

2.3.7.4 Electrode Materials and Membrane Coatings

A significant amount of work has been done to characterize carbon-based materials for in vivo studies. Materials such as diamond,77 carbon nanotubes,27,78 nanofibers,79 and nanotube yarns26,58 have all been utilized as sensors, and can offer improved performance in certain experimental situations. In addition, it is becoming more common to exploit nanoscale electrode materials for in vivo studies, as the smaller size domain can offer enhanced sensitivity and selectivity though and enhanced mass transfer and faster electron transfer kinetics.27,28,78 It is also common to coat the sensor surface with chemically selective polymer membranes to enhance sensitivity, selectivity, or stability. Nafion is a perfluorinated ion-exchange polymer that is often used, sometimes with polyethylenedioxythiophene (PEDOT), on the CFME surface to enhance sensitivity to dopamine and to decrease sensitivity to ascorbic acid.80,81 Nanoparticles, and carbon nanomaterials can be incorporated into these membranes to further enhance electrode performance.78,82 Chitosan hydrogels have been used to entrap various oxidase at the carbon electrode surface.83,84 This has allowed for indirect detection of non-electroactive molecules, such as glucose and lactate, by way of voltammetric measurements of hydrogen peroxide; eg., glucose oxidase produces gluconic acid and hydrogen peroxide in the presence of glucose. With this approach, it is possible to simultaneously record both electroactive and non- electroactive species at the same recording site.85

50

2.3.8 Critical Considerations and Alternative Methods

2.3.8.1 Calibration and Principal Component Regression for In Vivo Recordings

When making FSCV measurements, it is necessary to calibrate each electrode for accurate quantification of neurochemical events. Individual electrode performance is variable and dependent on the chemical nature of the electrode at the microscopic level, as well as the nature of the insulating seal that defines the sensor size at the macroscopic level. Acutely implanted sensors can be removed and post-calibrated in a flow cell. This approach has been demonstrated to be effective; however, it does not account for possible matrix effects. Analyte(s), macromolecules, small organic molecules, and even spectator ions can interact with the carbon surface through potential-specific reactions to affect performance. The use of permanently implanted sensors presents additional analytical challenges, as electrodes can become significantly fouled,53,86 can be encapsulated in response to the implantation (gliosis),3 and are difficult to remove for calibration.

Thus, electrode performance is assumed to be constant for the entirety of the experiment, which can last months.

Overall, these issues complicate the construction of appropriate training sets for multivariate analysis, typically achieved using principal component regression (PCR). A training set for dopamine can be collected in vivo using stimulated release, or using dopamine transients that are reliably evoked by introduction of an unexpected reward stimulus (ex., introduction of a sugar pellet).87 However, ensuring that a chemically independent signal is obtained in vivo is difficult. Alternatively, it is also common to utilize data recorded at another electrode in another animal, or collected ex vivo in a flow-cell to serve as training data for PCR.7,67,88 This approach, although in direct contrast to proper quantitative calibration, has proved effective if significant care is taken in analysis. However, it is important to note that the analyte-specific voltammograms

51

and scaling factors that lie at the core of PCR are dependent on the chemical nature of the recording environment. CVs recorded in vivo can differ in overall shape compared to those for the same analyte recorded at another time in the living brain, or when compared to CVs recorded in a bench- top flow cell. If there is significant mismatch between the training CVs and the CVs inherent to the unknown data set, model validation will fail. Misidentification of analyte contributions will lead to misestimation of chemical dynamics.89 In response to this, it has become standard to apply a potential offset (~200 mV) to the applied waveform during in vivo measurements in order to artificially shift in vivo CVs such that they appear more like those recorded in vitro.22,67,90

However, it is in this author’s opinion that this approach should be avoided, as application of a potential offset shifts the applied waveform, significantly influencing performance.

There have been recent advances in electrode calibration strategies for in vivo measurements. It has been shown that the shape and magnitude of the background signal serves as a predictor of both electrode sensitivity and the overall cyclic voltammogram shape for multiple analytes.51,91 Another study has demonstrated that exposing the sensor to a buffer containing bovine serum albumin before pre-calibration gives similar results as conventional post-calibration strategies that follow implantation.92

2.3.8.2 Background Subtraction and Electrochemical Drift

FSCV is inherently a differential technique, because the background current is subtracted to reveal faradaic contributions to the current that result from the oxidation or reduction of chemical species present in the vicinity of the electrode. Recently, an approach termed fast-scan controlled-adsorption voltammetry (FSCAV) has allowed direct determination of neurotransmitter basal levels by manipulation of analyte adsorption on the electrode surface.93,94 FSCAV functions by applying an FSCV waveform at a high frequency (~100 Hz) to thoroughly electrolyze any

52

adsorbed species on the electrode surface. This is followed by a seconds-long period at an electrostatically advantageous potential to adsorb analyte. Finally, there is a period of typical

FSCV measurements to quantify material adsorbed to the electrode surface. The first CVs collected after the controlled- adsorption period contain information related to the basal levels of neurotransmitter in the vicinity of the electrode. This technique is only applicable to species that adsorb to the sensor surface, such as dopamine and serotonin, with low nanomolar limits of detection. However, this multi-step technique offers significantly less temporal resolution than

FSCV.

In FSCV, the background signal is only stable for ~60 s. Following this period, significant and diverse background subtraction artifacts, termed electrochemical drift, complicate data interpretation. Thus, in a continuous recording, the reference (background) must be repositioned in ~60 s increments to analyze the entirety of the recording. Assessment of more gradual changes in analyte concentrations occurring over several minutes cannot be reliably evaluated. A number of strategies have been developed to address the limited time window of FSCV. A charge-balanced waveform can be paired with background subtraction to permit the evaluation of more gradual changes in analyte concentration over an hour, albeit with a significant decrease in temporal resolution.95 Another study has incorporated a small potential-step to sample the impedance state of the sensor prior to application of each triangular waveform. When combined with a convolution-based approach to data analysis, this strategy allows prediction and removal of nonfaradaic contributions to electrode drift.35 This method is particularly effective for removal of drift attributed to rapid changes in local ionic composition. A third method involves the use of a paired voltammetric waveform in which a smaller triangular sweep is used to sample the contribution of electrode drift to the signal immediately prior to each full voltammetric sweep.

53

Partial-least-squares regression is used to predict and subtract the drift contribution to date recorded in each potential sweep.36 This method provides a simple manner to evaluate both rapid and gradual molecular fluctuations in ~10-30 min in vivo recordings, while maintaining the temporal resolution.

2.3.8.3 Time-Consuming Data Analysis

The complex and seemingly random nature of the phasic events monitored during in vivo studies can easily result in a significant amount of time spent in data analysis, particularly in light of the issues discussed in 3.8.1 and 3.8.2. The general complexity of the signal necessitates appropriate background subtraction and careful construction and validation of training sets during data analysis. This can introduce human bias and error when analyzing data. An exciting step has been made toward automation of data analysis.96 In this work, an algorithm was constructed using

MATLAB to evaluate extended (>10 min) recordings of adenosine transients in live rats. The script was used to adjust the background position in a controlled manner based on well-defined criteria. This approach was used to successfully identfy transient neurochemical events with a high degree of accuracy, as compared to the tradiational approach to data analysis by way of a human analyst. In this case, the human analysts spent >10 hours evaluating the recording, whereas the algorithm was used to achieve similar results in <1 hour. Significant expansion of these methods will be vital for increasing the throughput of FSCV analysis.

2.3.8.4 Data Density and Analog/Digital Filtering

During a typical in vivo FSCV experiment, it is common to make both short measurements

(~1-2 min) and continuous recordings (10 mins to >hour). The default sampling rate and waveform application rate in HDCV software are 100 kHz and 10 Hz, respectively. Thus, the amount of data recorded becomes quite significant and requires adequate data storage. It has been demonstrated

54

that reduced sampling and waveform application rates can provide adequate temporal and electrochemical resolution for monitoring subsecond chemical dynamics in many experimental paradigms.97,98

It is common to use an analog filter to smooth the digitized waveform output. This is often a low-pass filter with a 2 kHz cutoff frequency. It is also common to digitally filter the recorded current response. For example, the default filtering setting in HDCV employs a 4th order low-pass

Bessel filter with a 2 kHz cutoff frequency. Fourier transforms can also be used to filter data sets.48

When used carefully, filtering can greatly enhance the quality of the recording and ease data analysis. However, it is imperative that the effects of analog and digital filtering be critically assessed in advance, and that filter settings remain consistent for every recording included in the experimental data set. This is because filtering can directly impact the apparent shape of a CV, the position of redox potentials, and current magnitudes (Figure 8). These issues are exacerbated when lower sampling rates and/or faster scan rates are employed.

55

Figure 2.8. Analog and digital filtering in FSCV. (left) Effects of low-pass analog filtering on the applied waveform and CVs for dopamine. (right) Effects of digital low-pass filtering on the dopamine response. The trace in red represents the standard setting in HDCV and the blue trace represents no digital filtering. These data are displayed for triangular waveforms applied at A) 400 V s-1 and B) 1000 V s-1.

2.3.8.5 Reference vs Working Electrode Driven Systems

In a two-electrode setup such as that which is commonly utilized in FSCV, the potentiostat can directly drive either the reference or the working electrode to achieve the same potential difference across the system. In a reference-driven system, an inverse waveform is applied at the reference electrode to achieve the target potential at the working electrode. Alternatively, the potential can be directly controlled at the working electrode. Both methods result in identical measurements. However, it is imperative to consider future needs before purchasing an instrument.

For example, waveforms are selected based on the analyte(s) targeted. A reference-driven system will not be capable of individually addressing more than one working electrode at a time.

56

2.3.8.6 Electrical/Optical Stimulation Using HDCV

HDCV software can be used to design and control the delivery of pulse sequences used in optical and electrical stimulations. It is important to note that the pulsed waveform is gated so as not to interfere with the electrochemical measurements; i.e., the pulse train can only be applied between voltammetric sweeps. The user must be cognizant of this, to ensure delivery of the desired stimulation pulse train. For example, a 60 Hz stimulation consisting of 60 biphasic pulses (each 2 ms in duration) should presumably be delivered over the course of 1 s. However, the true stimulus delivered is dependent on the manner in which the stimulation pulses are gated relative to the applied waveform (Figure 9). In this case, if the ‘stim delay’ time is 20 ms, then the true stimulus will more closely approximate a 10 Hz application of five 60 Hz pulses over 1 s, and only 50 total pulses will be delivered.

Figure 2.9. Stimulation pulses are gated to prevent overlap with application of the voltammetric waveform. This can inadvertently result in stimulation parameters that differ from the assumed output.

57

2.4 Conclusions

Fast-scan cyclic voltammetry (FSCV) is a powerful electroanalytical method that provides the means to monitor rapid (subsecond), low-level (nano/micro molarity) molecular species in real time. When paired with carbon-based microelectrode sensors, reliable measurements can be made at single cells, in live tissue slices, and in freely-moving animal subjects over the course of weeks to months. FSCV has provided remarkable insight regarding molecular mechanisms underlying specific aspects of goal-directed behavior and associative learning when appropriate behavior and pharmacological paradigms have been utilized. Advances in pharmacology and genetic methods to reliably manipulate brain circuitry have enhanced the power of behavioral and pharmacological studies. Recent advances in sensor materials and data analysis paradigms have expanded the detectable pool of analytes and enabled simultaneous monitoring of multiple species, both electroactive and non-electroactive, at a single recording site. Additionally, methods that extend the time-window from background-subtraction have allowed for evaluation of more gradual changes in analyte concentration, expanding the scope of FSCV studies. Miniaturization of the instrumentation and the development of microelectrode arrays are enabling the regional mapping of molecular dynamics, on a much broader spatial scale. Continued advances in semi-automated algorithms and multivariate strategies for higher throughput data analysis will empower researchers to draw more impactful conclusions from FSCV recordings.

58

2.5 References

1. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, 2006.

2. Bucher, E. S.; Wightman, R. M. Electrochemical Analysis of Neurotransmitters. Annu. Rev. Anal. Chem. 2015, 8, 239–261.

3. Peters, J. L.; Miner, L. H.; Michael, A. C.; Sesack, S. R. Ultrastructure at Carbon Fiber Microelectrode Implantation Sites after Acute Voltammetric Measurements in the Striatum of Anesthetized Rats. J. Neurosci. Methods 2004, 137 (1), 9–23. https://doi.org/10.1016/j.jneumeth.2004.02.006.

4. Day, J. J.; Roitman, M. F.; Wightman, R. M.; Carelli, R. M. Associative Learning Mediates Dynamic Shifts in Dopamine Signaling in the Nucleus Accumbens. Nat. Neurosci. 2007, 10 (8), 1020.

5. Stuber, G. D.; Klanker, M.; de Ridder, B.; Bowers, M. S.; Joosten, R. N.; Feenstra, M. G.; Bonci, A. Reward-Predictive Cues Enhance Excitatory Synaptic Strength onto Midbrain Dopamine Neurons. Science 2008, 321 (5896), 1690. https://doi.org/10.1126/science.1160873.

6. Gan, J. O.; Walton, M. E.; Phillips, P. E. M. Dissociable Cost and Benefit Encoding of Future Rewards by Mesolimbic Dopamine. Nat. Neurosci. 2010, 13 (1), 25–27. https://doi.org/10.1038/nn.2460.

7. Howe, M. W.; Tierney, P. L.; Sandberg, S. G.; Phillips, P. E.; Graybiel, A. M. Prolonged Dopamine Signalling in Striatum Signals Proximity and Value of Distant Rewards. nature 2013, 500 (7464), 575.

8. Nicola, S. M. The Flexible Approach Hypothesis: Unification of Effort and Cue-Responding Hypotheses for the Role of Nucleus Accumbens Dopamine in the Activation of Reward- Seeking Behavior. J. Neurosci. 2010, 30 (49), 16585. https://doi.org/10.1523/JNEUROSCI.3958-10.2010.

9. Robbins, T. W.; Everitt, B. J. A Role for Mesencephalic Dopamine in Activation: Commentary on Berridge (2006). Psychopharmacology (Berl.) 2007, 191 (3), 433–437. https://doi.org/10.1007/s00213-006-0528-7.

10. Flagel, S. B.; Clark, J. J.; Robinson, T. E.; Mayo, L.; Czuj, A.; Willuhn, I.; Akers, C. A.; Clinton, S. M.; Phillips, P. E. M.; Akil, H. A Selective Role for Dopamine in Stimulus–Reward Learning. Nature 2011, 469 (7328), 53–57. https://doi.org/10.1038/nature09588.

11. Mazzoni, P.; Hristova, A.; Krakauer, J. W. Why Don’t We Move Faster? Parkinson’s Disease, Movement Vigor, and Implicit Motivation. J. Neurosci. 2007, 27 (27), 7105. https://doi.org/10.1523/JNEUROSCI.0264-07.2007.

12. Howe, M. W.; Dombeck, D. A. Rapid Signalling in Distinct Dopaminergic Axons during Locomotion and Reward. Nature 2016, 535 (7613), 505–510. https://doi.org/10.1038/nature18942.

59

13. Kumar, U.; Medel-Matus, J.-S.; Redwine, H. M.; Shin, D.; Hensler, J. G.; Sankar, R.; Mazarati, A. Effects of Selective Serotonin and Norepinephrine Reuptake Inhibitors on Depressive- and Impulsive-like Behaviors and on Monoamine Transmission in Experimental Temporal Lobe Epilepsy. Epilepsia 2016, 57 (3), 506–515. https://doi.org/10.1111/epi.13321.

14. Wood, K. M.; Zeqja, A.; Nijhout, H. F.; Reed, M. C.; Best, J.; Hashemi, P. Voltammetric and Mathematical Evidence for Dual Transport Mediation of Serotonin Clearance in Vivo. J. Neurochem. 2014, 130 (3), 351–359. https://doi.org/10.1111/jnc.12733.

15. Dankoski, E. C.; Agster, K. L.; Fox, M. E.; Moy, S. S.; Wightman, R. M. Facilitation of Serotonin Signaling by SSRIs Is Attenuated by Social Isolation. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2014, 39 (13), 2928–2937. https://doi.org/10.1038/npp.2014.162.

16. Kirkpatrick, D. C.; McKinney, C. J.; Manis, P. B.; Wightman, R. M. Expanding Neurochemical Investigations with Multi-Modal Recording: Simultaneous Fast-Scan Cyclic Voltammetry, Iontophoresis, and Patch Clamp Measurements. Analyst 2016, 141 (16), 4902– 4911. https://doi.org/10.1039/C6AN00933F.

17. Belle, A. M.; Owesson-White, C.; Herr, N. R.; Carelli, R. M.; Wightman, R. M. Controlled Iontophoresis Coupled with Fast-Scan Cyclic Voltammetry/Electrophysiology in Awake, Freely Moving Animals. ACS Chem. Neurosci. 2013, 4 (5), 761–771. https://doi.org/10.1021/cn400031v.

18. Gómez-A, A.; Shnitko, T. A.; Barefoot, H. M.; Brightbill, E. L.; Sombers, L. A.; Nicola, S. M.; Robinson, D. L. Local μ-Opioid Receptor Antagonism Blunts Evoked Phasic Dopamine Release in the Nucleus Accumbens of Rats. ACS Chem. Neurosci. 2019, 10 (4), 1935–1940. https://doi.org/10.1021/acschemneuro.8b00437.

19. Duwensee, H.; Vázquez-Alvarez, T.; Flechsig, G.-U.; Wang, J. Thermally Induced Electrode Protection against Biofouling. Talanta 2009, 77 (5), 1757–1760. https://doi.org/10.1016/j.talanta.2008.10.011.

20. Kuhlmann, J.; Dzugan, L. C.; Heineman, W. R. Comparison of the Effects of Biofouling on Voltammetric and Potentiometric Measurements. Electroanalysis 2012, 24 (8), 1732–1738. https://doi.org/10.1002/elan.201200194.

21. Patel, J.; Radhakrishnan, L.; Zhao, B.; Uppalapati, B.; Daniels, R. C.; Ward, K. R.; Collinson, M. M. Electrochemical Properties of Nanostructured Porous Gold Electrodes in Biofouling Solutions. Anal. Chem. 2013, 85 (23), 11610–11618. https://doi.org/10.1021/ac403013r.

22. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat. Methods 2010, 7 (2), 126.

23. Park, J.; Aragona, B. J.; Kile, B. M.; Carelli, R. M.; Wightman, R. M. In Vivo Voltammetric Monitoring of Release in Subterritories of the Nucleus Accumbens Shell. Neuroscience 2010, 169 (1), 132–142. https://doi.org/10.1016/j.neuroscience.2010.04.076.

60

24. Arbuthnott, G. W.; Wickens, J. Space, Time and Dopamine. Trends Neurosci. 2007, 30 (2), 62–69.

25. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

26. Schmidt, A. C.; Wang, X.; Zhu, Y.; Sombers, L. A. Carbon Nanotube Yarn Electrodes for Enhanced Detection of Neurotransmitter Dynamics in Live Brain Tissue. ACS Nano 2013, 7 (9), 7864–7873. https://doi.org/10.1021/nn402857u.

27. Jacobs, C. B.; Vickrey, T. L.; Venton, B. J. Functional Groups Modulate the Sensitivity and Electron Transfer Kinetics of Neurochemicals at Carbon Nanotube Modified Microelectrodes. Analyst 2011, 136 (17), 3557–3565. https://doi.org/10.1039/C0AN00854K.

28. Puthongkham, P.; Yang, C.; Venton, B. J. Carbon Nanohorn-Modified Carbon Fiber Microelectrodes for Dopamine Detection. Electroanalysis 2018, 30 (6), 1073–1081. https://doi.org/10.1002/elan.201700667.

29. Unwin, P. R.; Güell, A. G.; Zhang, G. Nanoscale Electrochemistry of Sp2 Carbon Materials: From and Graphene to Carbon Nanotubes. Acc. Chem. Res. 2016, 49 (9), 2041–2048. https://doi.org/10.1021/acs.accounts.6b00301.

30. Cryan, M. T.; Ross, A. E. Scalene Waveform for Codetection of Guanosine and Adenosine Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2019, 91 (9), 5987–5993. https://doi.org/10.1021/acs.analchem.9b00450.

31. Calhoun, S. E.; Meunier, C. J.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin. ACS Chem. Neurosci. 2018, 10 (4), 2022–2032.

32. Ross, A. E.; Venton, B. J. Sawhorse Waveform Voltammetry for Selective Detection of Adenosine, ATP, and Hydrogen Peroxide. Anal. Chem. 2014, 86 (15), 7486–7493. https://doi.org/10.1021/ac501229c.

33. Jackson, B. P.; Dietz, S. M.; Wightman, R. Mark. Fast-Scan Cyclic Voltammetry of 5- Hydroxytryptamine. Anal. Chem. 1995, 67 (6), 1115–1120. https://doi.org/10.1021/ac00102a015.

34. Keithley, R. B.; Carelli, R. M.; Wightman, R. M. Rank Estimation and the Multivariate Analysis of in Vivo Fast-Scan Cyclic Voltammetric Data. Anal. Chem. 2010, 82 (13), 5541– 5551. https://doi.org/10.1021/ac100413t.

35. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal. Chem. 2017, 89 (11), 6166–6174. https://doi.org/10.1021/acs.analchem.7b01005.

61

36. Meunier, C. J.; McCarty, G. S.; Sombers, L. A. Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double-Waveform Partial-Least-Squares Regression. Anal. Chem. 2019, 91 (11), 7319–7327. https://doi.org/10.1021/acs.analchem.9b01083.

37. Johnson, J. A.; Gray, J. H.; Rodeberg, N. T.; Wightman, R. M. Multivariate Curve Resolution for Signal Isolation from Fast-Scan Cyclic Voltammetric Data. Anal. Chem. 2017, 89 (19), 10547–10555. https://doi.org/10.1021/acs.analchem.7b02771.

38. Root, D. H.; Hoffman, A. F.; Good, C. H.; Zhang, S.; Gigante, E.; Lupica, C. R.; Morales, M. Norepinephrine Activates Dopamine D4 Receptors in the Rat Lateral Habenula. J. Neurosci. 2015, 35 (8), 3460. https://doi.org/10.1523/JNEUROSCI.4525-13.2015.

39. Wolf, K.; Zarkua, G.; Chan, S.-A.; Sridhar, A.; Smith, C. Spatial and Activity-Dependent Catecholamine Release in Rat Adrenal Medulla under Native Neuronal Stimulation. Physiol. Rep. 2016, 4 (17), e12898. https://doi.org/10.14814/phy2.12898.

40. Hensley, A. L.; Colley, A. R.; Ross, A. E. Real-Time Detection of Melatonin Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2018, 90 (14), 8642–8650. https://doi.org/10.1021/acs.analchem.8b01976.

41. Jackson, B. P.; Dietz, S. M.; Wightman, R. Mark. Fast-Scan Cyclic Voltammetry of 5- Hydroxytryptamine. Anal. Chem. 1995, 67 (6), 1115–1120. https://doi.org/10.1021/ac00102a015.

42. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2013, 4 (5), 782–789. https://doi.org/10.1021/cn4000499.

43. Zachek, M. K.; Takmakov, P.; Moody, B.; Wightman, R. M.; McCarty, G. S. Simultaneous Decoupled Detection of Dopamine and Oxygen Using Pyrolyzed Carbon Microarrays and Fast-Scan Cyclic Voltammetry. Anal. Chem. 2009, 81 (15), 6258–6265. https://doi.org/10.1021/ac900790m.

44. Van Gompel, J. J.; Bower, M. R.; Worrell, G. A.; Stead, M.; Chang, S.-Y.; Goerss, S. J.; Kim, I.; Bennet, K. E.; Meyer, F. B.; Marsh, W. R.; et al. Increased Cortical Extracellular Adenosine Correlates with Seizure Termination. Epilepsia 2014, 55 (2), 233–244. https://doi.org/10.1111/epi.12511.

45. Pihel, K.; Schroeder, T. J.; Wightman, R. M. Rapid and Selective Cyclic Voltammetric Measurements of Epinephrine and Norepinephrine as a Method to Measure Secretion from Single Bovine Adrenal Medullary Cells. Anal. Chem. 1994, 66 (24), 4532–4537.

46. Phillips, P. E. M.; Wightman, R. M. Critical Guidelines for Validation of the Selectivity of In- Vivo Chemical Microsensors. TrAC Trends Anal. Chem. 2003, 22 (8), 509–514. https://doi.org/10.1016/S0165-9936(03)00907-5.

62

47. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108 (7), 2646–2687. https://doi.org/10.1021/cr068076m.

48. Bucher, E. S.; Brooks, K.; Verber, M. D.; Keithley, R. B.; Owesson-White, C.; Carroll, S.; Takmakov, P.; McKinney, C. J.; Wightman, R. M. Flexible Software Platform for Fast-Scan Cyclic Voltammetry Data Acquisition and Analysis. Anal. Chem. 2013, 85 (21), 10344–10353. https://doi.org/10.1021/ac402263x.

49. Takmakov, P.; McKinney, C. J.; Carelli, R. M.; Wightman, R. M. Instrumentation for Fast- Scan Cyclic Voltammetry Combined with Electrophysiology for Behavioral Experiments in Freely Moving Animals. Rev. Sci. Instrum. 2011, 82 (7), 074302. https://doi.org/10.1063/1.3610651.

50. Bath, B. D.; Michael, D. J.; Trafton, B. J.; Joseph, J. D.; Runnels, P. L.; Wightman, R. M. Subsecond Adsorption and Desorption of Dopamine at Carbon-Fiber Microelectrodes. Anal. Chem. 2000, 72 (24), 5994–6002.

51. Roberts, J. G.; Toups, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode Calibration Strategy for Voltammetric Measurements In Vivo. Anal. Chem. 2013, 85 (23), 11568–11575. https://doi.org/10.1021/ac402884n.

52. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir 2010, 26 (11), 9116– 9122. https://doi.org/10.1021/la9048924.

53. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal. Chem. 2010, 82 (5), 2020–2028. https://doi.org/10.1021/ac902753x.

54. Trevathan, J. K.; Yousefi, A.; Park, H. O.; Bartoletta, J. J.; Ludwig, K. A.; Lee, K. H.; Lujan, J. L. Computational Modeling of Neurotransmitter Release Evoked by Electrical Stimulation: Nonlinear Approaches to Predicting Stimulation-Evoked Dopamine Release. ACS Chem. Neurosci. 2017, 8 (2), 394–410. https://doi.org/10.1021/acschemneuro.6b00319.

55. Rodeberg, N. T. Improving the Collection and Calibration of Voltammetric Measurements of Striatal Dopamine Release, The University of North Carolina at Chapel Hill, 2017.

56. Strand, A. M.; Venton, B. J. Flame Etching Enhances the Sensitivity of Carbon-Fiber Microelectrodes. Anal. Chem. 2008, 80 (10), 3708–3715. https://doi.org/10.1021/ac8001275.

57. Poon, Melanie.; McCreery, R. L. In Situ Laser Activation of Glassy Carbon Electrodes. Anal. Chem. 1986, 58 (13), 2745–2750. https://doi.org/10.1021/ac00126a036.

58. Yang, C.; Trikantzopoulos, E.; Nguyen, M. D.; Jacobs, C. B.; Wang, Y.; Mahjouri-Samani, M.; Ivanov, I. N.; Venton, B. J. Laser Treated Carbon Nanotube Yarn Microelectrodes for Rapid and Sensitive Detection of Dopamine in Vivo. ACS Sens. 2016, 1 (5), 508–515. https://doi.org/10.1021/acssensors.6b00021.

63

59. Feng, J. Xing.; Brazell, Michael.; Renner, Kenneth.; Kasser, Richard.; Adams, R. N. Electrochemical Pretreatment of Carbon Fibers for in Vivo Electrochemistry: Effects on Sensitivity and Response Time. Anal. Chem. 1987, 59 (14), 1863–1867. https://doi.org/10.1021/ac00141a028.

60. Resendez, S. L.; Keyes, P. C.; Day, J. J.; Hambro, C.; Austin, C. J.; Maina, F. K.; Eidson, L. N.; Porter-Stransky, K. A.; Nevárez, N.; McLean, J. W.; et al. Dopamine and Opioid Systems Interact within the Nucleus Accumbens to Maintain Monogamous Pair Bonds. eLife 2016, 5, e15325. https://doi.org/10.7554/eLife.15325.

61. Xiao, L.; Chattree, G.; Oscos, F. G.; Cao, M.; Wanat, M. J.; Roberts, T. F. A Basal Ganglia Circuit Sufficient to Guide Birdsong Learning. Neuron 2018, 98 (1), 208-221.e5. https://doi.org/10.1016/j.neuron.2018.02.020.

62. Rees, H. R.; Anderson, S. E.; Privman, E.; Bau, H. H.; Venton, B. J. Carbon Nanopipette Electrodes for Dopamine Detection in Drosophila. Anal. Chem. 2015, 87 (7), 3849–3855. https://doi.org/10.1021/ac504596y.

63. Shon, Y.-M.; Lee, K. H.; Goerss, S. J.; Kim, I. Y.; Kimble, C.; Van Gompel, J. J.; Bennet, K.; Blaha, C. D.; Chang, S.-Y. High Frequency Stimulation of the Subthalamic Nucleus Evokes Striatal Dopamine Release in a Large Animal Model of Human DBS Neurosurgery. Neurosci. Lett. 2010, 475 (3), 136–140. https://doi.org/10.1016/j.neulet.2010.03.060.

64. Schwerdt, H. N.; Shimazu, H.; Amemori, K.-I.; Amemori, S.; Tierney, P. L.; Gibson, D. J.; Hong, S.; Yoshida, T.; Langer, R.; Cima, M. J.; et al. Long-Term Dopamine Neurochemical Monitoring in Primates. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (50), 13260–13265. https://doi.org/10.1073/pnas.1713756114.

65. Kishida, K. T.; Saez, I.; Lohrenz, T.; Witcher, M. R.; Laxton, A. W.; Tatter, S. B.; White, J. P.; Ellis, T. L.; Phillips, P. E. M.; Montague, P. R. Subsecond Dopamine Fluctuations in Human Striatum Encode Superposed Error Signals about Actual and Counterfactual Reward. Proc. Natl. Acad. Sci. U. S. A. 2016, 113 (1), 200–205. https://doi.org/10.1073/pnas.1513619112.

66. Votaw, J.; Byas-Smith, M.; Hua, J.; Voll, R.; Martarello, L.; Levey, A. I.; Bowman, F. D.; Goodman, M. Interaction of Isoflurane with the Dopamine Transporter. Anesthesiol. J. Am. Soc. Anesthesiol. 2003, 98 (2), 404–411.

67. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E. M.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234. https://doi.org/10.1021/acschemneuro.6b00393.

68. Deisseroth, K. Optogenetics. Nat. Methods 2011, 8 (1), 26.

69. Adamantidis, A. R.; Tsai, H.-C.; Boutrel, B.; Zhang, F.; Stuber, G. D.; Budygin, E. A.; Touriño, C.; Bonci, A.; Deisseroth, K.; de Lecea, L. Optogenetic Interrogation of

64

Dopaminergic Modulation of the Multiple Phases of Reward-Seeking Behavior. J. Neurosci. 2011, 31 (30), 10829. https://doi.org/10.1523/JNEUROSCI.2246-11.2011.

70. Kozai, T. D.; Vazquez, A. L. Photoelectric Artefact from Optogenetics and Imaging on Microelectrodes and Bioelectronics: New Challenges and Opportunities. J. Mater. Chem. B 2015, 3 (25), 4965–4978.

71. Armbruster, B. N.; Li, X.; Pausch, M. H.; Herlitze, S.; Roth, B. L. Evolving the Lock to Fit the Key to Create a Family of G Protein-Coupled Receptors Potently Activated by an Inert Ligand. Proc. Natl. Acad. Sci. 2007, 104 (12), 5163–5168.

72. Calipari, E. S.; Juarez, B.; Morel, C.; Walker, D. M.; Cahill, M. E.; Ribeiro, E.; Roman-Ortiz, C.; Ramakrishnan, C.; Deisseroth, K.; Han, M.-H. Dopaminergic Dynamics Underlying Sex- Specific Cocaine Reward. Nat. Commun. 2017, 8, 13877.

73. Zachek, M. K.; Takmakov, P.; Park, J.; Wightman, R. M.; McCarty, G. S. Simultaneous Monitoring of Dopamine Concentration at Spatially Different Brain Locations in Vivo. Biosens. Bioelectron. 2010, 25 (5), 1179–1185. https://doi.org/10.1016/j.bios.2009.10.008.

74. C. I. Dorta-Quiñones; X. Y. Wang; R. K. Dokania; A. Gailey; M. Lindau; A. B. Apsel. A Wireless FSCV Monitoring IC With Analog Background Subtraction and UWB Telemetry. IEEE Trans. Biomed. Circuits Syst. 2016, 10 (2), 289–299. https://doi.org/10.1109/TBCAS.2015.2421513.

75. M. Roham; D. P. Daberkow; E. S. Ramsson; D. P. Covey; S. Pakdeeronachit; P. A. Garris; P. Mohseni. A Wireless IC for Wide-Range Neurochemical Monitoring Using Amperometry and Fast-Scan Cyclic Voltammetry. IEEE Trans. Biomed. Circuits Syst. 2008, 2 (1), 3–9. https://doi.org/10.1109/TBCAS.2008.918282.

76. Zamani, H.; Bahrami, H. R.; Chalwadi, P.; Garris, P. A.; Mohseni, P. C–FSCV: Compressive Fast-Scan Cyclic Voltammetry for Brain Dopamine Recording. IEEE Trans. Neural Syst. Rehabil. Eng. 2017, 26 (1), 51–59.

77. Bennet, K. E.; Tomshine, J. R.; Min, H.-K.; Manciu, F. S.; Marsh, M. P.; Paek, S. B.; Settell, M. L.; Nicolai, E. N.; Blaha, C. D.; Kouzani, A. Z.; et al. A Diamond-Based Electrode for Detection of Neurochemicals in the Human Brain. Front. Hum. Neurosci. 2016, 10, 102. https://doi.org/10.3389/fnhum.2016.00102.

78. Ross, A. E.; Venton, B. J. Nafion–CNT Coated Carbon-Fiber Microelectrodes for Enhanced Detection of Adenosine. Analyst 2012, 137 (13), 3045–3051. https://doi.org/10.1039/C2AN35297D.

79. Zhang, D. A.; Rand, E.; Marsh, M.; Andrews, R. J.; Lee, K. H.; Meyyappan, M.; Koehne, J. E. Carbon Nanofiber Electrode for Neurochemical Monitoring. Mol. Neurobiol. 2013, 48 (2), 380–385.

80. Brazell, M. P.; Kasser, R. J.; Renner, K. J.; Feng, J.; Moghaddam, B.; Adams, R. N. Electrocoating Carbon Fiber Microelectrodes with Nafion Improves Selectivity for

65

Electroactive Neurotransmitters. J. Neurosci. Methods 1987, 22 (2), 167–172. https://doi.org/10.1016/0165-0270(87)90011-2.

81. Vreeland, R. F.; Atcherley, C. W.; Russell, W. S.; Xie, J. Y.; Lu, D.; Laude, N. D.; Porreca, F.; Heien, M. L. Biocompatible PEDOT:Nafion Composite Electrode Coatings for Selective Detection of Neurotransmitters in Vivo. Anal. Chem. 2015, 87 (5), 2600–2607. https://doi.org/10.1021/ac502165f.

82. Taylor, I. M.; Robbins, E. M.; Catt, K. A.; Cody, P. A.; Happe, C. L.; Cui, X. T. Enhanced Dopamine Detection Sensitivity by PEDOT/Graphene Oxide Coating on in Vivo Carbon Fiber Electrodes. 2D Mater. Biosens. Bioelectron. 2017, 89, 400–410. https://doi.org/10.1016/j.bios.2016.05.084.

83. Smith, S. K.; Lugo‐Morales, L. Z.; Tang, C.; Gosrani, S. P.; Lee, C. A.; Roberts, J. G.; Morton, S. W.; McCarty, G. S.; Khan, S. A.; Sombers, L. A. Quantitative Comparison of Immobilization Strategies for Glucose Biosensing in Real‐Time Using Fast‐Scan Cyclic Voltammetry Coupled with Carbon‐Fiber Microelectrodes. ChemPhysChem 2018, 19 (10), 1197–1204.

84. Smith, S. K.; Gosrani, S. P.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Carbon-Fiber Microbiosensor for Monitoring Rapid Lactate Fluctuations in Brain Tissue Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2018, 90 (21), 12994–12999.

85. Smith, S. K.; Lee, C. A.; Dausch, M. E.; Horman, B. M.; Patisaul, H. B.; McCarty, G. S.; Sombers, L. A. Simultaneous Voltammetric Measurements of Glucose and Dopamine Demonstrate the Coupling of Glucose Availability with Increased Metabolic Demand in the Rat Striatum. ACS Chem. Neurosci. 2017, 8 (2), 272–280. https://doi.org/10.1021/acschemneuro.6b00363.

86. Logman, M. J.; Budygin, E. A.; Gainetdinov, R. R.; Wightman, R. M. Quantitation of in Vivo Measurements with Carbon Fiber Microelectrodes. J. Neurosci. Methods 2000, 95 (2), 95–102. https://doi.org/10.1016/S0165-0270(99)00155-7.

87. Rodeberg, N. T.; Johnson, J. A.; Cameron, C. M.; Saddoris, M. P.; Carelli, R. M.; Wightman, R. M. Construction of Training Sets for Valid Calibration of in Vivo Cyclic Voltammetric Data by Principal Component Analysis. Anal. Chem. 2015, 87 (22), 11484–11491. https://doi.org/10.1021/acs.analchem.5b03222.

88. Willuhn, I.; Burgeno, L. M.; Groblewski, P. A.; Phillips, P. E. Excessive Cocaine Use Results from Decreased Phasic Dopamine Signaling in the Striatum. Nat. Neurosci. 2014, 17 (5), 704.

89. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2016, 7 (3), 349–359. https://doi.org/10.1021/acschemneuro.5b00302.

90. Michael L. A. V. Heien; Khan, A. S.; Ariansen, J. L.; Cheer, J. F.; Paul E. M. Phillips; Wassum, K. M.; Wightman, R. M.; Murray, R. W. Real-Time Measurement of Dopamine Fluctuations

66

after Cocaine in the Brain of Behaving Rats. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (29), 10023–10028.

91. Meunier, C. J.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Background Signal as an in Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2017, 8 (2), 411–419. https://doi.org/10.1021/acschemneuro.6b00325.

92. Liu, X.; Zhang, M.; Xiao, T.; Hao, J.; Li, R.; Mao, L. Protein Pretreatment of Microelectrodes Enables in Vivo Electrochemical Measurements with Easy Precalibration and Interference- Free from Proteins. Anal. Chem. 2016, 88 (14), 7238–7244. https://doi.org/10.1021/acs.analchem.6b01476.

93. Abdalla, A.; Atcherley, C. W.; Pathirathna, P.; Samaranayake, S.; Qiang, B.; Peña, E.; Morgan, S. L.; Heien, M. L.; Hashemi, P. In Vivo Ambient Serotonin Measurements at Carbon-Fiber Microelectrodes. Anal. Chem. 2017, 89 (18), 9703–9711. https://doi.org/10.1021/acs.analchem.7b01257.

94. Atcherley, C. W.; Laude, N. D.; Parent, K. L.; Heien, M. L. Fast-Scan Controlled-Adsorption Voltammetry for the Quantification of Absolute Concentrations and Adsorption Dynamics. Langmuir 2013, 29 (48), 14885–14892. https://doi.org/10.1021/la402686s.

95. Oh, Y.; Park, C.; Kim, D. H.; Shin, H.; Kang, Y. M.; DeWaele, M.; Lee, J.; Min, H.-K.; Blaha, C. D.; Bennet, K. E.; et al. Monitoring In Vivo Changes in Tonic Extracellular Dopamine Level by Charge-Balancing Multiple Waveform Fast-Scan Cyclic Voltammetry. Anal. Chem. 2016, 88 (22), 10962–10970. https://doi.org/10.1021/acs.analchem.6b02605.

96. Borman, R. P.; Wang, Y.; Nguyen, M. D.; Ganesana, M.; Lee, S. T.; Venton, B. J. Automated Algorithm for Detection of Transient Adenosine Release. ACS Chem. Neurosci. 2017, 8 (2), 386–393. https://doi.org/10.1021/acschemneuro.6b00262.

97. Lee, C. A.; Qi, L.; Amos, A.; Blanton, K.; McCarty, G. S.; Sombers, L. A. Reducing Data Density in Fast-Scan Cyclic Voltammetry Measurements of Dopamine Dynamics. J. Electrochem. Soc. 2018, 165 (12), G3042–G3050.

98. A. N. Amos; J. G. Roberts; L. Qi; L. A. Sombers; G. S. McCarty. Reducing the Sampling Rate of Biochemical Measurements Using Fast-Scan Cyclic Voltammetry for In Vivo Applications. IEEE Sens. J. 2014, 14 (9), 2975–2980. https://doi.org/10.1109/JSEN.2014.2321479.

67

CHAPTER 3

Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time

Detection of Met-Enkephalin

The following work was reprinted with permission from: Sarah E. Calhoun, Carl J. Meunier,

Christie A. Lee, Gregory S. McCarty, and Leslie A. Sombers, ACS Chemical Neuroscience, 2018,

10, 2022-2032., Copyright 2018 American Chemical Society. This work was part of a collaborative effort and a significant portion constitutes the author’s personal dissertation research; however, all results are shown to allow for a complete interpretation of the findings.

3.1 Introduction

Endogenous opioid peptides modulate a wide range of physiological functions including pain and reward processing, emotion, feeding behavior, and gastrointestinal activity by acting on the mu, delta, and kappa receptors (MOR, DOR, and KOR, respectively).1–5 These receptors are differentially expressed throughout the brain and peripheral nervous system.6–8 Mesolimbic opioid peptides are important mediators of hedonic and motivational aspects of reward processing,1–5 and aberrant opioid activity in the mesolimbic region is heavily implicated in drug addiction and drug- mediated reinforcing behaviors.9–14 However, the precise action of these opioid peptides and their interaction with mesolimbic dopamine (DA) remains ambiguous, despite nearly four decades of research. This is largely due to the paucity of techniques for direct detection of opioid peptides in situ.

The complexity of the endogenous opioid system presents many hurdles when using existing approaches to dissect the action of specific opioid peptides at specific receptors (or

68

receptor subtypes). Unlike receptors for many other small molecular transmitters, no endogenous opioid peptide family is associated exclusively with any one receptor type.15 Thus, even the complete and instantaneous blockade of a specific receptor type does not necessarily eliminate the action of a given opioid peptide. Similarly, quantification of mRNA expression and immunohistochemistry are often used to identify neurons that presumably contain the known opioid precursors: preproENK, pre-proopiomelanocortin, and pre-prodynorphin. However, cleavage of these larger molecules yields a variety of peptides that elicit a range of physiological responses.16 In fact, individual fragments from a given prohormone can generate opposing actions at postsynaptic cells in the brain.15,17 As such, identification of neurons that synthesize precursor molecules provides no information on the contribution of specific opioid peptides to brain function or whether release occurs in the somatodendritic region or at projection targets. Finally, it has even been shown that individual dopaminergic neurons in the mesolimbic circuitry can respond differentially to distinct DOR agonists.18 New analytical tools that are capable of directly monitoring rapid opioid peptide fluctuations in situ are needed to elucidate the contribution of individual opioid peptides to brain function.

Estimates of neuropeptide concentration are generally accomplished by coupling a sampling technique, such as microdialysis, to an ex situ analytical measurement of the collected fraction.19–27 This approach is widely used for monitoring small molecule neurotransmitters, but is difficult to apply to neuropeptides. These low abundance molecules are presumably released from diffuse fibers, and the probe volume is large when compared to the volume of the nerve terminal, resulting in substantial dilution as steady-state concentrations are reached. Further, microdialysis recovery efficiencies are low, particularly for “sticky” neuropeptides that readily adhere to polymeric materials (typically <5% recovery).28 These issues are significant because the

69

sensitivity of the detection method and the absolute recovery of the probe ultimately limit the temporal resolution.29 Recent studies have coupled microdialysis to liquid chromatography−mass spectrometry (LCMS)30 to detect M-ENK in dialysate from the dorsal striatum,21 globus pallidus,31 and hippocampus32 of rodents. However, this approach necessitates sampling periods of at least tens of minutes. This exacerbates the analytical challenge, as peptides are known to rapidly undergo cleavage and oxidation during sample collection.33,34

Background-subtracted, fast-scan cyclic voltammetry (FSCV) is commonly used to monitor DA fluctuations in the striatum during specific behavioral tasks related to reward seeking and consumption.35–39 We previously developed a modified-sawhorse waveform (MSW) for

FSCV that incorporates three different scan rates in each sweep to address several challenges associated with the electrochemical detection of tyrosine-containing peptides, such as M-ENK.40

Herein, we systematically investigated the waveform parameters for the electrochemical detection of M-ENK in the presence of catecholamine (CA). Incorporation of the optimized parameters increased sensitivity to M-ENK by more than 3-fold and enabled simultaneous monitoring of M-

ENK and CA at single, micrometer-scale recording sites in the rat dorsal striatum. The differential nature of this approach enables measurement of chemical fluctuations without interference from relatively stable or slowly changing electrochemical species that do not contribute to transient surges in chemical neurotransmitter. Thus, it provides the ability to reveal critical mechanistic details about rapid neuropeptide signaling and promises to considerably advance understanding of peptidergic mechanisms implicated in normal physiological function and in maladaptive behaviors such as drug addiction.

70

3.2 Methods

3.2.1 Chemicals

All chemicals were obtained with ≥95% purity (HPLC assay) and were purchased from

Sigma-Aldrich (St. Louis, MO) unless otherwise stated. The M-ENK acetate salt hydrate was obtained from LKT Laboratories (St. Paul, MN). Phosphate buffered saline (PBS; 10 mM

Na2HPO4, 138 mM NaCl, and 2.7 mM KCl) was utilized for all in vitro experiments. Adrenal slice experiments were completed in bicarbonate buffered saline (BBS; 125 mM NaCl, 26 mM

NaHCO3, 2.5 mM KCl, 2.0 mM CaCl2, 1.0 mM MgCl2, 1.3 mM NaH2PO4, 10 mM HEPES, and

10 mM glucose) saturated with 95% O2 and 5% CO2. All buffers were made using ultrapure >18.2

MΩ water (Millipore, Billerica, MA) and were adjusted to pH 7.4 with 1 M NaOH and 1 M HCl.

3.2.2 Microelectrode Fabrication

In vitro and ex vivo electrochemical experiments were carried out with glass-insulated T-

650 carbon-fiber microelectrodes (Cytec Industries, Woodland Park, NJ), fabricated as previously described.41 Briefly, a 7-μm diameter carbon fiber was aspirated into a glass capillary, and the glass was sealed around the carbon using a micropipette puller (Narishige, Tokyo, Japan). The fiber extending past the seal was cut to 100 μm. An electrical connection with the carbon fiber was established using a high ionic strength solution (4 M potassium acetate, 150 mM KCl) to backfill the capillary. A lead wire (Squires Electronics, Cornelius) was inserted to connect the electrode to custom instrumentation.

For experiments in animals, silica-insulated carbon-fiber microelectrodes were fabricated as described previously.42 Briefly, fused-silica tubing (90 μm outer diameter, 20 μm inner diameter) with a polyimide coating (Polymicro Technologies, Phoenic Arizona) was cut to 1−1.5 cm in length and placed in a bath of 70% isopropyl alcohol. A T-650 polyacrylonitrile carbon fiber

71

was inserted through the tubing and allowed to dry. An epoxy seal (McMaster Carr, Atlanta, GA) was created at one end, and an electrical connection was completed with conductive silver epoxy

(MG Chemical, Thief River Falls, MN) and a gold pin (Newark Element 14, Palatine, IL). A second layer of insulation was established around the connection using liquid insulting tape (GC

Electronics, Rockford, IL). Exposed carbon fibers were cut to 100−150 μm.

For the experiments in the anesthetized animal, an “injectrode” device was fabricated, as described previously.43 The fused silica insulation was initially cut to 3 cm (164.7 μm OD and 98.6

μm ID), and all other aspects of silica-insulated microelectrode fabrication remained unchanged.

The microelectrode was placed side by side with a guide cannula (26 GA, 11 mm from pedestal;

Plastics One, Roanoke, VA), and epoxy was used to secure them together as one device. Injection needles (33 GA, extending 1 mm beyond the guide; Plastics One) were positioned in the guide cannula for at least 1 min prior to infusions and remained in place for at least 1 min after infusions.

Reference electrodes were fabricated using a chloridized 0.25 mm diameter silver wire. A connection was made using a gold pin insulated with heat shrink. The silver wire and gold pin were positioned through a modified guide cannula stylet cap, and epoxy was used to secure it in place.

3.2.3 Electrochemical Data Acquisition In Vitro

All in vitro data were collected in a custom-built flow-injection apparatus housed within a

Faraday cage. A syringe pump (New Era Pump Systems, Inc., Wantagh, NY) was utilized to enable a continuous buffer flow of 1 mL min−1 across the working and reference electrodes. A micromanipulator (World Precision Instruments, Inc., Sarasota, FL) allowed for precise positioning of the working microelectrode into the electrochemical cell. A Ag/AgCl pellet reference electrode (World Precision Instruments, Inc., Sarasota, FL) was used to complete the

72

two-electrode cell. A six-port HPLC valve mounted on a two-position actuator controlled by a digital pneumatic solenoid valve (Valco Instruments, Houston, TX) enabled two-second bolus injections of analyte to be presented to the electrode.

Waveforms were applied at 3−20 Hz, and data were acquired at a sampling rate of 100 kHz using a custom-built instrument for potential application and current transduction (University of

North Carolina at Chapel Hill, Department of Chemistry, Electronics Facility) or a WaveNeuro

FSCV Potentiostat System (Pine Research Instrumentation, Durham, NC). High Definition Cyclic

Voltammetry software (HDCV; University of North Carolina at Chapel Hill) was used in conjunction with data acquisition cards (National Instruments, Austin TX) to control waveform output, as well as to acquire and process resulting signals, including background subtraction.

Electrodes in all experiments were electrochemically conditioned at 25 Hz until stable and then conditioned at the respective collection frequency needed per experiment for at least 10 additional min.

3.2.4 Animal Subjects and Care

Drug-naive, adult male Sprague− Dawley rats (275−300 g, Charles River Laboratories,

Raleigh, NC; n=5) were allowed to acclimate to the facility for several days. One animal received a unilateral 6-hydroxydopamine (6-OHDA) lesion of the substantia nigra by the vendor prior to receipt. Animals were individually housed on a 12:12 h light/dark cycle with free access to food and water. Animal care and use was in complete accordance with the North Carolina State

University institutional guidelines (IACUC) and the NIH’s Guide for the Care and Use of

Laboratory Animals.

73

3.2.5 Ex Vivo Adrenal Slice Preparation

Reserpine and α-MPT were prepared in 50% saline/50% dimethyl sulfoxide and in saline, respectively. Resperine (5.0 mg/kg, ip) was administered daily for 2 days prior to tissue removal and again 90 min prior to tissue removal. The α-MPT (250 mg/kg, ip) was administered 30 min prior to tissue removal. On the day of the experiment, the animal (n=1) was deeply anesthetized with urethane (1.5 g/kg ip) and rapidly decapitated. The adrenal glands were removed and embedded in 3% agarose in BBS. The agarose gel blocks were placed in ice-cold BBS, and 400-

μm thick slices were cut using a vibratome (World Precision Instruments, Sarasota, FL). Slices were allowed to rest in buffer for at least 1 h prior to placement in a recording chamber (Warner

Instruments, Hamden, CT) that was superfused with BBS buffer maintained at 34 °C. They were maintained there for at least 30 min before FSCV recordings.

Glass-insulated, carbon-fiber microelectrodes were placed approximately 100 μm below the surface of each slice with the aid of a microscope (Nikon Instruments, Inc., Melville, NY), and a Ag/AgCl pellet reference electrode (World Precision Instruments, Inc., Sarasota, FL) was placed in the tissue chamber to complete the electrochemical cell. A stimulating electrode comprised of two tungsten microelectrodes (FHC, Bowdoin, ME) was positioned 1 mm away from the working electrode. Electrical stimulations were carried out with a DS-4 Biphasic Stimulus Isolator

(Digitimer Ltd., Welwyn Garden City, England) controlled by the HDCV software. Stimulation consisted of 165 biphasic 500 μA pulses at a frequency of 165 Hz with a pulse width of 1.5 ms.

The MSW 2.0 waveform was applied at 10 Hz for this experiment.

3.2.6 In Vivo Experiments

Animals (n=3) were anesthetized with isoflurane (4% for induction and 1−3% for maintenance) and surgically prepared for electrode placement, as described.63 A heating pad

74

(Harvard Apparatus, Holliston, MA) was used to maintain body temperature at 37 °C throughout the duration of the procedure. Briefly, holes for electrodes were drilled in the skull according to coordinates from the rat brain atlas of Paxinos and Watson.44 A guide cannula for a Ag/AgCl reference electrode was placed in the contralateral forebrain (BASi Instruments, West Lafayette,

IN). The components were permanently affixed to screws in the skull with dental cement. The animals were allowed to recover for a minimum of 4 weeks prior to experiments, with daily handling. On the day of the experiment, a fresh Ag/AgCl reference electrode was inserted into the guide cannula.

For the anesthetized animal experiments, electrochemical data were collected in the dorsal striatum (+1.6 mm anteroposterior (AP); +2.0 mm mediolateral (ML) relative to bregma; −4.5 to

−5.0 mm DV from skull) using the MSW 2.0 applied at 5 Hz. On the experiment day, the animal

(n = 1) was anesthetized with isoflurane (as described above), and a heating pad was used to maintain body temperature at 37 °C. The subject received intrastriatal microinfusions (0.75 μL over 1 min) of PBS or a peptidase inhibitor cocktail (20 μM bestatin HCl added to a commercially available cocktail (Sigma-Aldrich) that contained 2 mM 4-(2-aminoethyl)benzenesulfonyl fluoride hydrochloride, 0.3 μM aprotinin, 116 μM bestatin, 14 μM E-64, 1 μM leupeptin, and 1 mM ethylenediaminetetraacetic acid).

Animals used in the awake, freely behaving experiments (n=2) were surgically prepared as described above, except silica-insulated electrodes were placed in the dorsal striatum (+1.2 mm

AP, +2.0 mm ML, and −4.5 DV). The animals were allowed to recover for a minimum of 4 weeks and were handled daily. The animal used in the fruit loops experiment received fruit loops daily after surgical recovery to prevent neophobia. On the day of the experiment, the animal was tethered and connected to a head-mounted voltammetric amplifier (current-to-voltage converter),

75

commonly referred to as a headstage (University of North Carolina at Chapel Hill, Department of

Chemistry, Electronics Facility). The headstage connects to the instrumentation via a swiveling commutator (SwivElectra; Crist Instument Co., Hagerstown, MD) to permit relatively unrestricted, free movement in the custom-built plexiglass chamber. Electrochemical data were collected using the MSW 2.0 applied at 5 Hz.

3.2.7 Statistics and Graphics

All data presented are shown as mean ± SEM, unless otherwise noted. Paired two-tailed

Student’s t tests, one-way repeated measures analysis of variance (ANOVA) with Bonferonni post hoc tests or analysis of covariance (ANCOVA) tests were used where appropriate. Significance was designated at p<0.05. Graphical depictions and statistical analyses were carried out using

GraphPad Prism 6 or 7 (GraphPad Software, Inc., La Jolla, CA) and HDCV. MATLAB R2016a was used to convert traditional CVs to mCVs for visualizing the data, and Microsoft Excel 2013 was used to calculate correlation values.

3.3 Results and Discussion

3.3.1 An Introduction to the Modified Sawhorse Waveform

The classic triangular waveform that is most frequently used in FSCV cannot be used to monitor opioid peptide fluctuations because of a plethora of issues previously described by our group.40 We overcame these limitations by designing a modified-sawhorse waveform (MSW), herein referred to as MSW 1.0 (Figure 5.1a). In the first segment of the forward scan, the potential is swept at 100 V s−1 from an accumulation potential of −0.2 V to a transition potential of +0.6 V.

The scan rate is increased to 400 V s−1 in the second segment of the forward sweep, which terminates at +1.2 V. The potential is held for 3 ms of measurements at 1.2 V before returning to

76

−0.2 at 100 V s−1.This waveform generates separate peaks for tyrosine and methionine moieties

(Figure 5.1b). Many peptides in the brain contain these residues; however, most of these are present at relatively constant concentrations over the time course of the measurement (tens of seconds).

As such, they are subtracted with the background signal. Only a peptide that contains both tyrosine and methionine and that surges in concentration over the seconds time scale is putatively identified as M-ENK with this approach.

In order to optimize and explore the full potential of the MSW for detection of M-ENK, the customizable waveform parameters were systematically evaluated. Application frequency, accumulation potential, scan rate, amperometric hold potential, and the transition potential that separates the first segment of the forward sweep from the second were varied to evaluate impact on electrochemical performance (Figure 5.1a). Individual CVs typically display potential and current on the abscissa and ordinate, respectively. As a result, current collected during the amperometric hold portion of the MSW collapses into a vertical line (Figure 5.1b, top). In this work, data collected during the amperometric hold were plotted with respect to hold time, using a scaling factor. The modified cyclic voltammograms (mCVs; Figure 5.1b, bottom) retain the conventional CV shape while also enabling visualization of currents generated during the amperometric hold period (shaded region). Chemical selectivity for M-ENK relies on the presence of distinct peaks generated in the oxidation of the tyrosine and methionine amino acid residues, which appear at ∼0.95 V and in the amperometric hold, respectively. Plotting the data in the mCV format facilitates visualization of both peaks, even enabling M-ENK to be distinguished from the closely related pentapeptide, leu-enkephalin (L-ENK), which differs by a single amino acid

(Figure 5.1b, bottom). Thus, the mCV convention is used throughout this work.

77

Figure 3.1. Introduction to the MSW. (a) A schematic of the MSW with the parameters of interest labeled. M-ENK, the target analyte, was used for waveform characterization. (b) Representative CVs (top) and mCVs (bottom) for M-ENK and L-ENK (normalized). Current collected during the amperometric period is plotted with respect to time (shaded region). M-ENK and L-ENK differ only at the C-terminus, with either a methionine or a leucine group, respectively. Both pentapeptides share the same first peak, but the presence of the second peak allows M-ENK to be visually distinguished.

3.3.2 Waveform Application Frequency

Adsorption plays a large role in the detection of many electroactive neurochemicals. For example, DA is positively charged at physiological pH, and it has been shown to concentrate at electrode surfaces during the period between voltammetric scans, when an electrode is negatively charged.45 Peptides contain many ionizable groups and other functionalities that influence

78

electrochemical detection. To investigate this, the electrochemical response to 1 μM M-ENK was recorded using waveform application frequencies of 3, 5, 10, and 20 Hz, which correspond to intersweep accumulation times of 307, 174, 94, and 24 ms, respectively (Figure 5.2a, top). All other waveform parameters were held constant. As accumulation time increased, anodic current for the oxidation of M-ENK increased (Figure 5.2a,b). This suggests that M-ENK concentrates on the carbon-fiber microelectrode surface between scans, thereby amplifying the signal generated upon . The 5 Hz waveform application frequency generated substantial current while maintaining subsecond temporal resolution. Thus, unless otherwise stated, 5 Hz was chosen as the waveform application frequency for subsequent experiments.

Figure 3.2. Waveform application frequency impacts sensitivity to MENK. (a) (top) Accumulation time as the MSW parameter under investigation. (bottom) Representative mCVs for bolus injections of 1 μM M-ENK. (b) Accumulation time (or application frequency) and current (tyrosine peak ∼0.95 V; dotted line) plotted on the abscissa and ordinate, respectively (n = 3 electrodes). Exponential line included as a guide for the eye.

3.3.3 Accumulation Potential and Scan Rate

The electrolysis of adsorption-controlled species is also impacted by accumulation potential. Thus, accumulation potential was systematically varied to investigate the impact on the voltammetric signal for M-ENK. Figure 5.3a (top) displays the waveforms used in this experiment.

Accumulation potential was varied (0.0, −0.2, −0.4, and −0.6 V); all other parameters were held constant. Figure 5.3a (middle) depicts representative mCVs for 1 μM M-ENK. Figure 5.3a

79

(bottom) presents a plot of the average peak current recorded in response to 1 μM M-ENK as a function of the accumulation potential (slope=−14.0±0.9 nA V−1, R2=0.93). Peak current clearly increased as the accumulation potential decreased (became more negative), consistent with adsorption-controlled electrolysis of M-ENK.

According to the Randles−Sevcik equation, current scales with scan rate.46 Thus, the scan rate in the potential window for tyrosine oxidation was systematically varied. Tyrosine oxidation occurs at approximately +1 V in the second segment of the forward sweep. Figure 5.3b (top) highlights how changing the scan rate in this window alters the time it takes to reach the amperometric hold potential, as well as the total duration of the waveform. Representative mCVs for 1 μM M-ENK collected using scan rates from 200 to 1200 V s−1 in this segment of the waveform are shown in Figure 5.3b (middle). Figure 5.3b (bottom) depicts the relationship between scan rate and the peak current generated by oxidation of 1 μM M-ENK (slope=0.0172 ±

0.0007 nA V−1, R2=0.99). It is important to note that increasing the scan rate shifts the peak attributed to tyrosine oxidation (at approximately +1 V), toward the amperometric hold. In fact, exceeding 800 V s−1 results in a complete loss of resolution, as the peaks that serve to identify the electroactive amino acids, tyrosine and methionine, completely overlap. Therefore, scan rates above 400 V s−1 should not be used to monitor M-ENK when using MSW 1.0.

80

Figure 3.3. Characterization of the voltammetric signal for M-ENK when varying accumulation potential (a) and scan rate (b). (top) Electrochemical waveforms investigated. The inset is an enlarged view of the region of interest. (middle) Representative mCVs collected in response to 1 μM M-ENK. (bottom) Peak anodic current generated in M-ENK oxidation (∼0.95 V) increased as the accumulation potential decreased and as the scan rate in the second segment of the forward scan increased (n=5 electrodes per parameter).

3.3.4 Amperometric Potential and Transition Potential

In an attempt to recover the characteristic two-peak signature of M-ENK that was lost with incorporation of higher scan rates (Figure 5.3b, middle), an amperometric potential of +1.3 V was investigated. In this experiment and in all subsequent analyses, unless otherwise stated, an accumulation potential of −0.4 V, a transition potential of +0.7 V, and a scan rate of 600 V s−1 in the second segment of the forward scan were employed. Figure 5.4a (top) shows the waveforms, which incorporate an amperometric potential of either +1.2 or +1.3 V. Figure 5.4a (middle) depicts representative mCVs for 1 μM M-ENK collected using both waveforms. The data clearly show 81

that extending the forward sweep to +1.3 V recovers peak resolution when using higher scan rates.

Furthermore, Figure 5.4a (bottom) demonstrates that this also improves sensitivity to M-ENK

(t(4)=3.803, p<0.05; n=5 electrodes). Based on these results, all subsequent analyses utilized an amperometric potential of +1.3 V.

Figure 3.4. Potentials selected at both nodes of the second segment of the forward scan influence the voltammetric response of M-ENK. (top) Schematic of the applied waveforms used to investigate amperometric potentials of +1.2 and +1.3 V (a) and transition potentials of +0.6, +0.65, and +0.7 V (b). (middle) Representative mCVs collected for 1 μM M-ENK. (Bottom) Increasing the amperometric potential from +1.2 to +1.3 V increased the peak signal. Increasing the transition potential significantly decreased signal amplitude. *p < 0.05, **p < 0.01, ***p < 0.001; n = 5 electrodes.

There is significant evidence that neuropeptide and small molecule neurotransmitters are copackaged within vesicles and that release can occur simultaneously.47,48 Furthermore, estimates of CA concentrations in the extracellular space, for instance, in striatum, are much higher than estimates of opioid peptide concentrations.21,49 Thus, CA molecules could easily interfere with the

82

detection of low abundance peptides. The CA neurotransmitters oxidize in the +0.5 to +0.7 V range. Ideally, CA oxidation would be completed prior to reaching the second segment of the forward scan to facilitate accurate quantification of both analytes. Transition potentials of +0.6,

+0.65, or +0.7 V (Figure 5.4b, top) were investigated with other parameters held constant. Figure

5.4b (middle) shows representative mCVs for M-ENK collected using the different waveforms.

Increasing the transition potential resulted in a decrease in the current generated for M-ENK oxidation, as shown in Figure 5.4b (bottom) (F(2,8)=95.44, p<0.0001). However, this drawback was offset by a significant benefit.

Figure 3.5. Transition potential that distinguishes the first segment ofthe forward scan from the second can be tailored to maximize sensitivity or selectivity. Representative mCVs collected with transition potentials of +0.6 V (a) or +0.7 V (b) for detection of 1 μM M-ENK, 500 nM DA, and a mixture of both species containing the same concentrations.

83

Figure 5.5 displays representative mCVs for 1 μM M-ENK, 500 nM DA (physiologically relevant for work in rodent striatum), and a mixture of the two analytes (at the same concentrations) collected with transition potentials of +0.6 and +0.7 V. The total charge contribution to the signal collected in each segment of the forward scan was examined (excluding the amperometric hold period). The results indicate that in the mixed signal, DA contributes 28.5% ± 0.9% and 18.5% ±

0.7% of the total charge collected in the second segment of the forward scan with transition potentials of +0.6 and +0.7 V, respectively (n=2 electrodes). Thus, increasing the voltage window of the initial segment (from +0.6 V to +0.7 V) allows for more complete electrolysis of DA. It should be noted that this issue is particularly important in adrenal tissue, where electrically stimulated CA release can substantially exceed 5 μM. Therefore, +0.7 V was selected as the transition potential for the subsequent studies. However, a trade-off clearly exists between sensitivity and selectivity, and determination of the most appropriate transition potential is dependent on the presence or absence of interfering chemical signals.

3.3.5 MSW 1.0 vs MSW 2.0: A Direct Comparison

Due to the coexistence of CAs and M-ENK in tissues targeted in this study (adrenal medulla and striatum),31,50–53 subsequent recordings employed a +0.7 V transition potential. Given all of the above characterizations, a waveform referred to as MSW 2.0 was employed for tissue measurements using a 5 Hz application frequency. The potential was swept from −0.4 V to +0.7

V at 100 V s−1 before a faster sweep to +1.3 V at 600 V s −1. The potential was then held at +1.3

V for 3 ms before returning to −0.4 V at 100 V s−1 (Figure 5.6a). MSW 2.0 improves upon the previous waveform (MSW 1.0). It results in defined and distinct tyrosine and methionine peaks vital for chemical selectivity, as well as a greater than 3-fold increase in sensitivity (Figure 5.6b,c, one-way ANCOVA, F(1,4)=304.9, p<0.0001; n=5 electrodes).

84

Figure 3.6. Improved detection of M-ENK with MSW 2.0. (a) Agraphic comparison of the two waveforms. (b) Representative mCVs collected for 1 μM M-ENK using MSW 1.040 and MSW 2.0. (c) A direct comparison of calibration plots for M-ENK using these waveforms. ***p<0.001; n=5 electrodes.

3.3.6 Simultaneous Measurements of CA and M-ENK Fluctuations in Living Adrenal Tissue

Endogenous peptides are critically involved in numerous physiological functions that promote survival, including the response to stress.6,54 For example, the electrically excitable chromaffin cells that make up the adrenal medulla secrete several neuropeptides (including M-

ENK) and relatively high concentrations of CAs (dopamine, norepinephrine, and epinephrine) during fight-or-flight behavior.6,55,56 For measurements in the adrenal medulla, rats were pretreated with α-methyl-DLtyrosine methyl ester hydrochloride (α-MPT) and reserpine to inhibit the

85

synthesis and vesicular packaging of CAs, respectively.57,58 This pretreatment effectively reduced

CA content, in order to facilitate detection of M-ENK. Without it, the CA signal dominated the color plot, exceeding concentrations of ∼5 μM (data not shown). Figure 5.7a displays representative color plots (top) and mCVs (bottom) for standards of 750 nM NE, 750 nM DA, and

500 nM M-ENK collected in vitro using MSW 2.0. Figure 5.7b (top) displays a representative color plot of electrically evoked CA release and a second analyte, putatively identified as M-ENK, collected in an adrenal slice (data collected at 10 Hz). A voltammogram directly following the stimulation was extracted and compared with a voltammogram for a M-ENK standard (Figure

5.7b, bottom). The voltammetric signature for CA is clearly evident, and there is good agreement between the normalized mCVs in the potential region where M-ENK is detected (0.7−1.3 V,

R2=0.83), providing electrochemical evidence for the identification of M-ENK. These data suggest that CA and M-ENK are released on a similar time scale upon electrical stimulation of adrenal tissue and establish the potential for MSW 2.0 in addressing a broad range of fundamental questions regarding endogenous opioid dynamics in live tissue.

86

Figure 3.7. Simultaneous monitoring of M-ENK and CA dynamics in an adrenal slice preparation with MSW 2.0. (a) Representative data for standards of 750 nM DA and 500 nM M-ENK, and (b) M-ENK and CA released following electrical stimulation (administered at the time indicated by the red dashed line). (top) Color plots of raw voltammetric data. (bottom) mCVs extracted from the color plots at the time indicated by the white dashed line. There is good agreement between the normalized mCVs in the potential range where M-ENK is evident (to the right of the dashed line, 0.7−1.3 V, R2=0.83).

3.3.7 Simultaneous Measurements of CA and M-ENK in the Dorsal Striatum

Enkephalins modulate motor output regions in the brain31,59–61 and nuclei involved in food intake,21,62 and they are involved in the integration of limbic information in the dorsal striatum.63

Approximately 90−95% of the cellular makeup of this region consists of medium spiny neurons, approximately half of which are known to express the DA D2 receptor and various opioid receptors and to contain enkephalin.64,65 An infusion of 1 μM M-ENK was delivered to the local vicinity of the electrode in the striatum (within ∼300−500 μm) to demonstrate the applicability of MSW 2.0 for voltammetric measurements of neuropeptide in brain tissue. Figure 5.8a demonstrates that

87

MSW 2.0 can clearly detect the infused M-ENK. Interestingly, this signal was followed by an increase in extracellular CA, and the extracted mCV (inset) exhibits the defining characteristics of both CA and the M-ENK standard. Next, PBS was locally microinfused, with no observable neurochemical effects (Figure 5.8b). Finally, a cocktail of peptidase inhibitors was microinfused into the vicinity of the recording site (Figure 5.8c). This solution contained 20 μM bestatin added to a commercially available cocktail of 2 mM 4-(2-aminoethyl)- benzenesulfonyl fluoride hydrochloride, 0.3 μM aprotinin, 116 μM bestatin, 14 μM E-64, 1 μM leupeptin, and 1 mM ethylenediaminetetraacetic acid. This manipulation should locally increase the concentration and extracellular lifetime of a variety of neuropeptides. The data show a voltammetric signal consistent with that recorded for infusion of exogenous M-ENK, as demonstrated by the Pearson’s correlation coefficient quantifying the covariance of the extracted mCVs (inset, R=0.88).

Figure 3.8. M-ENK recorded in the dorsal striatum of an anesthetized rat. (top) Representative color plots collected during microinfusion of M-ENK (a), PBS (b), or a cocktail of enkephalinase inhibitors (c). Infusion is marked with the orange bar on the concentration vs time plots (bottom). Inset mCVs were extracted at the time point indicated by the corresponding white dashed lines. Microinfusion of PBS did not result in any observable neurochemical changes, but local infusion of the protease inhibitor cocktail resulted in voltammograms that correlated with those collected after infusion of exogenous M-ENK (R = 0.88).

88

When assessing a new strategy to detect endogenous molecules in the brain, an important step is to validate the approach by selectively modulating the signal using known pharmacology.

However, to date, drugs to unambiguously and selectively manipulate the concentration of endogenous opioid peptides in the extracellular space do not exist. However, 6- OHDA lesioned animals exhibit increased pre-proENK mRNA expression.66 Furthermore, because the manipulation destroys the majority of the DA neurons in the substantia nigra that project to the dorsal striatum, there is also less interference from endogenous CA. Thus, putative M-ENK signals were monitored in the dorsal striatum of an intact and a 6-OHDA lesioned rat (both awake and freely moving).

Figure 5.9a (left) presents a representative color plot collected in the intact animal at rest.

The voltammograms (inset) are indicative of spontaneous, dynamic DA fluctuations, and there are also signals that correlate well with the voltammogram for exogenous M-ENK infused into the striatum (Figure 5.8a, R=0.85). By contrast, the representative color plot collected in the 6-OHDA lesioned animal (Figure 5.9a, right) contains spontaneous signals consistent with voltammograms for a MENK standard (Figure 5.7, R=0.84), with no evidence of a phasic DA signal.

Previous work has demonstrated a role for surges in ENK in the rat anterior dorsomedial striatum in the modulation of food-motivated behavior (albeit on the tens-of-minutes time scale).21

Thus, voltammetric data were collected in anterior dorsomedial striatum of a male rat in response to unexpected palatable food reward (fruit loops, Figure 5.9b). Little electrochemical signal was recorded under baseline conditions, until the subject was presented with food reward. The subsequent bouts of food interaction and consumption were monitored by a trained observer while electrochemistry was recorded. The voltammograms correlate with those for the M-ENK standard

(Figure 5.7, R=0.80−0.88), consistent with a role for ENK surges in motivation to consume reward.

89

Taken together, the data in Figure 5.9 suggest that endogenous M-ENK can be monitored in the striatum of awake animals; however, further investigation is needed to unambiguously identify M-

ENK and to clarify the role that opioid peptides play in motivated behaviors.

Figure 3.9. Neurochemical fluctuations recorded in the dorsomedial striatum of awake, freely behaving rats. Representative color plots are shown, with concentration vs time traces below. Inset mCVs were extracted at the time point indicated by the white dashed lines. (a) Voltammograms that correlate with those collected after infusion of exogenous M-ENK into striatal tissue were evident in the intact striatum (left, R=0.85), but no CA signal was evident in the 6-OHDA lesioned animal (right). These voltammograms correlate with the M-ENK standard (R=0.84). (b) A voltammetric signal that correlates with M-ENK fluctuations was recorded in response to the presentation (left) and consumption (middle and right) of unexpected food reward (R=0.80−0.88).

90

3.4 Conclusions

Direct, real-time measurements of opioid neuropeptides in live tissue and behaving animals will provide an improved understanding of their actions in both the peripheral and central nervous systems. This work takes a significant step toward that goal. The results demonstrate that the MSW allows for direct detection of tyrosine-containing peptides, such as M-ENK, in live tissue. Through a systematic characterization of the waveform parameters, we have enhanced selectivity and sensitivity for M-ENK and demonstrated the simultaneous release of M-ENK and CA in both adrenal and striatal tissue. Importantly, this approach is not limited to endogenous opioids, and we have demonstrated how voltammetric waveforms can be customized to enhance detection of specific target analytes, broadly speaking. Overall, this work lays a foundation that can ultimately enable researchers to make direct, real-time measurements of tyrosine-containing endogenous peptides. It has the potential to enable investigation of questions concerning the specific conditions required for peptide release, peptidergic lifetime in the extracellular space, and how peptidergic flux is paired with specific behaviors. Further, the approach promises to provide valuable and unprecedented information to inform the development of therapeutic treatments of a myriad of physiological dysfunctions.

91

3.5 References

1. Johnson, P. I.; Stellar, J. R.; Paul, A. D. Regional Reward Differences within the Ventral Pallidum Are Revealed by Microinjections of a Mu Opiate Receptor Agonist. Neuropharmacology 1993, 32 (12), 1305–1314. https://doi.org/10.1016/0028-3908(93)90025- X.

2. Smith, K. S.; Berridge, K. C. The Ventral Pallidum and Hedonic Reward: Neurochemical Maps of Sucrose “Liking” and Food Intake. J. Neurosci. 2005, 25 (38), 8637. https://doi.org/10.1523/JNEUROSCI.1902-05.2005.

3. Smith, K. S.; Tindell, A. J.; Aldridge, J. W.; Berridge, K. C. Ventral Pallidum Roles in Reward and Motivation. Behav. Brain Res. 2009, 196 (2), 155–167. https://doi.org/10.1016/j.bbr.2008.09.038.

4. Kelley, A. E.; Bakshi, V. P.; Haber, S. N.; Steininger, T. L.; Will, M. J.; Zhang, M. Opioid Modulation of Taste Hedonics within the Ventral Striatum. Proc. 2001 Meet. Soc. Study Ingestive Behav. SSIB 2002, 76 (3), 365–377. https://doi.org/10.1016/S0031-9384(02)00751- 5.

5. McBride, W. J.; Murphy, J. M.; Ikemoto, S. Localization of Brain Reinforcement Mechanisms: Intracranial Self-Administration and Intracranial Place-Conditioning Studies. Behav. Brain Res. 1999, 101 (2), 129–152. https://doi.org/10.1016/S0166-4328(99)00022-4.

6. Akil, H.; Watson, S. J.; Young, E.; Lewis, M. E.; Khachaturian, H.; Walker, J. M. Endogenous Opioids: Biology and Function. Annu. Rev. Neurosci. 1984, 7 (1), 223–255. https://doi.org/10.1146/annurev.ne.07.030184.001255.

7. Mansour, A.; Burke, S.; Pavlic, R. J.; Akil, H.; Watson, S. J. Immunohistochemical Localization of the Cloned Κ1 Receptor in the Rat CNS and Pituitary. Neuroscience 1996, 71 (3), 671–690. https://doi.org/10.1016/0306-4522(95)00464-5.

8. Le Merrer, J.; Becker, J. A. J.; Befort, K.; Kieffer, B. L. Reward Processing by the Opioid System in the Brain. Physiol. Rev. 2009, 89 (4), 1379–1412. https://doi.org/10.1152/physrev.00005.2009.

9. Gerrits, M. A.; van Ree, J.; Patkina, N.; Zvartau, E. Opioid Blockade Attenuates Acquisition and Expression of Cocaine-Induced Place Preference Conditioning in Rats. Psychopharmacology (Berl.) 1995, 119 (1), 92–98.

10. Mitchell, J. M.; Bergren, L. J.; Chen, K. S.; Rowbotham, M. C.; Fields, H. L. Naltrexone Aversion and Treatment Efficacy Are Greatest in Humans and Rats That Actively Consume High Levels of Alcohol. Neurobiol. Dis. 2009, 33 (1), 72–80. https://doi.org/10.1016/j.nbd.2008.09.018.

11. Mitchell, J. M.; O’Neil, J. P.; Janabi, M.; Marks, S. M.; Jagust, W. J.; Fields, H. L. Alcohol Consumption Induces Endogenous Opioid Release in the Human Orbitofrontal Cortex and

92

Nucleus Accumbens. Sci. Transl. Med. 2012, 4 (116), 116ra6. https://doi.org/10.1126/scitranslmed.3002902.

12. Olive, M. F.; Maidment, N. T. Repeated Heroin Administration Increases Extracellular Opioid Peptide-like Immunoreactivity in the Globus Pallidus/Ventral Pallidum of Freely Moving Rats. Psychopharmacology (Berl.) 1998, 139 (3), 251–254.

13. Skoubis, P. D.; Maidment, N. T. Blockade of Ventral Pallidal Opioid Receptors Induces a Conditioned Place Aversion and Attenuates Acquisition of Cocaine Place Preference in the Rat. Neuroscience 2003, 119 (1), 241–249. https://doi.org/10.1016/S0306-4522(03)00121-0.

14. Tang, X.-C.; McFarland, K.; Cagle, S.; Kalivas, P. W. Cocaine-Induced Reinstatement Requires Endogenous Stimulation of μ-Opioid Receptors in the Ventral Pallidum. J. Neurosci. 2005, 25 (18), 4512. https://doi.org/10.1523/JNEUROSCI.0685-05.2005.

15. Mansour, A.; Hoversten, M. T.; Taylor, L. P.; Watson, S. J.; Akil, H. The Cloned μ, δ and κ Receptors and Their Endogenous Ligands: Evidence for Two Opioid Peptide Recognition Cores. Brain Res. 1995, 700 (1), 89–98. https://doi.org/10.1016/0006-8993(95)00928-J.

16. Breslin, M. B.; Lindberg, I.; Benjannet, S.; Mathis, J. P.; Lazure, C.; Seidah, N. G. Differential Processing of Proenkephalin by Prohormone Convertases 1(3) and 2 and Furin. J. Biol. Chem. 1993, 268 (36), 27084–27093.

17. Lewis, R. Enkephalin Biosynthesis in the Adrenal Medulla. Adv. Biochem. Psychopharmacol. 1982, 33, 167–174.

18. Margolis, E. B.; Fujita, W.; Devi, L. A.; Fields, H. L. Two Delta Opioid Receptor Subtypes Are Functional in Single Ventral Tegmental Area Neurons, and Can Interact with the Mu Opioid Receptor. Neuropharmacology 2017, 123, 420–432. https://doi.org/10.1016/j.neuropharm.2017.06.019.

19. Freed, A. L.; Cooper, J. D.; Davies, M. I.; Lunte, S. M. Investigation of the Metabolism of Substance P in Rat Striatum by Microdialysis Sampling and Capillary Electrophoresis with Laser-Induced Fluorescence Detection. J. Neurosci. Methods 2001, 109 (1), 23–29. https://doi.org/10.1016/S0165-0270(01)00397-1.

20. Zhou, Y.; Mabrouk, O. S.; Kennedy, R. T. Rapid Preconcentration for Liquid Chromatography–Mass Spectrometry Assay of Trace Level Neuropeptides. J. Am. Soc. Mass Spectrom. 2013, 24 (11), 1700–1709.

21. DiFeliceantonio, A. G.; Mabrouk, O. S.; Kennedy, R. T.; Berridge, K. C. Enkephalin Surges in Dorsal Neostriatum as a Signal to Eat. Curr. Biol. 2012, 22 (20), 1918–1924. https://doi.org/10.1016/j.cub.2012.08.014.

22. Kostel, K. L.; Lunte, S. M. Evaluation of Capillary Electrophoresis with Post-Column Derivatization and Laser-Induced Fluorescence Detection for the Determination of Substance P and Its Metabolites. J. Chromatogr. B. Biomed. Sci. App. 1997, 695 (1), 27–38.

93

23. Haskins, W. E.; Wang, Z.; Watson, C. J.; Rostand, R. R.; Witowski, S. R.; Powell, D. H.; Kennedy, R. T. Capillary LC−MS2 at the Attomole Level for Monitoring and Discovering Endogenous Peptides in Microdialysis Samples Collected in Vivo. Anal. Chem. 2001, 73 (21), 5005–5014. https://doi.org/10.1021/ac010774d.

24. Desiderio, D. M.; Zhu, X. Quantitative Analysis of Methionine Enkephalin and β-Endorphin in the Pituitary by Liquid Secondary Ion Mass Spectrometry and Tandem Mass Spectrometry. J. Chromatogr. A 1998, 794 (1–2), 85–96.

25. Sinnaeve, B. A.; Storme, M. L.; Van Bocxlaer, J. F. Capillary Liquid Chromatography and Tandem Mass Spectrometry for the Quantification of Enkephalins in Cerebrospinal Fluid. J. Sep. Sci. 2005, 28 (14), 1779–1784. https://doi.org/10.1002/jssc.200500114.

26. Dawson, R.; Steves, J. P.; Lorden, J. F.; Oparil, S. Reverse-Phase Separation and Electrochemical Detection of Neuropeptides. Peptides 1985, 6 (6), 1173–1178. https://doi.org/10.1016/0196-9781(85)90446-2.

27. Shen, H.; Lada, M. W.; Kennedy, R. T. Monitoring of Met-Enkephalin in Vivo with 5-Min Temporal Resolution Using Microdialysis Sampling and Capillary Liquid Chromatography with Electrochemical Detection. J. Chromatogr. B. Biomed. Sci. App. 1997, 704 (1–2), 43–52.

28. Roy, M. C.; Ikimura, K.; Nishino, H.; Naito, T. A High Recovery Microsampling Device Based on a Microdialysis Probe for Peptide Sampling. Anal. Biochem. 2010, 399 (2), 305–307. https://doi.org/10.1016/j.ab.2009.12.036.

29. Li, Q.; Zubieta, J.-K.; Kennedy, R. T. Practical Aspects of in Vivo Detection of Neuropeptides by Microdialysis Coupled Off-Line to Capillary LC with Multistage MS. Anal. Chem. 2009, 81 (6), 2242–2250. https://doi.org/10.1021/ac802391b.

30. Baseski, H. M.; Watson, C. J.; Cellar, N. A.; Shackman, J. G.; Kennedy, R. T. Capillary Liquid Chromatography with MS3 for the Determination of Enkephalins in Microdialysis Samples from the Striatum of Anesthetized and Freely–Moving Rats. J. Mass Spectrom. 2005, 40 (2), 146–153. https://doi.org/10.1002/jms.733.

31. Mabrouk, O. S.; Li, Q.; Song, P.; Kennedy, R. T. Microdialysis and Mass Spectrometric Monitoring of Dopamine and Enkephalins in the Globus Pallidus Reveal Reciprocal Interactions That Regulate Movement. J. Neurochem. 2011, 118 (1), 24–33. https://doi.org/10.1111/j.1471-4159.2011.07293.x.

32. Wilson, R. E.; Jaquins-Gerstl, A.; Weber, S. G. On-Column Dimethylation with Capillary Liquid Chromatography-Tandem Mass Spectrometry for Online Determination of Neuropeptides in Rat Brain Microdialysate. Anal. Chem. 2018, 90 (7), 4561–4568. https://doi.org/10.1021/acs.analchem.7b04965.

33. Bergès, J.; de Oliveira, P.; Fourré, I.; Houée-Levin, C. The One-Electron Reduction Potential of Methionine-Containing Peptides Depends on the Sequence. J. Phys. Chem. B 2012, 116 (31), 9352–9362. https://doi.org/10.1021/jp304741e.

94

34. Devi, L.; Goldstein, A. Conversion of Leumorphin (Dynorphin B-29) to Dynorphin B and Dynorphin B-14 by Thiol Protease Activity. J. Neurochem. 1986, 47 (1), 154–157. https://doi.org/10.1111/j.1471-4159.1986.tb02843.x.

35. Cameron, C. M.; Wightman, R. M.; Carelli, R. M. Dynamics of Rapid Dopamine Release in the Nucleus Accumbens during Goal-Directed Behaviors for Cocaine versus Natural Rewards. Neuropharmacology 2014, 86, 319–328. https://doi.org/10.1016/j.neuropharm.2014.08.006.

36. Willuhn, I.; Wanat, M. J.; Clark, J. J.; Phillips, P. E. Dopamine Signaling in the Nucleus Accumbens of Animals Self-Administering Drugs of Abuse. In Behavioral neuroscience of drug addiction; Springer, 2010; pp 29–71.

37. Phillips, P. E. M.; Stuber, G. D.; Heien, M. L. A. V.; Wightman, R. M.; Carelli, R. M. Subsecond Dopamine Release Promotes Cocaine Seeking. Nature 2003, 422 (6932), 614–618. https://doi.org/10.1038/nature01476.

38. Roitman, M. F.; Stuber, G. D.; Phillips, P. E. M.; Wightman, R. M.; Carelli, R. M. Dopamine Operates as a Subsecond Modulator of Food Seeking. J. Neurosci. 2004, 24 (6), 1265. https://doi.org/10.1523/JNEUROSCI.3823-03.2004.

39. Stuber, G. D.; Roitman, M. F.; Phillips, P. E. M.; Carelli, R. M.; Wightman, R. M. Rapid Dopamine Signaling in the Nucleus Accumbens during Contingent and Noncontingent Cocaine Administration. Neuropsychopharmacology 2005, 30 (5), 853–863. https://doi.org/10.1038/sj.npp.1300619.

40. Schmidt, A. C.; Dunaway, L. E.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Multiple Scan Rate Voltammetry for Selective Quantification of Real-Time Enkephalin Dynamics. Anal. Chem. 2014, 86 (15), 7806–7812. https://doi.org/10.1021/ac501725u.

41. Roberts, J. G.; Lugo-Morales, L. Z.; Loziuk, P. L.; Sombers, L. A. Real-Time Chemical Measurements of Dopamine Release in the Brain. In Dopamine; Springer, 2013; pp 275–294.

42. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V.; et al. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat. Methods 2010, 7 (2), 126–129. https://doi.org/10.1038/nmeth.1412.

43. Wilson, L. R.; Panda, S.; Schmidt, A. C.; Sombers, L. A. Selective and Mechanically Robust Sensors for Electrochemical Measurements of Real-Time Hydrogen Peroxide Dynamics in Vivo. Anal. Chem. 2018, 90 (1), 888–895. https://doi.org/10.1021/acs.analchem.7b03770.

44. Paxinos, G.; Watson, C. The Rat Brain in Stereotaxic Coordinates: Hard Cover Edition; Elsevier, 2006.

45. Heien, M. L. A. V.; Phillips, P. E. M.; Stuber, G. D.; Seipel, A. T.; Wightman, R. M. Overoxidation of Carbon-Fiber Microelectrodes Enhances Dopamine Adsorption and Increases Sensitivity. Analyst 2003, 128 (12), 1413–1419. https://doi.org/10.1039/B307024G.

95

46. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

47. Nguyen, T. T.; Léan, A. D. Nonadrenergic Modulation by Clonidine of the Cosecretion of Catecholamines and Enkephalins in Adrenal Chromaffin Cells. Can. J. Physiol. Pharmacol. 1987, 65 (5), 823–827. https://doi.org/10.1139/y87-132.

48. Kawagoe, K. T.; Garris, P. A.; Wiedemann, D. J.; Wightman, R. M. Regulation of Transient Dopamine Concentration Gradients in the Microenvironment Surrounding Nerve Terminals in the Rat Striatum. Neuroscience 1992, 51 (1), 55–64. https://doi.org/10.1016/0306- 4522(92)90470-M.

49. Howe, M. W.; Tierney, P. L.; Sandberg, S. G.; Phillips, P. E. M.; Graybiel, A. M. Prolonged Dopamine Signalling in Striatum Signals Proximity and Value of Distant Rewards. Nature 2013, 500 (7464), 575–579. https://doi.org/10.1038/nature12475.

50. Klein, R. L.; Wilson, S. P.; Dzielak, D. J.; Yang, W.-H.; Viveros, O. H. Opioid Peptides and Noradrenaline Co-Exist in Large Dense-Cored Vesicles from Sympathetic Nerve. Neuroscience 1982, 7 (9), 2255–2261. https://doi.org/10.1016/0306-4522(82)90135-X.

51. Kelle, A. E.; Stinus, L.; Iversen, S. D. Interactions between D-Ala-Met-Enkephalin, A10 Dopaminergic Neurones, and Spontaneous Behaviour in the Rat. Behav. Brain Res. 1980, 1 (1), 3–24. https://doi.org/10.1016/0166-4328(80)90043-1.

52. Kalivas, P. W.; Widerlöv, E.; Stanley, D.; Breese, G.; Prange, A. J. Enkephalin Action on the Mesolimbic System: A Dopamine-Dependent and a Dopamine-Independent Increase in Locomotor Activity. J. Pharmacol. Exp. Ther. 1983, 227 (1), 229–237.

53. Johnson, S.; North, R. Opioids Excite Dopamine Neurons by Hyperpolarization of Local Interneurons. J. Neurosci. 1992, 12 (2), 483. https://doi.org/10.1523/JNEUROSCI.12-02- 00483.1992.

54. Pasternak, G. W.; Pan, Y.-X. Mu Opioids and Their Receptors: Evolution of a Concept. Pharmacol. Rev. 2013, 65 (4), 1257. https://doi.org/10.1124/pr.112.007138.

55. Toneff, T.; Funkelstein, L.; Mosier, C.; Abagyan, A.; Ziegler, M.; Hook, V. Beta-Amyloid Peptides Undergo Regulated Co-Secretion with Neuropeptide and Catecholamine Neurotransmitters. Peptides 2013, 46, 126–135. https://doi.org/10.1016/j.peptides.2013.04.020.

56. Verhofstad, A.; Coupland, R.; Parker, T.; Goldstein, M. Immunohistochemical and Biochemical Study on the Development of the Noradrenaline-and Adrenaline-Storing Cells of the Adrenal Medulla of the Rat. Cell Tissue Res. 1985, 242 (2), 233–243.

57. Martínez-Olivares, R.; Villanueva, I.; Racotta, R.; Piñón, M. Depletion and Recovery of Catecholamines in Several Organs of Rats Treated with Reserpine. Auton. Neurosci. 2006, 128 (1), 64–69. https://doi.org/10.1016/j.autneu.2006.04.004.

96

58. Kulagina, N. V.; Michael, A. C. Monitoring Hydrogen Peroxide in the Extracellular Space of the Brain with Amperometric Microsensors. Anal. Chem. 2003, 75 (18), 4875–4881. https://doi.org/10.1021/ac034573g.

59. Koshimizu, Y.; Wu, S.-X.; Unzai, T.; Hioki, H.; Sonomura, T.; Nakamura, K. C.; Fujiyama, F.; Kaneko, T. Paucity of Enkephalin Production in Neostriatal Striosomal Neurons: Analysis with Preproenkephalin–Green Fluorescent Protein Transgenic Mice. Eur. J. Neurosci. 2008, 28 (10), 2053–2064. https://doi.org/10.1111/j.1460-9568.2008.06502.x.

60. Henry, B.; Crossman, A. R.; Brotchie, J. M. Effect of Repeatedl-DOPA, Bromocriptine, or Lisuride Administration on Preproenkephalin-A and Preproenkephalin-B MRNA Levels in the Striatum of the 6-Hydroxydopamine-Lesioned Rat. Exp. Neurol. 1999, 155 (2), 204–220. https://doi.org/10.1006/exnr.1998.6996.

61. Henry, B.; Duty, S.; Fox, S. H.; Crossman, A. R.; Brotchie, J. M. Increased Striatal Pre- Proenkephalin B Expression Is Associated with Dyskinesia in Parkinson’s Disease. Exp. Neurol. 2003, 183 (2), 458–468. https://doi.org/10.1016/S0014-4886(03)00064-5.

62. Pratt, W. E.; Kelley, A. E. Striatal Muscarinic Receptor Antagonism Reduces 24-h Food Intake in Association with Decreased Preproenkephalin Gene Expression. Eur. J. Neurosci. 2005, 22 (12), 3229–3240. https://doi.org/10.1111/j.1460-9568.2005.04489.x.

63. Banghart, M. R.; Neufeld, S. Q.; Wong, N. C.; Sabatini, B. L. Enkephalin Disinhibits Mu Opioid Receptor-Rich Striatal Patches via Delta Opioid Receptors. Neuron 2015, 88 (6), 1227– 1239. https://doi.org/10.1016/j.neuron.2015.11.010.

64. Kemp, J. M.; Powell, T. P. S. The Connexions of the Striatum and Globus Pallidus: Synthesis and Speculation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1971, 262 (845), 441–457.

65. Gerfen, C. R.; Engber, T. M.; Mahan, L. C.; Susel, Z.; Chase, T. N.; Monsma, F. J.; Sibley, D. R. D$_1$ and D$_2$ Dopamine Receptor-Regulated Gene Expression of Striatonigral and Striatopallidal Neurons. Science 1990, 250 (4986), 1429–1432.

66. Gerfen, C.; McGinty, J.; Young, W. Dopamine Differentially Regulates Dynorphin, Substance P, and Enkephalin Expression in Striatal Neurons: In Situ Hybridization Histochemical Analysis. J. Neurosci. 1991, 11 (4), 1016. https://doi.org/10.1523/JNEUROSCI.11-04- 01016.1991.

97

CHAPTER 4

Electrochemical Selectivity Achieved Using a Double Voltammetric Waveform and Partial

Least Squares Regression: Differentiating Endogenous Hydrogen Peroxide Fluctuations

from Shifts in pH

The following work was reprinted with permission from: Carl J. Meunier, Edwin C. Mitchell,

James G. Roberts, Jonathan V. Toups, Gregory S. McCarty, and Leslie A. Sombers, Analytical

Chemistry, 2017, 90, 1767-1776., Copyright 2017 American Chemical Society. Supplemental information is found in Appendix A.

4.1 Introduction

Hydrogen peroxide (H2O2) is an important, membrane permeable, reactive oxygen species

(ROS) present in the brain that is commonly regarded as a potential toxin, as it can react to form

1 aggressive hydroxyl radicals that can irreversibly alter DNA, lipids, and protein structures. H2O2 also serves as an indicator of the upstream production of more detrimental ROS by mitochondria, where elevated concentrations are indicative of cellular malfunction. Pathological oxidative stress leads to neuronal degeneration that is symptomatic of several neurodegenerative disease states and

2–6 aging. Importantly, H2O2 also serves as a subcellular signaling molecule in redox signaling pathways important to normal cell function, as well as a modulator of dopamine transmission in

7–9 the dorsolateral striatum. Finally, the electrochemical detection of H2O2 is also of great interest because H2O2 can be exploited to serve as a reporter molecule in the indirect detection of nonelectroactive species such as glucose, glutamate, and acetylcholine using oxidase-based biosensors.10–14

98

H2O2 oxidation is electrocatalyzed at platinum, which is widely used to construct electrodes for these applications.15,16 However, recent work from our group has demonstrated the utility of fast-scan cyclic voltammetry (FSCV) coupled with carbon-fiber microelectrodes for the detection

17 of H2O2 at a biosensor surface, as well as for the detection of endogenous H2O2 fluctuations in brain slice and in vivo.14,18 This approach offers both sensitivity and some selectivity, making it ideal for monitoring molecular fluctuations in vivo. By employing a waveform extended to +1.4 V

(vs Ag/AgCl), the carbon surface is oxidized, rendering it more sensitive to many analytes,

17,19–21 including H2O2. However, endogenous H2O2 events are often accompanied by other chemical events that are detected by the electrode, including small shifts in pH, that render quantification of in vivo data difficult. Much work has gone into developing paradigms to improve selectivity in voltammetric experiments. These range from the incorporation of modified data collection strategies to the incorporation of statistical paradigms to deconvolute complex signals.

For instance, electrodes can be coated with ion-selective polymers,22–24 a double electrochemical waveform can be employed,25 or the voltammetric sweep can be combined with a small potential step to help remove unwanted signal.26 Alternatively, voltammograms for individual chemical contributors to complex FSCV signals can be resolved using multivariate data analysis paradigms, such as principal component regression (PCR), elastic net regression,27 or multivariate curve resolution.28 One example that has been generally accepted by the field is PCR, a combination of principal component analysis with inverse least-squares regression.29–37 PCR utilizes information collected across the entire potential window for a training set comprised of cyclic voltammograms

(CVs) of known analytes (and concentrations) to determine principal components (PCs), or basis eigenvectors, that describe the variance in the training data.34,37 Then, concentration dynamics for these species can be predicted for unknown data by projecting the data onto the PCs to extrapolate

99

the contribution of each analyte contained in the training set. This has been shown to work well for discriminating DA from other signals, including ΔpH, in data collected in vivo.27,29,36 Yet, PCR struggles when two analytical signals have similar sources of variance; i.e. have similar

34,35 voltammetric features. Such is the case for H2O2 and ΔpH signals. The voltammetric signature for H2O2 consists of a single, well-defined peak at ∼1.3 V. By contrast, that for ΔpH exhibits current across the entire potential window, including a peak at ∼1.3 V. To further confound the situation, the ΔpH signal varies widely across electrodes due to a dependence on the chemical composition of the recording environment and the surface state of the carbon, which contains many oxygen-containing functional groups.38,39 Thus, PCR is not always adequate to reliably distinguish current contributions from H2O2 and ΔpH, as these are not reliably defined in factor space.

Herein, we employ a double voltammetric waveform to more accurately quantify H2O2 fluctuations that occur at the electrode surface simultaneously with small shifts in pH. To accomplish this, a smaller waveform in which H2O2 is electrochemically silent is scanned immediately prior to the standard, larger waveform that collects current contributions from both

H2O2 and ΔpH. PLSR is a logical and easy to understand variant of PCR that is commonly used in analytical chemistry due to readily available software and advantages in predictive power in some applications.40–45 It is used herein to reduce the dimensionality of calibration sets of CVs collected using both waveforms. Calibration sets collected with the small waveform (sWF) contain predictor variables to describe response variables that are evident in calibration sets collected with the large waveform (lWF). Thus, PLSR was used to exploit information on ΔpH collected with the small waveform and to predict and subtract the current response due to ΔpH from voltammograms collected with the larger waveform, without altering information related to H2O2 dynamics. The validity of the DW-PLSR was assessed using 5-fold cross validation and cumulative variance.

100

Finally, the model was used to subtract ΔpH (and dopamine, DA) contributions from complex signals recorded in the dorsal striatum, clearly revealing endogenous and stimulated H2O2 and fluctuations. The model reliably discriminates between H2O2 and ΔpH signals both in vitro and in vivo. Thus, DW-PLSR promises to serve as a reliable tool to increase confidence in the quantification of endogenous H2O2 dynamics.

4.2 Experimental Section

4.2.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. In vitro electrochemical experiments were performed in TRIS buffered saline (15 mM 2-amino-2-hydroxymethylpropane-

1,3- diol, 3.25 mM KCl, 1.20 mM CaCl2, 1.2 mM MgCl2, 2 mM Na2SO4, 1.25 mM NaH2PO4, and

145 mM NaCl) adjusted to pH 7.4 with 1 M NaOH and 1 M HCl. All aqueous solutions were made using ultrapure 18.2 MΩ water

4.2.2 Microelectrode Fabrication

Cylindrical carbon-fiber microelectrodes were fabricated using T-650 carbon fibers (Cytec

Industries, Inc., Woodland Park, NJ) as previously described.46 In short, a single carbon fiber was aspirated into a glass capillary, a glass seal was created using a micropipette puller (Narishige,

Tokyo, Japan), and the fiber extending past the seal was cut to 100 μm in length. An electrical connection to the carbon fiber was established with conductive silver epoxy and a wire lead. For experiments in freely moving animals, electrodes were insulated in polyimide-coated, fused-silica tubing (Polymicro Technologies, Phoenix, AZ; 164.7 μm outer diameter/ 98.6 μm inner diameter), as previously described.47 Briefly, the tubing was cut to 4 cm in length and placed in an aqueous

101

70% isopropyl alcohol solution. A T-650 carbon fiber was inserted and allowed to dry. An epoxy seal was created at one end, and an electrical connection to the fiber was completed via conductive silver epoxy. Electrodes were then cured at 100 °C for 20 min. Finally, the exposed carbon fiber was cut to length.

4.2.3 Flow-Injection System

Working electrodes were positioned in a custom electrochemical flow cell using a micromanipulator (World Precision Instruments, Sarasota, FL). A syringe pump (New Era Pump

Systems, Wantagh, NY) supplied a continuous buffer stream (1 mL min−1) across the working and reference electrodes (Ag/AgCl pellet, World Precision Instruments, Inc., Sarasota, FL). Two- second bolus injections of analyte were accomplished with a six-port HPLC valve mounted on a two-position actuator controlled by a digital pneumatic solenoid (Valco Instruments, Houston,

TX). The entire apparatus was enclosed in a custom-built, grounded Faraday cage.

4.2.4 Data Acquisition

The double triangular waveform consisted of a small cyclic waveform ranging from −0.4

V to +0.8 V, and a larger cyclic waveform ranging from −0.4 to +1.4 V, applied using a scan rate of 400 V s−1 and separated by a 12 ms hold at −0.4. The waveform was applied with an application frequency of 10 Hz and a sampling frequency of ∼37 kHz, using a custom-built instrument for potential application and current transduction (Universal Electrochemistry Instrument, University of North Carolina at Chapel Hill, Department of Chemistry, Electronics Facility). This resulted in small and large CVs consisting of 223 and 334 data points, respectively. High Definition Cyclic

Voltammetry software (HDCV, University of North Carolina at Chapel Hill) was used for waveform output, signal processing, and data analysis, in conjunction with data acquisition cards

(National Instruments, Austin TX) used for measuring current and synchronizing the

102

electrochemical cell with the flow-injection system. Analog filtering was accomplished with a 2- pole Sallen-Key, low-pass filter at 2 kHz. Data were then digitally filtered with a first-order low- pass Bessel filter; cutoff frequencies of 2 and 100 kHz in HDCV.

4.2.5 Animal Experiments

Animal care and use was in accordance with North Carolina State University Institutional

Animal Care and Use Committee (IACUC) guidelines. Male Sprague−Dawley rats (250−300 g) were purchased with or without a unilateral 6-hydroxydopamine (6-OHDA) lesion (Charles River

Laboratories, Raleigh, NC), individually housed on a 12:12 h light/dark cycle with free access to food and water, and allowed to acclimate to the facility for several days. On surgery day, animals designated for freely moving experiments were anesthetized with 4% isoflurane during an induction period, then isoflurane was adjusted to 2.5% for electrode placement. A silica-insulated working electrode was placed in the dorsal striatum of the lesioned hemisphere (+1.2 mm AP, +

2.0 mm ML, + 5.0 mm DV relative to bregma) using a flat-skull stereotaxic apparatus (Kopf

Instrumentation; Tujunga, CA), as previously described.46 A Ag/AgCl reference electrode was placed in the contralateral forebrain. Electrodes were permanently affixed with dental cement, and animals were allowed to recover for 4 weeks. On experiment day, the electrodes were connected to a head-mounted amplifier and electrical connections passed through a commutator that allowed free movement in an open field chamber.

Animals designated for acute experiments were deeply anesthetized with urethane (1.5 g kg−1), a heating pad (Harvard Apparatus, Holliston, MA) was used to maintain body temperature at 37 °C, and electrodes were implanted as described above with the following exceptions: the carbon-fiber working electrode positioned in the dorsal striatum was insulated with glass, and a bipolar stimulation electrode was implanted in the medial forebrain bundle (+2.7 mm AP, + 1.7

103

mm ML, and −7.5 mm DV, relative to bregma). Electrical stimulations consisted of trains of 100 biphasic, 300 μA pulses at a frequency of 60 Hz with a pulse width of 2 ms. All working electrodes were conditioned at 30 Hz for 15 min with the double waveform prior to data collection with the same waveform.

4.2.6 Data Processing and Analysis

The first 10 CVs (1 s) of each 30 s data file were averaged to obtain the background voltammogram used to produce background-subtracted data for that file. For in vitro experiments, the maximum oxidation current was evaluated for each analyte (H2O2 and pH). Voltammograms

(and corresponding small waveform voltammograms) containing the maximal current, as well as two CVs immediately before and after, were averaged to produce a set of CVs for calibration, and also training data for PLSR. In vivo data were examined manually in HDCV using the tool described in to extract CVs in the same manner as the training set.

PLSR was performed using the “plsregress” function in MATLAB R2014a (Mathworks,

Natick, MA), along with the SIMPLS algorithm to center the data. This approach was computationally efficient, and served to maximize covariance between the predictor and response variables.48 Mean-squared error was determined using a 5-fold cross validation. All statistical analyses and graphical depiction of the data were carried out using GraphPad Prism 6 (GraphPad

Software, Inc., La Jolla, CA) and MATLAB.

4.3 Results and Discussion

4.3.1. FSCV Detection of H2O2 and ΔpH

The standard approach to the voltammetric detection of H2O2 using FSCV at a carbon-fiber microelectrode employs a cyclic triangular waveform that scans from a −0.4 V holding potential

104

to a + 1.4 V switching potential and back to −0.4 V at a constant scan-rate of 400 V s−1 , applied at a frequency of 10 Hz.17,18,49 This generates a large but stable background signal composed predominately of nonfaradaic capacitive (charging) current,50 contributions arising from oxidation and reduction of groups inherent to the electrode surface,39,51 and faradaic contributions from electroactive species that are statically present in the extracellular space.46 The background current can be subtracted to reveal faradaic current generated in response to an electroactive species transiently present in the immediate vicinity of the electrode. This current is plotted versus the applied potential to create a CV, where peak position(s) serve as a chemical identifier and the intensity is used for quantification. Due to the volume of data collected, CVs are concatenated with respect to time to create a “color plot”, which displays three dimensions of data in a simplified manner. Figure 4.1A contains representative color plots and inset CVs for bolus injections of H2O2

(left), basic ΔpH (middle), and a mixture of H2O2 and basic ΔpH (right). The overall shape of the

CVs for H2O2 and ΔpH differ, as evidenced by the patterns visible in the color plots. Importantly, shifts in pH generate current across the entire potential window.38 However, the peak current is generated at ∼1.3 V (reverse scan) for both H2O2 and ΔpH, convolving identification of the species contributing to the mixed voltammogram and hindering reliable calibration.

105

Figure 4.1. (A) Representative color plots and CVs (inset) for of 80 μM H2O2 (left), + 0.25 basic ΔpH (middle) and a mixture of both (right). Time and potential are plotted as the ordinate and abscissa, respectively. Current is depicted in false color, as indicated by the scale bar. CVs are plotted with positive potentials to the right. Arrows indicate the direction of the scan. Representative CVs (left) and a calibration plot (right) for (B) H2O2, and (C) ΔpH (calibration carried out using the peak at ∼0.6 V). Data are presented as mean ± standard deviation (n=5 electrodes). All data presented were collected using the larger scan of the double waveform.

In voltammetry, peak current is directly proportional to analyte concentration, and the slope of a regression line through the calibration data can be used as a calibration factor to describe sensitivity to a given analyte. Figure 4.1B,C displays representative CVs (left) collected in vitro

17,18,30,38 for physiologically relevant concentrations of H2O2 and ΔpH, as well as calibration plots

(right) where the slopes of the regression models are 0.131 ± 0.001 nA/μM for H2O2, and 53 ± 3 nA/ΔpH unit. However, univariate calibration is not adequate for quantification when multiple unresolved signals overlap, which is often the case when making measurements in complex in vivo environments. The current produced by individual analytes is additive at each potential with

106

contributions proportional to the concentration of each species, as long as there is no or competing surface interaction for the constituents. If the ΔpH were known, then it would be straightforward to determine and subtract its contribution at a specific potential.

However, if the contributions of both ΔpH and H2O2 are unknown, then it becomes impossible to accurately assign values for each component using univariate calibration.

4.3.2 Principal Component Regression: Complications in Distinguishing H2O2 and ΔpH

To use PCR with FSCV, principal component analysis (PCA) is performed by compiling a training set composed of CVs for each analyte expected to contribute to the signal, with intensities encompassing the relevant magnitudes.33,34 PCA calculates a ranked set of principal component vectors describing the entirety of the training set, with the first component capturing the maximum variance inherent to the training set; and remaining components each describing the remaining variance, always orthogonal to the previous component.37,52 Then scores, or scaled projections of the training CVs onto the individual components, can be regressed against concentration. This allows signal contributions from different analytes, whose voltammograms differ in factor (component) space, to be individually quantified. A useful tool to evaluate the efficacy of PCR models is the Cook’s score (distance) plot.30,34,53 A good PCR model distinguishes each analyte well in factor space, and will generally result when the training set consists of analytes with voltammograms that are sufficiently unique. Such is the case for a training set composed of

H2O2 and DA CVs. Example Cook’s plots for two PCR training sets are displayed in Figure 4.2A

(H2O2 and DA) and B (H2O2 and ΔpH). The data shown in Figure 4.2A were recorded using a conventional catecholamine waveform spanning −0.4 to +1.3 V with a scan rate of 400 V s−1. For two analytes, one expects to observe two score vectors, or linear groups of data. An ideal model would produce components that are analyte-pure, with sample scores resulting in orthogonal

107

(uncorrelated) linear groups lying on the coordinates, rather than within the quadrants. This would indicate that each component describes the variance of one individual analyte.52–54 This PCR model for H2O2 and DA results in nearly orthogonal (dashed line) score vector components (∠83°), as shown in Figure 4.2A. DA’s score vector lies close to the abscissa, indicating that most of the variance is captured by PC1, and H2O2’s score vector spreads predominately along the ordinate.

In both cases, the adjacent PC describes some of the variance, signifying that the voltammograms for these species share at least some common variance. Figure 4.2C plots the two PC loadings versus the applied potential. PC1 (black) is similar to dopamine’s CV, and PC2 (red) resembles the CV for H2O2. This indicates that the model sufficiently separated the analytes in factor space.

Figure 4.2. Principal component regression is insufficient to resolve individual contributions to the voltammetric signal from ΔpH and H2O2. (A) Score plot for a training set comprised of voltammograms for H2O2 (red) and DA (blue) collected using the catecholamine waveform (−0.4 to +1.3 V, 400 V s−1). The sample scores on the first and second components are plotted on the abscissa and ordinate, respectively. The angle between the two score vectors (red and blue lines) is inset. (B) Score plot for a training set comprised of voltammograms for H2O2 (red) and ΔpH −1 (blue) collected using the H2O2 waveform (−0.4 to +1.4 V, 400 V s ). (C,D) Principal component loadings (ordinate) plotted vs potential (abscissa) for the first (black) and second (red) components in (A) and (B), respectively.

108

However, if voltammograms for two analytes are not sufficiently dissimilar (i.e., there are similar sources of variance), it is likely that the two analytes cannot be accurately resolved using

PCR. Significant deviations from orthogonal sample score vectors indicate multicollinearity, and

54 thus highly correlated factors. In the case of voltammograms for ΔpH and H2O2 recorded using the standard H2O2 waveform spanning −0.4 to +1.4 V, at least one component contains significant contributions from both analytes. Figure 4.2B shows that the sample scores for both analytes lie well within a quadrant, indicating that the first two PCs contain substantial contributions from both species. Additionally, the angle between the two score vectors (∠60°) deviates significantly from orthogonality, indicating highly correlated factors. Indeed, the first PC (Figure 4.2D, black) resembles the voltammogram for ΔpH; it is not visually discernible that it contains contributions from H2O2. By contrast, the second PC (red) contains characteristics of both analytes. Ultimately this confounds reliable separation of contributions from ΔpH from those of H2O2 using PCR. In an effort to more accurately distinguish and quantify voltammetric signals comprised of both ΔpH and H2O2, alternate strategies were investigated.

4.3.3 Double Waveform-Partial Least Squares Regression Model

A model to reliably identify, predict, and subtract current contributions due to local shifts in pH requires gathering information specific to the ΔpH signal. To achieve this, we employed a novel voltammetric waveform that incorporates two triangular sweeps of differing magnitude, as depicted in Figure 4.3A. Figure 4.3B displays representative CVs collected using the small (left) and large (right) waveforms for H2O2 (top), ΔpH (middle), and a mixture of the two analytes

−1 (bottom). The sWF scans from −0.4 V to +0.8 V and back at 400 V s , as H2O2 is electrochemically silent in this potential window and ΔpH generates substantial current. Thus, the sWF captures information specific to ΔpH that can serve as a predictor of the ΔpH contribution to

109

voltammograms collected 12 ms later with the lWF, which scans to +1.4 at 400 V s−1, generating

CVs that contain contributions from both analytes.

Data collected using the double waveform were used to produce a multivariate calibration model to determine the ΔpH contribution to voltammograms subsequently collected using the lWF. PCR and PLSR are similar in that they both utilize predictor components to describe the observed data, but they differ in the manner in which these components are constructed. PCR creates PCs to describe variance in the training set voltammograms (predictor) without considering the response (analyte concentration), and is therefore considered to be an unsupervised dimensionality reduction technique.52,55 In the context of the double waveform described herein, predictor-sWF and response-lWF voltammograms differ in length and number of data points).

Therefore, PCR is not suitable, as the size of the sWF predictor differs from that of the lWF- response. By contrast, PLSR is a supervised dimensionality reduction technique that projects both the sWF-predictor and lWFresponse variables onto a new vector space to find components that maximize the covariance of the projected structures.55,56 This generally allows PLSR models to describe the training data more efficiently with fewer components (than PCR), resulting in an output prediction model that is often (but not always) more reliable when adjusted.40,42,56

When using PLSR, the total number of components produced by the model will be one less than the size of the training set. However, only a fraction of the components will describe the “real” signal (PCs), with the remaining components describing noise.37,52 Supervised predictive modeling techniques, such as PLSR, generally employ cross-validation approaches to determine the proper number of PCs to retain.41 As such, k-fold cross validation was used to determine the number of

PCs retained in the PLSR analysis. This was done by training the model following removal of a fraction of the training set. The removed data were then used to compute the mean-square error

110

(MSE) of the predicted lWF-response. Finally, cumulative variance was used as a figure of merit to assess the level of noise in the data.44,57 A training set was constructed of ΔpH (including

ΔpH=0.0) voltammograms collected with both the sWF and the corresponding lWF using five electrodes in vitro. Representative training set data are depicted in Figure 4.3B (middle). Figure

4.3C displays the 5-fold cross validation and cumulative percent variance for both the sWF predictor (black and gray) and lWF response (red and pink). The MSE associated with both the sWF-predictor and the lWF-response variables show no significant improvement in predictive power by retaining greater than five components. Therefore, five components were retained for this in vitro model, which capture an excess of 99.5% of the training set variance. To further validate the model, sWF voltammograms from every electrode for ΔpH, H2O2, and mixtures

(+0.25 ΔpH with five H2O2 concentrations) were input into the model. The PLSR-predicted ΔpH signal was subtracted from voltammograms collected using the lWF. ΔpH-subtracted CVs for a single electrode are displayed in Figure 4.3D for H2O2 (top), ΔpH (middle), and known mixtures

(bottom). Importantly, the model effectively removed ΔpH contributions to signals that contained a known contribution from ΔpH. The coefficient of determination (R2) between the training CVs for H2O2 collected using the lWF and the ΔpH-subtracted H2O2 and mixture CVs are 0.988 ± 0.008 and 0.96 ± 0.02, respectively (Figure 4.3D, inset text). In fact, calibration data for H2O2 (Figure

4.1B), and for ΔpH-subtracted H2O2 and mixture samples were not significantly different (data not shown, slope: F(2,39)=0.025, p=0.98, intercept: F(2,41)=0.86, p=0.43). Furthermore, the model effectively removed ΔpH signals that were ∼20× larger than the magnitude of the smallest H2O2 contributions. This is important because H2O2 signals recorded in vivo are often smaller in magnitude than voltammetric signals for local shifts in pH. Thus, the DW-PLSR model should serve as a useful tool in vivo.

111

Figure 4.3. A double waveform-partial least-squares regression (DW-PLSR) model. (A) Double waveform containing a small triangular waveform that scans from −0.4 to +0.8 V and back at 400 V s−1 , and a larger one that extends to +1.4 V. (B) Representative CVs collected with the small (left) and large (right) waveforms for H2O2 (top), basic ΔpH (middle), and known mixtures of these two analytes. Note that H2O2 (top) does not generate significant current using the sWF. (C) Cross validation and cumulative percent variance. The number of components retained is plotted on the ordinate. Left: Cumulative variance for the sWF-predictor (black) and lWF-response (red) CVs. A dashed line indicates the position of 99.5% variance. Right: average mean-squared error (nA2 /CV) for the sWF-predictor (gray) and lWF-response (pink) CVs. 5-fold cross-validation was done by training the model without data from each electrode. (D) ΔpH-subtracted CVs for a single electrode, using a five-component PLSR model for H2O2 (top), ΔpH (middle), and known mixtures (bottom). Inset are coefficients of determination (n=15 CVs) for the H2O2 (B, top-right) and the ΔpH-subtracted CVs collected using the lWF.

112

4.3.4 Construction of an In Vivo Training Set

Whenever possible, a training set or calibration should be generated in a recording environment that matches the experimental environment so as to eliminate matrix effects, or changes in the analytical signal caused by anything in the sample other than the analyte.58

Constructing a training set for in vitro studies is straightforward, as the experimenter knows what constitutes each sample. However, constructing an appropriate training set of data collected in vivo presents a bigger challenge, as the chemical complexity of living tissue is immense, dynamic, and has not been accurately replicated. In traditional in vivo studies, DA-specific voltammograms have been obtained by electrical stimulation of dopaminergic afferents to evoke DA release in the vicinity of the electrode, generating a well-characterized CV that is easily recognized.59 Electrical stimulation also generates a hemodynamic response that includes a large contribution from a local pH change,60 but also likely contains contributions from other electroactive substances including

H2O2, O2, and adenosine. Thus, a method to aid in selection of CVs collected in vivo that best represent pure shifts in pH was devised by examining relationships inherent to the ΔpH data collected in vitro. Figure 4.4A displays correlation plots (n=5 electrodes) for current collected at any two potentials of the lWF for ΔpH (top) and H2O2 (bottom). Red indicates that currents collected at two distinct potentials simultaneously increase in intensity (disregarding sign) with increasing analyte concentration, and thus are correlated; blue indicates regions in the voltammogram in which there is little to no correlation. ΔpH generates a strong signal at potentials less than +0.8 V, and there is a high correlation (R2≈0.99) between currents collected at +0.6 V

(forward-scan) and +1.3 V (reverse scan). Little to no current is generated at +0.6 V in the voltammetric detection of H2O2, and thus the correlation between currents collected at these

2 specific potentials is much weaker for H2O2 (R ≈0.5). As such, the relationship between currents

113

collected at +0.6 V (forwardscan) and +1.3 V (reverse scan) can be used to differentiate voltammograms resulting from a local shift in pH from those due to local fluctuations in H2O2.

To demonstrate this, in vitro data were collected for standards of ΔpH, H2O2, and mixtures of these two species, and the currents recorded at +0.6 V and +1.3 V were normalized to the maximum currents recorded at these potentials (across all samples). Figure 4.4B displays the results for the ΔpH voltammograms. The current at both potentials increases as ΔpH increases, resulting in a ΔpH-only vector distributed about the ordinate and abscissa. Figure 4.4C displays the results for all samples: ΔpH (red), H2O2 (blue), H2O2/+0.1 ΔpH (dark purple), and H2O2/+0.2

ΔpH (light purple). In short, the data group nicely. Increases increases along the ordinate indicate pure shifts in pH. By contrast, H2O2 is electrochemically silent below +0.6 V, and thus has little spread along the ordinate. Therefore, these relationships can be used to facilitate selection of ΔpH

CVs collected in vivo, and to exclude voltammograms that most likely result from mixtures of

ΔpH and H2O2.

114

Figure 4.4. Training set selection tool. (A) Correlation plots describe current collected at any two potentials in the voltammetric detection of ΔpH (top) and H2O2 (bottom). Large waveform potentials are plotted on the ordinate and abscissa. Red coloring indicates that currents collected at two potentials are correlated, i.e., they simultaneously increase in intensity with increasing analyte concentration. Blue indicates little to no correlation. For clarity, example current traces (linear voltammograms) for each analyte are shown along the ordinate and abscissa. Stars indicate positions of significantly contrasting correlation discussed in the text. (B) Normalized current at +0.6 and +1.3 V extracted from the ΔpH calibration data for one electrode. The magnitude of the pH shift is labeled next to each data cluster. Inset is a CV with arrows that indicate the potentials investigated. (C) Displays the normalized currents collected at +0.6 V and +1.3 data for standards of ΔpH (red), H2O2 (blue), and two sets of mixtures of these two analytes (purple).

4.3.5 Applying the DW-PLSR Model to In Vivo Data

The DW-PLSR model was used to evaluate H2O2 fluctuations in the dorsal striatum of an awake, freely moving, 6-OHDA lesioned rat. This is a common model for Parkinson’s disease, and data were recorded on the lesioned side which contains very few dopaminergic terminals.61,62

Local shifts in pH can generate significant current and, as a result, H2O2 fluctuations can be easily masked. Thus, it is necessary to reliably subtract ΔpH contributions in order to more accurately assess underlying H2O2 dynamics. The in vivo data were investigated (as described above) to select

115

appropriate ΔpH voltammograms to serve as “standards” in the training set. Figure 4.5A shows voltammograms for acidic ΔpH (left; 12 CVs) and blanks (right; 3 CVs) collected using both the sWF (top) and lWF (bottom). These CVs and their inversions were used for PLSR, as inverting the voltammogram for an acidic pH shift results in the voltammogram for a basic shift in pH.

Figure 4.5. In vivo training set for PLSR model. (A) sWF (top) and corresponding lWF (bottom) CVs were extracted from in vivo data for acidic ΔpH (left; 12 CVs), and blank (right; 3 CVs) signals. The inversion (not shown) of each training CV was also included to produce a training set of 30 CVs. (B) Cross validation and cumulative percent variance. The number of components retained is plotted on the ordinate. Left: Cumulative variance for the sWF-predictor (black) and lWF-response (red) CVs. A dashed line indicates the position of 99.5% variance. Right: average mean-squared error (nA2 /CV) for the sWF-predictor (gray) and lWF-response (pink) CVs. [5- Fold cross validation was performed by randomly dividing up the training databases].

Figure 4.5B displays the cross-validation results used to determine the number of PCs to retain. The MSE for both the sWF predictor (gray) and lWF-response (pink) variables were not improved when greater than five components were retained. Consequently, five components were kept for this in vivo model. This five-component model also contained in excess of 99.5% of the

116

variance inherent to the original in vivo data set. As further evidence of proper variable selection, the sWF-predictor and lWF-response loadings for the first five components are plotted versus the corresponding waveform potentials in Figure S1A of the Supporting Information (SI). Loadings for the first PC (top) resemble the ΔpH training data. Loadings for PCs 2−5 (bottom) for the sWF predictor (left) appear to be predominately noise, whereas those for the lWF-response (right) contain prominent features near +1.3 V. Additionally, the residual signal for each training CV, not captured by the first five components, appears to be noise and does not contain features reminiscent of either analyte (Figure S1B).

Figure 4.6 summarizes the results obtained for ΔpH subtraction from data collected in vivo.

Figure 4.6A,B (top) shows color plots for 30 s recordings - for simplicity, only the data collected using the lWF are presented. The data presented in Figure 4.6A contain information on H2O2 fluctuations, as indicated by the blue/green/purple colors at ∼+1.3 V. The data presented in Figure

4.6B are similar, but also contain significant contributions from ΔpH - note the different current scale. The PLSR-predicted, lWF response (describing the ΔpH contribution) is presented in the middle row, and the ΔpH-subtracted color plots are displayed at the bottom. In Figure 4.6A, where no interfering signal was evident, the raw data and ΔpH-subtracted color plots are visually similar.

They describe the H2O2 activity, which is minimal in this recording. In Figure 4.6B, where an interfering ΔpH signal was evident, the raw data and ΔpH-subtracted color plots are substantially different, highlighting the ability to model and subtract contributions from ΔpH and retain information selectively describing H2O2 dynamics. Voltammograms extracted from the raw data

(black), the PLSR-predicted ΔpH response (blue), and the ΔpH-subtracted data (red) are compared in Figures 4.6C,D. The raw and ΔpH-subtracted voltammograms in Figure 4.6C are similar, as the model determined that there is not significant interference from ΔpH. By contrast, the PLSR-

117

predicted response in Figure 4.6B is similar to the raw voltammogram, which contains a significant

ΔpH contribution. The ΔpH-subtracted voltammogram for the time point indicated by the dashed vertical line reveals a well-defined, H2O2 signal. Similarly, the corrected color plot provides for easy visualization and quantification of the H2O2 dynamics across this 30 s period.

Figure 4.6. Application of DW-PLSR model to in vivo data collected in the striatum of a freely moving, 6-OHDA lesioned rat. (A,B) Color plots for two different 30 s recordings. For simplicity, only the data collected using the large waveform are shown. Raw data (top), the PLSR ΔpH- predicted response (middle), and the ΔpH-subtracted data (bottom) are shown. Colormap scaling is shown above each set. (C,D) Voltammograms extracted from color plots in (A) and (B), respectively. Dashed white lines indicate the position of the extracted CVs.

When working in the striatum, DA is commonly monitored. Thus, it is important to demonstrate that the DW-PLSR model is effective for elucidating H2O2 dynamics, even in the presence of DA. Striatal recordings from an intact, anesthetized rat were used to construct a

ΔpH/DA training set for PLSR, as done in Figure 4.6. The training set CVs (8 ΔpH, 8 DA, and 3

118

blank CVs), PC loadings, cumulative variance, and cross-validation data are displayed in Figure

S-2. The cross-validation shows that the MSE begins to level off when greater than four components are retained. Thus, four PCs were retained for analysis of the data shown in Figure

4.7. It is evident from the PC loadings that PC1 is predominately indicative of a ΔpH signal, PC2 contains features indicative of both ΔpH and DA, and PC3−5 resemble noise. Figure 4.7A displays the raw data (top), the PLSR-predicted ΔpH/DA response (middle), and ΔpH/DAsubtracted color plots (bottom) for a 60 s recording. The ΔpH/ DA-subtracted data retain features clearly indicative of H2O2, and the PLSR-predicted, lWF response contains features inherent to ΔpH and DA. Figure

4.7B displays a comparison of CVs extracted immediately following an electrical stimulation

(top), as well as from a later time point (bottom). DA and contributions from ΔpH are evident following the stimulation (at the time indicated by the white line in panel A), and the model clarifies the H2O2 component at this time point (top, panel B). A more substantial ΔpH signal is observed ∼30 s later (black line, panel A). The DW-PLSR can also predict and subtract this component from the data, clarifying the H2O2 information (bottom, panel B). Thus, the DW-PLSR model is applicable for the analysis of complex data containing multiple chemical contributions.

119

Figure 4.7. Application of the DW-PLSR model to data collected in the striatum of an intact rat. (A) Color plot for a 60 s recording, the time of electrical stimulation is indicated by the arrow. Raw data (top), the PLSR-predicted ΔpH/DA response (middle), and the ΔpH/DA-subtracted data (bottom) are shown. (B) CVs extracted from color plots in (A), at the positions indicated by the vertical dashed lines.

4.4 Conclusion

In all electrochemical studies carried out in vivo, chemical selectivity is a primary concern.

Herein, we demonstrate that PLSR can be combined with a double voltammetric waveform to provide an effective means of quantitatively resolving the electrochemical contribution of an analyte in a complex mixture. This can be done even in the presence of an interferent that also exhibits substantial redox activity at the same potential. By using a sWF that captures information specific to the primary interferents (in this case ΔpH and DA) as a predictor variable, the contribution of these species in the analytical region of interest can be subtracted nonsubjectively from the lWF signal, revealing the underlying signal for the analyte of interest. A major advantage of this technique is that it simplifies time consuming experimental protocols for building in vivo

120

training sets - researchers are no longer required to search in vivo recordings for voltammograms representative of all analytes present; only interferent CVs are necessary. As such, the inclusion of a stimulating electrode to elicit an analyte response may no longer be necessary, simplifying surgical procedures and limiting tissue damage. Importantly, the model does not subtract information in the absence of the interfering species. Finally, the potential limits of the double waveform can be easily adjusted for broad applicability. This data analysis tool should prove useful in distinguishing the contribution of interferents from that of any other analyte with redox activity that appears outside the potential window encompassed by the smaller waveform.

121

4.5 References

1. Halliwell, B. Oxidative Stress and Neurodegeneration: Where Are We Now? J. Neurochem. 2006, 97 (6), 1634–1658. https://doi.org/10.1111/j.1471-4159.2006.03907.x.

2. Adam-Vizi, V. Production of Reactive Oxygen Species in Brain Mitochondria: Contribution by Electron Transport Chain and Non–Electron Transport Chain Sources. Antioxid. Redox Signal. 2005, 7 (9–10), 1140–1149. https://doi.org/10.1089/ars.2005.7.1140.

3. Fridovich, I. Superoxide Anion Radical (O·̄2), Superoxide Dismutases, and Related Matters. J. Biol. Chem. 1997, 272 (30), 18515–18517. https://doi.org/10.1074/jbc.272.30.18515.

4. Mattson, M. P.; Liu, D. Energetics and Oxidative Stress in Synaptic Plasticity and Neurodegenerative Disorders. Neuromolecular Med. 2002, 2 (2), 215–231.

5. Zhang, Y.; Dawson, V. L.; Dawson, T. M. Oxidative Stress and Genetics in the Pathogenesis of Parkinson’s Disease. Neurobiol. Dis. 2000, 7 (4), 240–250. https://doi.org/10.1006/nbdi.2000.0319.

6. Kishida, K. T.; Klann, E. Sources and Targets of Reactive Oxygen Species in Synaptic Plasticity and Memory. Antioxid. Redox Signal. 2006, 9 (2), 233–244. https://doi.org/10.1089/ars.2007.9.233.

7. Bao, L.; Avshalumov, M. V.; Patel, J. C.; Lee, C. R.; Miller, E. W.; Chang, C. J.; Rice, M. E. Mitochondria Are the Source of Hydrogen Peroxide for Dynamic Brain-Cell Signaling. J. Neurosci. 2009, 29 (28), 9002. https://doi.org/10.1523/JNEUROSCI.1706-09.2009.

8. Stone, J. R.; Yang, S. Hydrogen Peroxide: A Signaling Messenger. Antioxid. Redox Signal. 2006, 8 (3–4), 243–270. https://doi.org/10.1089/ars.2006.8.243.

9. Rhee, S. G. H2O2, a Necessary Evil for Cell Signaling. Science 2006, 312 (5782), 1882. https://doi.org/10.1126/science.1130481.

10. Burmeister, J. J.; Gerhardt, G. A. Self-Referencing Ceramic-Based Multisite Microelectrodes for the Detection and Elimination of Interferences from the Measurement of l-Glutamate and Other Analytes. Anal. Chem. 2001, 73 (5), 1037–1042. https://doi.org/10.1021/ac0010429.

11. Parikh, V.; Kozak, R.; Martinez, V.; Sarter, M. Prefrontal Acetylcholine Release Controls Cue Detection on Multiple Timescales. Neuron 2007, 56 (1), 141–154. https://doi.org/10.1016/j.neuron.2007.08.025.

12. Wassum, M. K.; Tolosa, M. V.; Wang, J.; Walker, E.; Monbouquette, G. H.; Maidment, T. N. Silicon Wafer-Based Platinum Microelectrode Array Biosensor for Near Real-Time Measurement of Glutamate in Vivo. Sensors 2008, 8 (8). https://doi.org/10.3390/s8085023.

13. Wilson, G. S.; Johnson, M. A. In-Vivo Electrochemistry: What Can We Learn about Living Systems? Chem. Rev. 2008, 108 (7), 2462–2481. https://doi.org/10.1021/cr068082i.

122

14. Smith, S. K.; Lee, C. A.; Dausch, M. E.; Horman, B. M.; Patisaul, H. B.; McCarty, G. S.; Sombers, L. A. Simultaneous Voltammetric Measurements of Glucose and Dopamine Demonstrate the Coupling of Glucose Availability with Increased Metabolic Demand in the Rat Striatum. ACS Chem. Neurosci. 2017, 8 (2), 272–280. https://doi.org/10.1021/acschemneuro.6b00363.

15. Zhang, Y.; Wilson, G. S. Electrochemical Oxidation of H2O2 on Pt and Pt + Ir Electrodes in Physiological Buffer and Its Applicability to H2O2-Based Biosensors. Int. J. Devoted Asp. Electrode Kinet. Interfacial Struct. Prop. Electrolytes Colloid Biol. Electrochem. 1993, 345 (1), 253–271. https://doi.org/10.1016/0022-0728(93)80483-X.

16. Hall, S. B.; Khudaish, E. A.; Hart, A. L. Electrochemical Oxidation of Hydrogen Peroxide at Platinum Electrodes. Part III: Effect of Temperature. Electrochimica Acta 1999, 44 (14), 2455– 2462. https://doi.org/10.1016/S0013-4686(98)00369-7.

17. Sanford, A. L.; Morton, S. W.; Whitehouse, K. L.; Oara, H. M.; Lugo-Morales, L. Z.; Roberts, J. G.; Sombers, L. A. Voltammetric Detection of Hydrogen Peroxide at Carbon Fiber Microelectrodes. Anal. Chem. 2010, 82 (12), 5205–5210. https://doi.org/10.1021/ac100536s.

18. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2013, 4 (5), 782–789. https://doi.org/10.1021/cn4000499.

19. Aoki, K.; Ishida, M.; Tokuda, K.; Hasebe, K. Electrode Kinetics of the Oxidation of Hydrogen Peroxide at Pretreated Glassy Carbon and Carbon Fiber Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1988, 251 (1), 63–71. https://doi.org/10.1016/0022-0728(88)80385- 1.

20. Nowall, W. B.; Kuhr, W. G. Detection of Hydrogen Peroxide and Other Molecules of Biologicl Importance at an Electrocataltic Surface on a Carbon Fiber Microelectrode. Electroanalysis 1997, 9 (2), 102–109. https://doi.org/10.1002/elan.1140090203.

21. Roberts, J. G.; Toups, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode Calibration Strategy for Voltammetric Measurements In Vivo. Anal. Chem. 2013, 85 (23), 11568–11575. https://doi.org/10.1021/ac402884n.

22. Vreeland, R. F.; Atcherley, C. W.; Russell, W. S.; Xie, J. Y.; Lu, D.; Laude, N. D.; Porreca, F.; Heien, M. L. Biocompatible PEDOT:Nafion Composite Electrode Coatings for Selective Detection of Neurotransmitters in Vivo. Anal. Chem. 2015, 87 (5), 2600–2607. https://doi.org/10.1021/ac502165f.

23. Szentirmay, M. N.; Martin, C. R. Ion-Exchange Selectivity of Nafion Films on Electrode Surfaces. Anal. Chem. 1984, 56 (11), 1898–1902. https://doi.org/10.1021/ac00275a031.

24. Gerhardt, G. A.; Oke, A. F.; Nagy, G.; Moghaddam, B.; Adams, R. N. Nafion-Coated Electrodes with High Selectivity for CNS Electrochemistry. Brain Res. 1984, 290 (2), 390– 395. https://doi.org/10.1016/0006-8993(84)90963-6.

123

25. Jang, D. P.; Kim, I.; Chang, S.-Y.; Min, H.-K.; Arora, K.; Marsh, M. P.; Hwang, S.-C.; Kimble, C. J.; Bennet, K. E.; Lee, K. H. Paired Pulse Voltammetry for Differentiating Complex Analytes. Analyst 2012, 137 (6), 1428–1435. https://doi.org/10.1039/C2AN15912K.

26. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal. Chem. 2017, 89 (11), 6166–6174. https://doi.org/10.1021/acs.analchem.7b01005.

27. Kishida, K. T.; Saez, I.; Lohrenz, T.; Witcher, M. R.; Laxton, A. W.; Tatter, S. B.; White, J. P.; Ellis, T. L.; Phillips, P. E. M.; Montague, P. R. Subsecond Dopamine Fluctuations in Human Striatum Encode Superposed Error Signals about Actual and Counterfactual Reward. Proc. Natl. Acad. Sci. U. S. A. 2016, 113 (1), 200–205. https://doi.org/10.1073/pnas.1513619112.

28. Johnson, J. A.; Gray, J. H.; Rodeberg, N. T.; Wightman, R. M. Multivariate Curve Resolution for Signal Isolation from Fast-Scan Cyclic Voltammetric Data. Anal. Chem. 2017, 89 (19), 10547–10555. https://doi.org/10.1021/acs.analchem.7b02771.

29. Keithley, R. B.; Carelli, R. M.; Wightman, R. M. Rank Estimation and the Multivariate Analysis of in Vivo Fast-Scan Cyclic Voltammetric Data. Anal. Chem. 2010, 82 (13), 5541– 5551. https://doi.org/10.1021/ac100413t.

30. Keithley, R. B.; Wightman, R. M. Assessing Principal Component Regression Prediction of Neurochemicals Detected with Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2011, 2 (9), 514–525. https://doi.org/10.1021/cn200035u.

31. Keithley, R. B.; Mark Wightman, R.; Heien, M. L. Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis. TrAC Trends Anal. Chem. 2009, 28 (9), 1127–1136. https://doi.org/10.1016/j.trac.2009.07.002.

32. Keithley, R. B.; Wightman, R. M.; Heien, M. L. Erratum to “Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis” [Trends Anal. Chem. 28 (2009) 1127–1136]. TrAC Trends Anal. Chem. 2010, 29 (1), 110. https://doi.org/10.1016/S0165-9936(09)00282-9.

33. Rodeberg, N. T.; Johnson, J. A.; Cameron, C. M.; Saddoris, M. P.; Carelli, R. M.; Wightman, R. M. Construction of Training Sets for Valid Calibration of in Vivo Cyclic Voltammetric Data by Principal Component Analysis. Anal. Chem. 2015, 87 (22), 11484–11491. https://doi.org/10.1021/acs.analchem.5b03222.

34. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2016, 7 (3), 349–359. https://doi.org/10.1021/acschemneuro.5b00302.

35. Heien, M. L. A. V.; Johnson, M. A.; Wightman, R. M. Resolving Neurotransmitters Detected by Fast-Scan Cyclic Voltammetry. Anal. Chem. 2004, 76 (19), 5697–5704. https://doi.org/10.1021/ac0491509.

124

36. Hermans, A.; Keithley, R. B.; Kita, J. M.; Sombers, L. A.; Wightman, R. M. Dopamine Detection with Fast-Scan Cyclic Voltammetry Used with Analog Background Subtraction. Anal. Chem. 2008, 80 (11), 4040–4048. https://doi.org/10.1021/ac800108j.

37. Kramer, R. Chemometric Techniques for Quantitative Analysis; CRC Press, 1998.

38. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Bucher, E. S.; McCarty, G. S.; Wightman, R. M. Characterization of Local PH Changes in Brain Using Fast-Scan Cyclic Voltammetry with Carbon Microelectrodes. Anal. Chem. 2010, 82 (23), 9892–9900. https://doi.org/10.1021/ac102399n.

39. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108 (7), 2646–2687. https://doi.org/10.1021/cr068076m.

40. Wentzell, P. D.; Vega Montoto, L. Comparison of Principal Components Regression and Partial Least Squares Regression through Generic Simulations of Complex Mixtures. Chemom. Intell. Lab. Syst. 2003, 65 (2), 257–279. https://doi.org/10.1016/S0169-7439(02)00138-7.

41. Geladi, P.; Kowalski, B. R. Partial Least-Squares Regression: A Tutorial. Anal. Chim. Acta 1986, 185, 1–17. https://doi.org/10.1016/0003-2670(86)80028-9.

42. Godoy, J. L.; Vega, J. R.; Marchetti, J. L. Relationships between PCA and PLS-Regression. Chemom. Intell. Lab. Syst. 2014, 130, 182–191. https://doi.org/10.1016/j.chemolab.2013.11.008.

43. Saurina, J.; Hernández-Cassou, S.; Saurina, J.; Fàbregas, E.; Alegret, S. Determination of in Feed Samples by Cyclic Voltammetry and Multivariate Calibration Methods. Analyst 1999, 124 (5), 733–737. https://doi.org/10.1039/A901149H.

44. Lavine, B.; Workman, J. Chemometrics. Anal. Chem. 2010, 82 (12), 4699–4711. https://doi.org/10.1021/ac101202z.

45. Sjöström, M.; Wold, S.; Lindberg, W.; Persson, J.-Å.; Martens, H. A Multivariate Calibration Problem in Analytical Chemistry Solved by Partial Least-Squares Models in Latent Variables. Anal. Chim. Acta 1983, 150, 61–70. https://doi.org/10.1016/S0003-2670(00)85460-4.

46. Roberts, J. G.; Lugo-Morales, L. Z.; Loziuk, P. L.; Sombers, L. A. Real-Time Chemical Measurements of Dopamine Release in the Brain. In Dopamine; Springer, 2013; pp 275–294.

47. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V.; et al. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat. Methods 2010, 7 (2), 126–129. https://doi.org/10.1038/nmeth.1412.

48. de Jong, S. SIMPLS: An Alternative Approach to Partial Least Squares Regression. Chemom. Intell. Lab. Syst. 1993, 18 (3), 251–263. https://doi.org/10.1016/0169-7439(93)85002-X.

125

49. Roberts, J. G.; Hamilton, K. L.; Sombers, L. A. Comparison of Electrode Materials for the Detection of Rapid Hydrogen Peroxide Fluctuations Using Background-Subtracted Fast Scan Cyclic Voltammetry. Analyst 2011, 136 (17), 3550–3556. https://doi.org/10.1039/C1AN15337D.

50. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

51. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal. Chem. 2010, 82 (5), 2020–2028. https://doi.org/10.1021/ac902753x.

52. Härdle, W.; Simar, L. Applied Multivariate Statistical Analysis; Springer, 2007; Vol. 22007.

53. Cook, R. D. Detection of Influential Observation in Linear Regression. Technometrics 1977, 19 (1), 15–18. https://doi.org/10.2307/1268249.

54. Loehlin, J. C. Latent Variable Models: An Introduction to Factor, Path, and Structural Equation Analysis; Psychology Press, 2004.

55. Maitra, S.; Yan, J. Applying Multivariate Statistical Models. Quebec City Quebec Can. Casualty Actuar. Soc. 2008.

56. Mehmood, T.; Liland, K. H.; Snipen, L.; Sæbø, S. A Review of Variable Selection Methods in Partial Least Squares Regression. Chemom. Intell. Lab. Syst. 2012, 118, 62–69. https://doi.org/10.1016/j.chemolab.2012.07.010.

57. Palermo, G.; Piraino, P.; Zucht, H.-D. Performance of PLS Regression Coefficients in Selecting Variables for Each Response of a Multivariate PLS for Omics-Type Data. Adv. Appl. Bioinforma. Chem. AABC 2009, 2, 57–70.

58. Harris, D. C. Quantitative Chemical Analysis; Macmillan, 2010.

59. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E. M.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234. https://doi.org/10.1021/acschemneuro.6b00393.

60. Venton, B. J.; Michael, D. J.; Wightman, R. M. Correlation of Local Changes in Extracellular Oxygen and PH That Accompany Dopaminergic Terminal Activity in the Rat Caudate– Putamen. J. Neurochem. 2003, 84 (2), 373–381. https://doi.org/10.1046/j.1471- 4159.2003.01527.x.

61. Deumens, R.; Blokland, A.; Prickaerts, J. Modeling Parkinson’s Disease in Rats: An Evaluation of 6-OHDA Lesions of the Nigrostriatal Pathway. Exp. Neurol. 2002, 175 (2), 303– 317. https://doi.org/10.1006/exnr.2002.7891.

126

62. Hoffman, A. F.; van Horne, C. G.; Eken, S.; Hoffer, B. J.; Gerhardt, G. A. In VivoMicrodialysis Studies of Somatodendritic Dopamine Release in the Rat Substantia Nigra: Effects of Unilateral 6-OHDA Lesions and GDNF. Exp. Neurol. 1997, 147 (1), 130–141. https://doi.org/10.1006/exnr.1997.6571.

127

CHAPTER 5

Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double Waveform Partial Least

Squares Regression

The following work was reprinted with permission from: Carl J. Meunier, Gregory S. McCarty, and Leslie A. Sombers, Analytical Chemistry, 2019, 91, 7319-7327., Copyright 2019 American

Chemical Society. Supplemental information is found in Appendix B.

5.1 Introduction

Fast-scan cyclic voltammetry (FSCV) at carbon-fiber microelectrodes (CFME) is a powerful method for assessing rapid neurochemical dynamics in biological preparations ranging from single cells to freely moving animals.1–3 When combined with pharmacological or behavioral paradigms, this approach has provided remarkable insight into the molecular mechanisms underlying goal-directed behavior.4–7 Scan rates in the hundreds of volts per second range have allowed sensitive, real-time measurements with molecular specificity for a wide variety of both electroactive and nonelectroactive (through enzyme-modified electrodes) neurochemicals.8–11

However, the rapid scan rates generate charging currents that dwarf the faradaic currents generated by target analytes. This necessitates the use of background subtraction to reveal faradaic currents generated by neurochemicals present at the electrode surface. As a result, the duration of the measurement window (from a single reference point) is limited, generally to about 90 s. After this point, background instability, or electrochemical drift, leads to subtraction artifacts. These complicate the evaluation of FSCV recordings, especially in intact animals.12,13

128

Drift is observed in practically all electrochemical systems ranging from pH probes to the electrospray emitters used in mass spectrometry.14,15 In FSCV, drift signals are diverse and stem from dynamic surface phenomena that predominately impact the double layer at the carbon electrode surface, shifting impedance. As a result, when sampling over an extended recording period (minutes to hours), a new background position must be selected for subtraction approximately every 90 s. This limits the duration of a stable measurement window and masks the detection of more gradual changes in neurotransmitter tone. It is important that the subtracted background signal is selected appropriately, because many neurochemical fluctuations are seemingly random and can be difficult to identify in the presence of drift or other interferents. As such, the use of traditional FSCV has been largely limited to studying phasic neurochemical events associated with some prominent external stimulus. This is generally accomplished over short measurement windows, precluding the study of gradual changes in analyte concentration.

Various strategies have been investigated to account for this problem. Early work by Millar and colleagues aimed to remove differential capacitance effects by characterizing nonfaradiac and faradaic responses using paired voltammetric waveforms.16 Frequency domain analyses have also proved useful, as nonfaradaic contributions can be isolated into the fundamental waveform frequency.17 More recently, an approach employing charge balanced waveforms was utilized to extend the measurement time window and evaluate changes in “tonic” extracellular dopamine; however, this required a significant reduction in temporal resolution.18 Fast-scan controlled adsorption voltammetry (FSCAV) is another approach that has been used to monitor resting levels of an analyte.19 It employs an alternating series of high and low frequency voltammetric pulses to manipulate adsorption at the electrode surface; however, this also requires a substantial reduction in temporal resolution. Wightman and colleagues developed a convolution-based approach that

129

incorporates a small amplitude potential step to sample the impedance state of the electrode before each sweep of the triangular voltammetric waveform.20 Finally, attempts have been made to extend the FSCV measurement window by using principal component regression (PCR) to quantify electrochemical drift for its removal from the overall signal, but this approach has had difficulty accounting for the wide variety of drift signals.12,21,22

Previously, our lab developed double-waveform partial-least squares regression (DW-

PLSR) to remove shifts in pH (ΔpH) and dopamine (DA) signals from in vivo recordings that

23 targeted detection of hydrogen peroxide (H2O2). In that work, voltammetric waveforms were applied in pairs. A smaller triangular sweep that excluded the redox window for the analyte of interest was scanned immediately prior to each conventional voltammetric sweep. Data collected in the smaller scan described the interferents and also served as a PLSR predictor for the interfering signals collected with the full waveform immediately thereafter. This approach enabled successful removal of interfering signals (pH and DA), improving quantification of H2O2. Herein, a DW-

PLSR strategy is used to predict and subtract electrochemical drift from extended FSCV recordings containing H2O2, DA, and adenosine fluctuations. Furthermore, we have demonstrated in vitro that drift signals from a set of electrodes can be used to create a model for drift removal that is not specific to a given electrode. The DW-PLSR model discriminates electrochemical drift from targeted analytes both in vitro and in vivo and serves as a reliable means to simultaneously evaluate multiple neurochemical fluctuations at the same sensor, from a single reference point. This permits evaluation of concentration changes occurring on both the second and minute time scales, alleviating critical experimental and analytical constraints associated with the practice of FSCV.

130

5.2 Materials and Methods

5.2.1 Data Acquisition

The double triangular waveform consisted of a small and large cyclic waveform ranging from −0.4 V to +0.3−0.8 V and −0.4 V to +1.4 V, respectively. This pair of triangular sweeps was applied using a scan rate of 400 V s −1 and separated by a 12 ms hold at −0.4 V. This waveform was applied with a frequency of 10 Hz, sampling at ∼37−50 kHz, using either a custom-built instrument for potential application and current transduction (Universal Electrochemistry

Instrument, University of North Carolina at Chapel Hill, Department of Chemistry, Electronics

Facility) or a WaveNeuro potentiostat (Pine Instruments, Durham, NC). High-Definition Cyclic

Voltammetry software (HDCV, University of North Carolina at Chapel Hill) was used for waveform output, signal processing, and data analysis, in conjunction with data acquisition cards

(National Instruments, Austin TX) used for measuring current and synchronizing the electrochemical cell with the flow-injection system. Analog filtering of the applied waveform was accomplished with a 2- pole Sallen-Key, low-pass filter at 2 kHz

5.2.2 Animal Experiments

Animal care and use was in accordance with North Carolina State University Institutional

Animal Care and Use Committee (IACUC) guidelines. Male Sprague−Dawley rats (n=2 animals;

250−300 g; Charles River Laboratories, Raleigh, NC) were individually housed on a 12:12 h light/dark cycle with free access to food and water and allowed to acclimate to the facility for several days. For acute electrode preparations, the animal was urethane anesthetized (4 g kg−1 intraperitoneal) and placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes were drilled in the skull using stereotaxic coordinates, as previously described.24,25 The working electrode was placed in the dorsal striatum (+1.2 mm AP, + 2.0 mm

131

ML, −5.0 mm DV relative to bregma) for H2O2 measurements and in the ventral striatum (+1.7 mm AP, +0.8 mm ML, −6.0 mm DV relative to bregma) for dopamine measurements; the Ag/AgCl reference electrode was placed in the contralateral forebrain. The animal’s body temperature was maintained at 37 °C by a heating pad. All working electrodes were conditioned by application of the waveform at 30 Hz for at least 30 min prior to data collection.

5.2.3 Data Processing and Analysis

All data were examined manually in HDCV before exporting to MATLAB R2018a

(Mathworks, Natick, MA), where data were digitally filtered (first-order, low pass Bessel filter with 2 kHz cutoff) prior to further analyses. PLSR was accomplished using the “plsregress” function along with the SIMPLS algorithm to center the data, to reduce data dimensionality, and to construct vector components describing the variance. Mean-squared error was determined using a 5-fold cross validation and was used to select the number of vector components retained in each

PLSR model.26 Pearson’s linear correlation coefficient was used to compare drift-subtracted CVs to those obtained using the conventional background subtraction approach. All statistical analyses and graphical depiction of data were carried out using GraphPad Prism 6 (GraphPad Software,

Inc., La Jolla, CA) or MATLAB R2018a.

5.3 Results and Discussion

5.3.1 Electrochemical Drift in FSCV Recordings

The term “electrochemical drift” is used herein to describe a variety of voltammetric signals recorded at carbon-based sensors during FSCV experiments lasting minutes or longer.

Some of these stem from physical changes. For example, etching of the carbon surface is known to occur over the lifetime of the sensor.27,28 This contributes to drift by altering the surface area

132

(affecting capacitance) and the composition of chemical functionalities on the sensor surface.27,29,30

Electrode fouling can occur via electropolymerization of redox-active species,31,32 tissue encapsulation, or adsorption of macromolecules to alter the active sensing surface (and its impedance properties).33–35 Additionally, ionic and molecular fluctuations are common in vivo and can transiently alter the double-layer structure, shifting capacitance.20,36,37 For example, the ΔpH signal contains current contributions across the entire potential window that are attributed to ion displacement by either hydroxide or hydronium ions, as well as shifts in the protoncoupled electron-transfer equilibria of surface-oxide functionalities.27,38 Furthermore, it has been shown that divalent cations (Ca2+ and Mg2+) interact with surface quinones in a potential-dependent manner.39 Due to the broad scope of sources that can potentially impact the carbon/solution interface, the drift signals observed in voltammetry data are highly diverse, can vary from one recording session to the next, and can contain multiple features across the potential-window.

5.3.2 The Double-Waveform Partial-Least-Squares Regression Model

Figure 6.1 displays an outline of the DW-PLSR model, as it applies to an investigation of electrochemical drift. To reliably identify, predict, and subtract drift from voltammetric data, a double-triangular waveform is employed (Figure 6.1A). The small waveform (sWF) scans from a

−0.4 V holding potential to +0.3/+0.8 at 400 V s−1, and the large waveform (lWF) extends to +1.4 at 400 V s−1. The switching potential of the sWF is chosen to exclude the oxidation potential of any analytes of interest and to capture information specific to interferents, in this case electrochemical drift. A sWF switching potential of +0.3 V is insufficient to substantially oxidize

DA, which prevents the model from treating it as an interferent.

PLSR is a method for multivariate statistical analysis that is similar to principal component regression (PCR), which is commonly used to isolate individual chemical signals from complex

133

FSCV data.21,40 Notably, the two approaches differ in the construction of the principal components

(PCs). PCR is an unsupervised dimensionality reduction method,41 as a set of predictors (CVs) are used to construct PCs without reference to the response (concentration). By contrast, PLSR is a supervised dimensionality reduction method that projects both predictor and response variables to a new vector space to determine the PCs that maximize the covariance of projected structures.26,41

As such, PLSR generally describes training data more efficiently with fewer PCs (than PCR), and output prediction is often more robust.42,43

Figure 5.1. Introduction to the double-waveform partial-least-squares regression (DW-PLSR) paradigm. (A) Double waveform comprised of a smaller triangular sweep (sWF) paired with a larger sweep (lWF). (B) Training set CVs describing drift (top; 8 CVs) and noise (bottom; 3 CVs) recorded on the sWF (left) and the lWF (right). (C) Cumulative percent variance (left) and 5-fold cross validation average mean-square error (right; nA2 /CV) for both the predictor (black) and response (red). (D) Color plots displaying raw data collected with the lWF (top), PLSRpredicted drift (middle; three components), and drift-subtracted (bottom) data for a 15 min in vitro recording. Time and potential are plotted as the abscissa and ordinate, respectively. Current is depicted in false color, as indicated by the color bar. Inset CVs were extracted from the data at the time point indicated by the white line. (E) Absolute current inherent to the CVs collected in the recordings displayed in (D), summed. Inset is a comparison of the total current in the drift-subtracted data to that in the training set describing the noise (bold black line is the mean; thinner lines represent the SD).

134

The DW-PLSR model is trained using data collected with the sWF and the lWF as the predictor and response, respectively. Figure 6.1B displays a representative PLSR training set comprised of sWF and lWF CVs for electrochemical drift and noise. 5-fold cross validation (Figure

6.1C) was used to determine principal component vectors that describe the training data, and cumulative percent variance (Figure 6.1C) was used to assess the noise in the data.43,44 Three principal components were retained in this model, as no improvement in the predictive power

(mean-square error) was observed by retaining more components. Figure 6.1D displays a lWF color plot for an eventless 15 min in vitro recording, composed of CVs recorded over time. The inset CV describes electrochemical drift with current contributions that span the entire potential window. The PLSR-predicted drift for this recording is also shown and is strikingly similar to the raw recording. Subtracting the PLSR-predicted drift from the raw recording results in the drift- subtracted color plot that is comprised solely of noise. To quantitatively assess the performance of

DWPLSR for drift subtraction, the absolute value of the current collected in each of these background-subtracted CVs was summed (Figure 6.1E). The CVs that comprise both the raw recording and the PLSR-predicted drift continually increase in current over the duration of the recording, indicative of electrochemical drift. Following drift subtraction, the CVs contain little current, indicating removal of drift. In fact, the total CV current following drift subtraction is comparable to the level of the noise (Figure 6.1E; inset). Thus, DW-PLSR can be used to account for electrochemical drift and should serve as a useful tool in extending the time window over which chemical fluctuations can be reliably monitored from a single reference point.

135

Figure 5.2. DW-PLSR effectively removes drift from a 15 min recording containing a rapid fluctuation of adenosine (40 μM). (A) Color plots containing raw data collected with the lWF, PLSR-predicted drift (three components), and drift-subtracted data. Insets provide an expanded view of the adenosine fluctuation (white box). (B) CVs extracted at the time point of the adenosine injection. The CV generated using DW-PLSR for drift subtraction (red) correlates well with that generated using the conventional background-subtraction approach (green, R=0.95). (C) Current versus time traces extracted at the potential corresponding to the primary (+1.35 V; black) and secondary (+1.05 V; red) oxidation of adenosine.

To investigate the ability of DW-PLSR to remove electrochemical drift that occurs concurrent with chemical fluctuations of redox-active species, a 15 min in vitro recording that contained a 2-s bolus injection of 40-μM adenosine was analyzed via DW-PLSR. Adenosine is a purine nucleoside that plays an important biological role in energy transfer, is a neuromodulator believed to play a role in regulation of blood flow, and is the subject of a number of studies incorporating FSCV.45–47 Adenosine undergoes two detectable, consecutive oxidations at a standard CFME. The first occurs at ∼+1.3 V and, following sufficient production of the first oxidation product, the second oxidation at ∼+1 V can also be monitored. Therefore, a +0.8 V upper potential limit was used for the sWF in this experiment. Figure 6.2A displays raw data collected with the lWF, PLSR-predicted drift, and drift-subtracted color plots. Insets are expanded color plots at the time point of the adenosine injection. The drift is readily identified by PLSR, and drift

136

subtraction clearly reveals the short adenosine fluctuation. A comparison of CVs from each color plot is displayed in Figure 6.2B. Importantly, the ability to use the conventional approach to background subtraction is still possible with this approach the data inherent to the initial, smaller waveform can simply be ignored. Figure 6.2B also contains a CV generated by subtracting a background CV that was collected immediately prior to the injection (conventional method). The currents reflected in the drift-subtracted CV only describe the two steps of adenosine oxidation, and the CV is consistent in shape and intensity with the CV generated by subtracting an adjacent background (R=0.95). These data demonstrate that the DW-PLSR model can effectively remove drift to reveal a rapid adenosine fluctuation occurring 12 min after the initial background position.

Furthermore, examination of the current versus time traces extracted at both oxidation potentials from the drift-subtracted data demonstrates that DW-PLSR is effective in preserving the time- delayed detection of the second oxidation step (Figure 6.2C). These data demonstrate the ability of DW-PLSR to evaluate an analyte whose signal subtly changes over time; a feature that PCR can struggle to handle.48,49

5.3.3 Subtraction of Drift from Extended Dopamine Recordings

To be broadly applicable, the DW-PLSR model must effectively remove drift from recordings that target a variety of analytes of interest. DA is one of the most widely studied neurotransmitters and is the species most commonly monitored by FSCV. H2O2 is a reactive oxygen species that is produced as a byproduct of normal mitochondrial function. It can readily

50 diffuse across the cell membrane and can be monitored via FSCV. H2O2 can also modulate striatal DA release.51 Furthermore, it is generated by oxidase enzymes permitting indirect detection of nonelectroactive molecules such as glucose, lactate, and acetyl- (choline).8–11,52 Thus, DW-

PLSR was used to remove drift from a 30 min in vitro recording containing both DA and H2O2

137

fluctuations. In this experiment, a limit of +0.3 V was used in the sWF in order to prevent the oxidation of DA in the PLSR predictor. If significant DA signal was generated with the sWF, it would be treated as an interferent and would be removed with electrochemical drift.23

Figure 6.3A displays a 30 min flow-cell recording in which calibrations of both DA and

H2O2 were performed, as well as color plots for PLSR-predicted drift, and resulting drift-subtracted data. Insets are expanded views of the final 1 μM DA (∼17 min) and 200 μM H2O2 (∼29 min) injections. The drift training set and the model validation data for this electrode are shown in

Supporting Information Figure S1. Drift subtraction clearly reveals the analyte injections that were masked by drift in the raw data. Figure 6.3B displays the sum of the absolute value of the current in each CV collected over time. The summed raw current continually increases over the duration of the recording, indicative of electrochemical drift. As a result, current fluctuations due to the redox activity of the analytes are increasingly masked. The total (summed) current for voltammograms of PLSR-predicted drift collected over time overlaps with that of the raw recording, except it does not spike during analyte injection. Following drift subtraction, the total current contained in the CVs reflects two sets of spikes with increasing intensity - these are the calibration data for each analyte, clearly revealed. Figure 6.3C displays the current versus time traces extracted from each of the color plots at the oxidation potential for DA and H2O2. In this case, the contribution of drift would lead to an underestimation of both DA and H2O2 concentrations. Following drift subtraction, the data are baseline corrected. Figure 6.3D displays

DA and H2O2 CVs extracted from the raw data, PLSR-predicted drift and drift-subtracted color plots at the time points indicated by the arrows in Figure 6.3A. CVs generated by adjacent background subtraction are overlaid with the drift-subtracted CVs, for comparison. This experiment was repeated on two additional electrodes, and corresponding CVs from those analyses

138

are displayed in Figure 6.3F. The training sets, model validation data, and color plots for the two additional electrodes are provided in Supporting Information Figure S1. It is important to note that although each drift signal exhibits unique features, the drift-subtracted CVs are consistently well correlated with those produced using adjacent background subtraction (R=0.97−0.99).

Furthermore, calibrations produced via adjacent background subtraction (black) for both DA (39.2

± 0.1 nA/ μM) and H2O2 (0.168 ± 0.002 nA/μM) were consistent with those produced when drift was subtracted using the DW-PLSR model for both DA (34.2 ± 0.1 nA/μM) and H2O2 (0.163 ±

0.003 nA/μM). Calibration data were fit using linear regression (Figure 6.3E). One-way ANCOVA revealed no significant differences in slope or intercept for DA (slope: F(1,20)=0.303, p>0.05; intercept: F(1,21)=0.901, p>0.05) and H2O2 (slope: F(1,20)=0.042, p>0.05; intercept:

F(1,21)=0.224, p>0.05) calibration lines. These data confirm that drift subtraction using the DW-

PLSR model does not skew quantification of DA and H2O2 signals, even at time-points >15 min from the initial background position. Thus, DWPLSR should provide a powerful tool for removing electrochemical drift from in vivo recordings.

139

Figure 5.3. DW-PLSR can be used to effectively remove drift from 30 min recordings containing DA (125, 250, 500, and 1000 nM) and H2O2 (25, 50, 100, and 200 μM) bolus injections in triplicate. (A) Color plots of raw data collected using the lWF, as well as PLSR-predicted drift (three components), and drift-subtracted data collected at a single electrode. An expanded view of the color plots for the last DA (1 μM, white arrow and box) and H2O2 (200 μM, red arrow and box) injections are shown as insets. (B) Sum of the currents inherent to each CV comprising the color plots in (A), over time. Tick marks on the abscissa indicate the time points of DA and H2O2 injections in the first and last half of the recording, respectively. (C) Current extracted at the oxidation potential for DA (top; + 0.5 V) and H2O2 (bottom; + 1.3 V), plotted as a function of time. (D) Calibration plots for DA (left) and H2O2 (right) created from the drift-subtracted data (initial background at 1 s), as well as by subtracting an adjacent background signal (black; conventional method). Data are displayed as the mean ± SD (n=3 electrodes). (E) CVs for 1 μM DA (left) and 200 μM H2O2 (right) extracted from the data at the time points indicated by the arrows in (A) for the raw (black), PLSR-predicted drift (blue), and drift-subtracted (red) data. A CV generated via adjacent background subtraction (green) is shown for comparison. (F) DA and H2O2 CVs extracted at the same time points as in (D) for the other two electrodes used in this study.

5.3.4 Drift Subtraction from Extended In Vivo Recordings

To evaluate effectiveness for in vivo recordings, the DW-PLSR model was used to analyze natural H2O2 fluctuations evident in a 15 min recording from the dorsal striatum of an anesthetized rat. A switching potential of +0.8 V was used in the sWF. Figure 6.4A displays the raw data,

140

PLSR-predicted drift, and drift-subtracted color plots. The data contain noticeable H2O2 dynamics, indicated by the blue/green colors at ∼+1.3 V, along with a relatively small contribution from drift.

CVs for the observed drift signals are displayed in Figure 6.4B. Training set CVs as well as cross- validation and cumulative percent variance are displayed in Supporting Information Figure S2.

Figure 6.4C displays current versus time traces taken at −0.2 V, where drift is evident, as well as at +1.3 V, where H2O2 oxidation occurs. At −0.2 V, the raw and PLSR-predicted currents are consistent, resulting in baseline correction with subtraction of drift. At +1.3 V, the raw and drift- subtracted traces correlate, indicating a minimal contribution from electrochemical drift at ∼ +1.3

V in this recording. Figure 6.4D compares CVs selected from various time points in the recording, demonstrating the effectiveness of the model. Overall, drift subtraction using the DW-PLSR model provides for easy visualization and quantification of striatal H2O2 dynamics across this 15 min window.

Next, the DW-PLSR model was employed for recordings in the ventral striatum of an anesthetized rat, a brain region containing significant dopaminergic input.53 Here, a +0.3 V switching potential was used in the sWF. The most striking data collected in this experiment are shown in Figure 6.5. These CVs were collected over 10 min at the very end of the experiment as the animal stopped breathing, due to extended anesthesia. This recording is a “worst-case scenario”, but it clearly demonstrates the utility of the model in even the most extreme case of ischemia. Ischemic events and spreading depression lead to the mass depolarization of neurons, dysregulation of oxygen and pH, and millimolar-scale changes in the concentrations of common extracellular ions (e.g., K+, Na+, Ca2+).54–57 These changes substantially affect the quality of the voltammetric data.

141

Figure 5.4. DW-PLSR model removes drift from data collected in the dorsal striatum of an anesthetized rat, improving visualization of endogenous H2O2 transients. (A) Raw data, PLSR- predicted drift (4 components), and drift-subtracted color plots collected in a 15 min recording. (B) lWF drift signals indicative of training set data (normalized to the max current). (C) Current versus time traces extracted from the color plots at −0.2 V (top) and +1.3 V (bottom). (D) CVs extracted from the color plots at the time points indicated by the white lines and lowercase letters.

The data in Figure 6.5A show a large DA signal (white line; b) collected during the ischemic event. It was recorded along with a significant amount of interfering drift. The drift was likely generated by the effects of ionic shifts on electrode impedance; variable impedance changes background size and shape.58 Training set CVs as well as cross-validation and cumulative percent variance data are displayed in Supporting Information Figure S2. Figure 6.5B plots the summed current contained in the CVs, over time. Following drift subtraction, the total current is baseline corrected and rises only at the time point of the DA fluctuation. Similarly, only features attributed to the DA redox couple are observed in the drift-subtracted color plot. The current versus times traces shown in Figure 6.5C were extracted at +0.7 V (DA oxidation), −0.2 V (dopamine-o- quinone reduction), and +1.4 V. In this case, the origin of the current observed at +1.4 V is not

142

clear. Figure 6.5D displays CVs extracted at various time points in the recording: prior to (a), during (b), and following (c) the large DA event, as well as at the end of the 10 min recording.

Electrochemical drift was a significant contributor to each of these signals, and in all cases the

DW-PLSR model was effective for drift subtraction.

Interestingly, drift subtraction revealed a gradual decrease in DA “tone” in the second half of the recording that had been masked by the much larger drift signal. This is evident as an undershoot in DA concentration in Figure 6.5C (top) and as an inverted DA CV in Figure 6.5D

(bottom). The decrease is likely due to the diffusion and breakdown of extracellular DA, concurrent with a lack of replenished vesicular stores. However, it cannot be ruled out that some portion of the DA signal was captured within the much larger drift signal used in the training set and erroneously subtracted from the data. Indeed, the maximum current of the largest drift CV in the training set is ∼700 nA (Supporting Information Figure S2), and the revealed decrease in DA is represented by only ∼−30 nA. This presents a potential complication in monitoring slower, more gradual concentration changes. In order to reliably investigate minute-to-minute chemical dynamics occurring far from the initial background position, it is necessary that the training CVs be purely representative of drift.

143

Figure 5.5. DW-PLSR model removes drift from voltammetric data collected in the ventral striatum of an anesthetized rat during a 10 min ischemic event. (A) Raw data, PLSR-predicted drift (3 components), and drift-subtracted color plots. (B) Total summed current of each CV collected over time in (A). (C) Current versus time traces extracted at +0.7 V (top), −0.2 V (middle), and +1.4 V (bottom) from each color plot. (D) CVs extracted from the color plots at the time points indicated by the white lines and lowercase letters (a−d). Inset is the rescaled drift-subtracted CV (d).

5.3.5 Integrating Drift Signals from Multiple Electrodes for Drift Subtraction

All models shown up to this point have been constructed using electrode-specific training sets. However, many studies employing FSCV incorporate permanently implanted electrodes that are used for extended recording sessions over the course of weeks to months.59,60 Electrochemical drift can vary from recording session to session, presenting a critical constraint; a drift training set may need to be made for every recording session.

144

Figure 5.6. DW-PLSR model constructed using drift signals from a collection of electrodes revealing both transient fluctuations and gradual increases in H2O2 concentration. (A) Representative raw data, PLSR-predicted drift (seven components), and drift-subtracted color plots for a 30 min recording in which both a bolus injection and a gradual increase in H2O2 occurred (same concentrations; 160 μM shown). Insets show an enhanced view of the injection. (B) The summed current of each CV as a function of time for the color plots in (A). (C) CVs extracted (white lines in A) from the injection (a; top) and near the end of the gradual increase in H2O2 concentration (b; bottom). In both plots, a CV generated by subtracting a background signal collected immediately prior to the time of the injection is included for comparison (green).

Figure 6.6A displays a color plot of raw voltammetric data collected in a 30 min flow-cell recording. A bolus injection of H2O2 was introduced at the time point indicated by the first white dashed line. In the second half of the recording, a multi-pump system was used to slowly increase

(15−30 min) the concentration of H2O2 in the flowing buffer to a final concentration equal to that of the bolus. The bolus injection of H2O2 can be deciphered at ∼1.3 V in the raw voltammetric data, even in the presence of drift. However, it is far less straightforward to distinguish H2O2 from

145

drift during the slow increase in H2O2 concentration. To evaluate if an electrode-specific training set is required for the effective removal of drift from these data, CVs describing drift were collected with multiple electrodes (n=10) and used to construct a DWPLSR model. Supporting Information

Figure S3 displays normalized CVs for drift collected with the sWF and lWF at each electrode that are representative of those used to build the PLSR training set, as well as cross-validation mean- squared error and cumulative percent variance. Seven components were retained and used to predict and subtract drift from voltammetric data. It is visually evident that the color plot describing the PLSR-predicted drift for the recording describes the observed electrochemical drift, as the overall signal across the entirety of this color plot is similar to that evident in the raw data.

Following drift subtraction, the color plot contains only two features at +1.3 V - the bolus injection and the slow increase in H2O2 concentration are clearly evident. A comparison of the total summed current of each CV over time for the raw, PLSR-predicted drift, and drift-subtracted data (Figure

6.6B) demonstrates that the model was effective for evaluating the slow increase in H2O2 concentration, as the total current gradually increases to the same magnitude as that of the bolus of H2O2 (same concentration). Figure 6.6C compares CVs extracted from the injection and near the end of the slow H2O2 increase. In both cases, the raw signal contains a significant drift contribution, which the PLSR prediction effectively captures. Following drift subtraction, the voltammetric signature for H2O2 is revealed, and CVs from both time points are of the same shape and intensity (R=0.99). Furthermore, a CV obtained by subtracting a background signal immediately prior to the bolus injection (green trace) is essentially indistinguishable from those obtained using DWPLSR for drift subtraction (R =0.99). Supporting Information Figure S3 displays CVs for bolus injections of different concentrations of H2O2 collected using three other electrodes that exhibited substantially different drift characteristics. These data reflect the ability

146

of this DW-PLSR model to predict a variety of drift signals, and to generate drift-subtracted CVs that are consistent with those from adjacent background subtraction (R2=0.95−0.99). Thus, it is clearly possible to develop effective models for subtraction of drift from fast voltammetric data that are not specific to any given electrode, to reveal both rapid and gradual changes in analyte concentration over time.

5.4 Conclusions

Electrochemical drift can be observed in any application of electrochemistry. In FSCV, drift occurs when changes at the carbon/solution interface impact the background signal. Drift not only limits the duration over which FSCV can accurately monitor rapid chemical fluctuations

(typically the background is stable for 30−90 s), but it also negatively impacts selectivity and complicates interpretation of the data. Herein, we demonstrate that PLSR can be combined with a double voltammetric waveform to provide an effective means of determining and subtracting drift contributions from in vivo recordings. By using data collected with a sWF that does not encompass the redox activity of the analyte(s) as a predictor variable, the drift inherent to data collected with the subsequent lWF can be predicted and nonsubjectively subtracted. A major advantage of this technique is that it simplifies training set construction for in vivo analyses, as researchers are no longer required to search for analyte-pure CVs for all components of the chemical signal; rather, only interferent CVs are necessary. The DW-PLSR model can be used to subtract drift signals and reveal chemical fluctuations reflected by currents that span multiple orders of magnitude. In fact, it was demonstrated in vitro that drift information can be collected at a set of many electrodes, pooled, and used for the effective subtraction of drift as long as sufficient variety is included in the training set. However, it is important to note that potential sources of electrochemical drift are

147

greater in vivo, and effective implementation of a training set that is not electrode-specific to the analysis of in vivo data will require more investigation. Importantly, drift subtraction did not alter the signal of any of the target analytes investigated. The DWPLSR paradigm for drift removal promises to simplify data analysis and reduce user bias, as there is no longer a need to subtract backgrounds at multiple points over the course of a long (minutes) recording. As a result, it is expected that this approach will broaden the applicability of FSCV for monitoring chemical fluctuations occurring on multiple time scales.

148

5.5 References

1. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, 2006.

2. Phillips, P. E.; Robinson, D. L.; Stuber, G. D.; Carelli, R. M.; Wightman, R. M. Real-Time Measurements of Phasic Changes in Extracellular Dopamine Concentration in Freely Moving Rats by Fast-Scan Cyclic Voltammetry. In Drugs of Abuse; Springer, 2003; pp 443–464.

3. Pihel, Karin.; Schroeder, T. J.; Wightman, R. Mark. Rapid and Selective Cyclic Voltammetric Measurements of Epinephrine and Norepinephrine as a Method To Measure Secretion from Single Bovine Adrenal Medullary Cells. Anal. Chem. 1994, 66 (24), 4532–4537. https://doi.org/10.1021/ac00096a021.

4. Cheer, J. F.; Aragona, B. J.; Heien, M. L. A. V.; Seipel, A. T.; Carelli, R. M.; Wightman, R. M. Coordinated Accumbal Dopamine Release and Neural Activity Drive Goal-Directed Behavior. Neuron 2007, 54 (2), 237–244. https://doi.org/10.1016/j.neuron.2007.03.021.

5. Cheer, J. F.; Wassum, K. M.; Heien, M. L. A. V.; Phillips, P. E. M.; Wightman, R. M. Cannabinoids Enhance Subsecond Dopamine Release in the Nucleus Accumbens of Awake Rats. J. Neurosci. Off. J. Soc. Neurosci. 2004, 24 (18), 4393–4400. https://doi.org/10.1523/JNEUROSCI.0529-04.2004.

6. Phillips, P. E. M.; Stuber, G. D.; Heien, M. L. A. V.; Wightman, R. M.; Carelli, R. M. Subsecond Dopamine Release Promotes Cocaine Seeking. Nature 2003, 422 (6932), 614–618. https://doi.org/10.1038/nature01476.

7. Roitman, M. F.; Stuber, G. D.; Phillips, P. E. M.; Wightman, R. M.; Carelli, R. M. Dopamine Operates as a Subsecond Modulator of Food Seeking. J. Neurosci. Off. J. Soc. Neurosci. 2004, 24 (6), 1265–1271. https://doi.org/10.1523/JNEUROSCI.3823-03.2004.

8. Wassum, M. K.; Tolosa, M. V.; Wang, J.; Walker, E.; Monbouquette, G. H.; Maidment, T. N. Silicon Wafer-Based Platinum Microelectrode Array Biosensor for Near Real-Time Measurement of Glutamate in Vivo. Sensors 2008, 8 (8). https://doi.org/10.3390/s8085023.

9. Smith, S. K.; Lee, C. A.; Dausch, M. E.; Horman, B. M.; Patisaul, H. B.; McCarty, G. S.; Sombers, L. A. Simultaneous Voltammetric Measurements of Glucose and Dopamine Demonstrate the Coupling of Glucose Availability with Increased Metabolic Demand in the Rat Striatum. ACS Chem. Neurosci. 2017, 8 (2), 272–280. https://doi.org/10.1021/acschemneuro.6b00363.

10. Smith, S. K.; Gosrani, S. P.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Carbon-Fiber Microbiosensor for Monitoring Rapid Lactate Fluctuations in Brain Tissue Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2018, 90 (21), 12994–12999. https://doi.org/10.1021/acs.analchem.8b03694.

11. Parikh, V.; Kozak, R.; Martinez, V.; Sarter, M. Prefrontal Acetylcholine Release Controls Cue Detection on Multiple Timescales. Neuron 2007, 56 (1), 141–154. https://doi.org/10.1016/j.neuron.2007.08.025.

149

12. Hermans, A.; Keithley, R. B.; Kita, J. M.; Sombers, L. A.; Wightman, R. M. Dopamine Detection with Fast-Scan Cyclic Voltammetry Used with Analog Background Subtraction. Anal. Chem. 2008, 80 (11), 4040–4048. https://doi.org/10.1021/ac800108j.

13. Howell, J. O.; Kuhr, W. G.; Ensman, R. E.; Mark Wightman, R. Background Subtraction for Rapid Scan Voltammetry. J. Electroanal. Chem. Interfacial Electrochem. 1986, 209 (1), 77– 90. https://doi.org/10.1016/0022-0728(86)80187-5.

14. Hunt, S. M.; Sheil, M. M.; Belov, M.; Derrick, P. J. Probing the Effects of Cone Potential in the Electrospray Ion Source: Consequences for the Determination of Molecular Weight Distributions of Synthetic Polymers. Anal. Chem. 1998, 70 (9), 1812–1822. https://doi.org/10.1021/ac971288m.

15. McLauchlan, G.; Rawlings, J. M.; Lucas, M. L.; McCloy, R. F.; Crean, G. P.; McColl, K. E. Electrodes for 24 Hours PH Monitoring--a Comparative Study. Gut 1987, 28 (8), 935. https://doi.org/10.1136/gut.28.8.935.

16. Millar, J.; Williams, G. V. Fast Differential Ramp Voltammetry: A New Voltammetric Technique Designed Specifically for Use in Neuronal Tissue. J. Electroanal. Chem. Interfacial Electrochem. 1990, 282 (1), 33–49. https://doi.org/10.1016/0022-0728(91)85088-7.

17. Brazill, S. A.; Bender, S. E.; Hebert, N. E.; Cullison, J. K.; Kristensen, E. W.; Kuhr, W. G. Sinusoidal Voltammetry:: A Frequency Based Electrochemical Detection Technique. J. Electroanal. Chem. 2002, 531 (2), 119–132. https://doi.org/10.1016/S0022-0728(02)01067-7.

18. Oh, Y.; Park, C.; Kim, D. H.; Shin, H.; Kang, Y. M.; DeWaele, M.; Lee, J.; Min, H.-K.; Blaha, C. D.; Bennet, K. E.; et al. Monitoring In Vivo Changes in Tonic Extracellular Dopamine Level by Charge-Balancing Multiple Waveform Fast-Scan Cyclic Voltammetry. Anal. Chem. 2016, 88 (22), 10962–10970. https://doi.org/10.1021/acs.analchem.6b02605.

19. Atcherley, C. W.; Laude, N. D.; Parent, K. L.; Heien, M. L. Fast-Scan Controlled-Adsorption Voltammetry for the Quantification of Absolute Concentrations and Adsorption Dynamics. Langmuir 2013, 29 (48), 14885–14892. https://doi.org/10.1021/la402686s.

20. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal. Chem. 2017, 89 (11), 6166–6174. https://doi.org/10.1021/acs.analchem.7b01005.

21. Keithley, R. B.; Wightman, R. M. Assessing Principal Component Regression Prediction of Neurochemicals Detected with Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2011, 2 (9), 514–525. https://doi.org/10.1021/cn200035u.

22. Keithley, R. B.; Mark Wightman, R.; Heien, M. L. Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis. TrAC Trends Anal. Chem. 2009, 28 (9), 1127–1136. https://doi.org/10.1016/j.trac.2009.07.002.

23. Meunier, C. J.; Mitchell, E. C.; Roberts, J. G.; Toups, J. V.; McCarty, G. S.; Sombers, L. A. Electrochemical Selectivity Achieved Using a Double Voltammetric Waveform and Partial

150

Least Squares Regression: Differentiating Endogenous Hydrogen Peroxide Fluctuations from Shifts in PH. Anal. Chem. 2018, 90 (3), 1767–1776. https://doi.org/10.1021/acs.analchem.7b03717.

24. Roberts, J.; Lugo-Morales, L.; Loziuk, P.; Sombers, L. Leading Methods in Dopamine Research. 2013.

25. Paxinos, G.; Watson, C. The Rat Brain in Stereotaxic Coordinates: Hard Cover Edition; Elsevier, 2006.

26. Mehmood, T.; Liland, K. H.; Snipen, L.; Sæbø, S. A Review of Variable Selection Methods in Partial Least Squares Regression. Chemom. Intell. Lab. Syst. 2012, 118, 62–69. https://doi.org/10.1016/j.chemolab.2012.07.010.

27. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal. Chem. 2010, 82 (5), 2020–2028. https://doi.org/10.1021/ac902753x.

28. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108 (7), 2646–2687. https://doi.org/10.1021/cr068076m.

29. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir 2010, 26 (11), 9116– 9122. https://doi.org/10.1021/la9048924.

30. Mitchell, E. C.; Dunaway, L. E.; McCarty, G. S.; Sombers, L. A. Spectroelectrochemical Characterization of the Dynamic Carbon-Fiber Surface in Response to Electrochemical Conditioning. Langmuir 2017, 33 (32), 7838–7846. https://doi.org/10.1021/acs.langmuir.7b01443.

31. Harreither, W.; Trouillon, R.; Poulin, P.; Neri, W.; Ewing, A. G.; Safina, G. Carbon Nanotube Fiber Microelectrodes Show a Higher Resistance to Dopamine Fouling. Anal. Chem. 2013, 85 (15), 7447–7453. https://doi.org/10.1021/ac401399s.

32. Hashemi, P.; Dankoski, E. C.; Petrovic, J.; Keithley, R. B.; Wightman, R. M. Voltammetric Detection of 5-Hydroxytryptamine Release in the Rat Brain. Anal. Chem. 2009, 81 (22), 9462– 9471. https://doi.org/10.1021/ac9018846.

33. Lempka, S. F.; Miocinovic, S.; Johnson, M. D.; Vitek, J. L.; McIntyre, C. C. In Vivoimpedance Spectroscopy of Deep Brain Stimulation Electrodes. J. Neural Eng. 2009, 6 (4), 046001. https://doi.org/10.1088/1741-2560/6/4/046001.

34. A. Mercanzini*; P. Colin; J. Bensadoun; A. Bertsch; P. Renaud. In Vivo Electrical Impedance Spectroscopy of Tissue Reaction to Microelectrode Arrays. IEEE Trans. Biomed. Eng. 2009, 56 (7), 1909–1918. https://doi.org/10.1109/TBME.2009.2018457.

151

35. Williams, J. C.; Hippensteel, J. A.; Dilgen, J.; Shain, W.; Kipke, D. R. Complex Impedance Spectroscopy for Monitoring Tissue Responses to Inserted Neural Implants. J. Neural Eng. 2007, 4 (4), 410–423. https://doi.org/10.1088/1741-2560/4/4/007.

36. Nicholson, C. Measurement of Extracellular Ions in the Brain. Trends Neurosci. 1980, 3 (9), 216–218. https://doi.org/10.1016/0166-2236(80)90081-8.

37. Rogers, M. L.; Feuerstein, D.; Leong, C. L.; Takagaki, M.; Niu, X.; Graf, R.; Boutelle, M. G. Continuous Online Microdialysis Using Microfluidic Sensors: Dynamic Neurometabolic Changes during Spreading Depolarization. ACS Chem. Neurosci. 2013, 4 (5), 799–807. https://doi.org/10.1021/cn400047x.

38. Deakin, M. R.; Kovach, P. M.; Stutts, K. J.; Wightman, R. Mark. Heterogeneous Mechanisms of the Oxidation of Catechols and Ascorbic Acid at Carbon Electrodes. Anal. Chem. 1986, 58 (7), 1474–1480. https://doi.org/10.1021/ac00298a046.

39. Kim, Y. J.; Wu, W.; Chun, S.-E.; Whitacre, J. F.; Bettinger, C. J. Catechol-Mediated Reversible Binding of Multivalent Cations in Eumelanin Half-Cells. Adv. Mater. 2014, 26 (38), 6572–6579. https://doi.org/10.1002/adma.201402295.

40. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E. M.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234. https://doi.org/10.1021/acschemneuro.6b00393.

41. Härdle, W.; Simar, L. Applied Multivariate Statistical Analysis; Springer, 2007; Vol. 22007.

42. Wentzell, P. D.; Vega Montoto, L. Comparison of Principal Components Regression and Partial Least Squares Regression through Generic Simulations of Complex Mixtures. Chemom. Intell. Lab. Syst. 2003, 65 (2), 257–279. https://doi.org/10.1016/S0169-7439(02)00138-7.

43. Geladi, P.; Kowalski, B. R. Partial Least-Squares Regression: A Tutorial. Anal. Chim. Acta 1986, 185, 1–17. https://doi.org/10.1016/0003-2670(86)80028-9.

44. Lavine, B.; Workman, J. Chemometrics. Anal. Chem. 2008, 80 (12), 4519–4531. https://doi.org/10.1021/ac800728t.

45. Dunwiddie, T. V.; Masino, S. A. The Role and Regulation of Adenosine in the Central Nervous System. Annu. Rev. Neurosci. 2001, 24 (1), 31–55. https://doi.org/10.1146/annurev.neuro.24.1.31.

46. Lee, S. T.; Venton, B. J. Regional Variations of Spontaneous, Transient Adenosine Release in Brain Slices. ACS Chem. Neurosci. 2018, 9 (3), 505–513. https://doi.org/10.1021/acschemneuro.7b00280.

47. Ross, A. E.; Venton, B. J. Adenosine Transiently Modulates Stimulated Dopamine Release in the Caudate–Putamen via A1 Receptors. J. Neurochem. 2015, 132 (1), 51–60. https://doi.org/10.1111/jnc.12946.

152

48. Borman, R. P.; Wang, Y.; Nguyen, M. D.; Ganesana, M.; Lee, S. T.; Venton, B. J. Automated Algorithm for Detection of Transient Adenosine Release. ACS Chem. Neurosci. 2017, 8 (2), 386–393. https://doi.org/10.1021/acschemneuro.6b00262.

49. Ross, A. E.; Venton, B. J. Sawhorse Waveform Voltammetry for Selective Detection of Adenosine, ATP, and Hydrogen Peroxide. Anal. Chem. 2014, 86 (15), 7486–7493. https://doi.org/10.1021/ac501229c.

50. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2013, 4 (5), 782–789. https://doi.org/10.1021/cn4000499.

51. Bao, L.; Avshalumov, M. V.; Patel, J. C.; Lee, C. R.; Miller, E. W.; Chang, C. J.; Rice, M. E. Mitochondria Are the Source of Hydrogen Peroxide for Dynamic Brain-Cell Signaling. J. Neurosci. 2009, 29 (28), 9002. https://doi.org/10.1523/JNEUROSCI.1706-09.2009.

52. Burmeister, J. J.; Gerhardt, G. A. Self-Referencing Ceramic-Based Multisite Microelectrodes for the Detection and Elimination of Interferences from the Measurement of l-Glutamate and Other Analytes. Anal. Chem. 2001, 73 (5), 1037–1042. https://doi.org/10.1021/ac0010429.

53. Robbins, T. W.; Everitt, B. J. Functions of Dopamine in the Dorsal and Ventral Striatum. Milest. Dopamine Res. 1992, 4 (2), 119–127. https://doi.org/10.1016/1044-5765(92)90010-Y.

54. Borjigin, J.; Leem, U.; Liu, T.; Pal, D.; Huff, S.; Klarr, D.; Sloboda, J.; Hernandez, J.; Wang, M. M.; Mashour, G. A. Surge of Neurophysiological Coherence and Connectivity in the Dying Brain. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (35), 14432–14437.

55. Hansen, A. J. Effect of Anoxia on Ion Distribution in the Brain. Physiol. Rev. 1985, 65 (1), 101–148. https://doi.org/10.1152/physrev.1985.65.1.101.

56. Siemkowicz, E.; Hansen, A. Brain Extracellular Ion Composition and EEG Activity Following 10 Minutes Ischemia in Normo-and Hyperglycemic Rats. Stroke 1981, 12 (2), 236–240.

57. Hobbs, C. N.; Holzberg, G.; Min, A. S.; Wightman, R. M. Comparison of Spreading Depolarizations in the Motor Cortex and Nucleus Accumbens: Similar Patterns of Oxygen Responses and the Role of Dopamine. ACS Chem. Neurosci. 2017, 8 (11), 2512–2521. https://doi.org/10.1021/acschemneuro.7b00266.

58. Meunier, C. J.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Background Signal as an in Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2017, 8 (2), 411–419. https://doi.org/10.1021/acschemneuro.6b00325.

59. Kishida, K. T.; Saez, I.; Lohrenz, T.; Witcher, M. R.; Laxton, A. W.; Tatter, S. B.; White, J. P.; Ellis, T. L.; Phillips, P. E. M.; Montague, P. R. Subsecond Dopamine Fluctuations in Human Striatum Encode Superposed Error Signals about Actual and Counterfactual Reward.

153

Proc. Natl. Acad. Sci. U. S. A. 2016, 113 (1), 200–205. https://doi.org/10.1073/pnas.1513619112.

60. Schwerdt, H. N.; Shimazu, H.; Amemori, K.-I.; Amemori, S.; Tierney, P. L.; Gibson, D. J.; Hong, S.; Yoshida, T.; Langer, R.; Cima, M. J.; et al. Long-Term Dopamine Neurochemical Monitoring in Primates. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (50), 13260–13265. https://doi.org/10.1073/pnas.1713756114.

154

CHAPTER 6

The Background Signal as an In Situ Predictor of Dopamine Oxidation Potential:

Improving Interpretation of Fast-Scan Cyclic Voltammetry Data

The following work was reprinted with permission from: Carl J. Meunier, James G. Roberts,

Gregory S. McCarty, and Leslie A. Sombers, ACS Chemical Neuroscience, 2017, 8, 411-419.,

Copyright 2017 American Chemical Society.

6.1 Introduction

Background-subtracted fast-scan cyclic voltammetry employing carbon-fiber microelectrodes (CFME) is a powerful tool for making real-time, subsecond measurements of electroactive molecules in discrete brain regions.1,2 This approach originated with the work of

Millar and colleagues,3,4 and has seen increasing use in the past decade (Figure 3.1) due to thorough characterization and technological improvements made largely by Wightman and colleagues.5–10

When combined with behavioral and pharmacological paradigms, it has provided remarkable information about the molecular mechanisms underlying specific aspects of goal-directed behavior and associative learning, significantly advancing this field.11–19 Despite this, broad acceptance by the neuroscience community has been hindered by remaining difficulties associated with its practice, such as reliable electrode calibration.

Individual CFME performance is variable and dependent on physical characteristics inherent to the handmade electrodes and the recording microenvironment. As a result, electrodespecific calibration is required to accurately relate the observed current to analyte concentration. Traditionally, calibration is done in a flow-injection system following an

155

experiment, as the electrode surface changes in response to tissue exposure and this alters electrode sensitivity.20,21 However, postcalibration assumes that the homogeneous buffered electrolyte solution used in vitro replicates the unknown, complex, and chemically dynamic environment of brain tissue. Traditional calibration schemes do not account for the involvement of spectator species, such as proteins, which can affect both working and reference electrodes in vivo. Also, traditional calibration assumes that electrode sensitivity is constant over the course of the experiment. These assumptions contradict fundamentals of analytical chemistry, which state that the calibration medium should resemble (as closely as possible) the environment in which the measurement was made,22 or calibration may result in significant misinterpretation of data.

Additionally, univariate calibration fails when the signals from multiple analytes overlap, which is often the case for in vivo electrochemical data. Principal component analysis combined with inverse least squares regression, also called principal component regression (PCR),23 is a multivariate calibration technique. When used with fast-scan cyclic voltammetry (FSCV), training sets are composed of voltammograms for individual electroactive analytes to resolve chemical contributions to the signal, as well as calibration factors to estimate sensitivity and enable quantification.24–26 A more detailed description of the use of PCR with FSCV can be found elsewhere.27

156

Figure 6.1. Number of references in PubMed using “fast-scan cyclic voltammetry” per year, since 1984.

PCR has been effectively employed in analysis of in vivo data for acutely implanted

CFMEs.24,28 The relatively brief exposure to tissue minimizes the neuroimmunological response,29 and acutely implanted electrodes can be easily removed for postcalibration to obtain in vitro calibration factors as estimates of sensitivity in vivo. However, chronically implanted electrodes introduce a significant procedural constraint with respect to calibration, due to the duration of the implant (weeks to months),30 and difficulties in safely recovering the electrode for postcalibration.

These constraints have led to use of “standard” or “universal” training sets, obtained from other electrodes, recording sessions, or subjects, to analyze in vivo data.31–33 Even when training sets are constructed from data collected in vitro using the electrode that was implanted in vivo, so-called

“matrix effects” can lead to inconsistencies between the in vivo data and the data that comprise the training set (i.e., changes in the shape and position of peaks in the voltammogram).34 For these reasons, analytical chemistry textbooks conclude that standards used for calibration should be run in the same matrix as the unknown.22,35 Thus, when monitoring molecules in brain tissue a reliable and straightforward method for in situ electrode calibration is essential.

157

Recent work in our group has moved toward this goal by demonstrating that the total background current can be used as a predictor of electrode sensitivity to multiple analytes, in situ.21 Herein, we expand upon this work by exploring changes in electrochemical performance resulting from variable impedance. Electrode fouling is mimicked by the addition of impedance to the electrochemical system. The resulting shifts in sensitivity, and in the oxidation potentials of both quinonelike moieties on the carbon surface (Ep,QH) and of dopamine (Ep,DA), are characterized.

Finally, an in vitro model that predicts Ep,DA based on Ep,QH is described and validated. This powerful tool will facilitate in situ determination of appropriate electrochemical parameters for accurate quantification. Additionally, the information within will be fundamental in the development of more automated data analysis methods.

6.2 Methods

6.2.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. In vitro electrochemical experiments were performed in either a phosphate-buffered saline (10 mM Na2HPO4, 138 mM

NaCl, and 2.7 mM KCl) or TRIS buffered saline (15 mM 2-amino-2-hydroxymethylpropane-1,3- diol, 3.25 mM KCl, 1.20 mM CaCl2, 1.2 mM MgCl2, 2 mM Na2SO4, 1.25 mM NaH2PO4, and 145 mM NaCl). Adrenal slice experiments used bicarbonate buffered saline (125 mM NaCl, 26 mM

NaHCO3, 2.5 mM KCl, 2.0 mM CaCl2, 1.0 mM MgCl2, 1.3 mM NaH2PO4, 10 mM HEPES, and

10 mM glucose) saturated with 95% O2 and 5% CO2. All buffers were adjusted to pH 7.4 with 1

M NaOH and 1 M HCl. Reserpine (10 mg mL−1) and α-methyl-DL-tyrosine methyl ester

158

hydrochloride (α-MPT, 250 mg mL−1) were prepared in 50% saline/50% dimethyl sulfoxide and in saline, respectively. All aqueous solutions were made using ultrapure 18.2 MΩ water.

6.2.2 Microelectrode Fabrication

Cylindrical carbon-fiber microelectrodes were fabricated using T-650 carbon fibers (Cytec

Industries, Inc., Woodland Park, NJ) as previously described.1 In short, a single carbon fiber was aspirated into a glass capillary, a glass seal was created using a micropipette puller (Narishige,

Tokyo, Japan), and the fiber extending past the seal was cut to 100 μm in length. Electrical connection to the carbon-fiber was established via a conductive silver epoxy. A Ag/AgCl pellet reference electrode (World Precision Instruments, Inc., Sarasota, FL) was used to complete the two-electrode electrochemical cell.

6.2.3 Model Circuit Fabrication and Characterization

Resistance and capacitance values for circuit components were confirmed with a voltmeter.

Five circuits with constant resistance (60, 110, 160, 210, and 310 kΩ) were produced in a manner that allowed capacitors (0, 3.93, 7.43, 10.19, and 13.82 nF) to be interchanged. Additionally, a circuit containing four 10 kΩ resistors in series was fabricated to add data points between 0 and

40 kΩ. This resulted in 29 different circuits. EIS Spectrum Analyzer software (available online at http://www. abc.chemistry.bsu.by/vi/analyser/) was used to simulate electrochemical impedance spectroscopic values for each circuit.36

6.2.4 Flow-Injection System

Working electrodes were positioned in a custom electrochemical flow-cell using a micromanipulator (World Precision Instruments, Sarasota, FL). A syringe pump (New Era Pump

Systems, Wantagh, NY) supplied a continuous buffer stream of 0.5 mL min−1 (1.0 mL min−1 for data in Figure 3.3) across the working and reference electrodes. Two-second bolus injections of

159

analyte to the microelectrode surface were accomplished with a six-port HPLC valve mounted on a two-position actuator controlled by a digital pneumatic solenoid valve (Valco Instruments,

Houston, TX). The entire apparatus was enclosed in a custom-built grounded Faraday cage.

6.2.5 Data Acquisition

A triangular waveform from −0.4 to +1.3 V with a −0.4 V holding potential was applied at a scan rate of 400 V s−1 , an application frequency of 10 Hz, and a sampling frequency of 200 kHz

(100 kHz for data in Figure 3.3) using a custom built instrument for potential application and current transduction (University of North Carolina at Chapel Hill, Department of Chemistry,

Electronics Facility). High Definition Cyclic Voltammetry software (University of North Carolina at Chapel Hill) was used for waveform output, data analysis and signal processing in conjunction with data acquisition cards (National Instruments, Austin TX) used for measuring current and synchronizing the electrochemical cell with the flow-injection system. Analog filtering was accomplished with a 2-pole Sallen-Key, low-pass filter at 2 kHz. There was no additional analog or digital filtering of the data.

6.2.6 Animal Experiments

Male Sprague−Dawley rats (250−300 g, Charles River Laboratories, Raleigh, NC) were used in all animal experiments. Animal care and use was in accordance with North Carolina State

University Institutional Animal Care and Use Committee (IACUC) guidelines. One animal was used as an acute electrode preparation. The animal was urethane anesthetized (4 g kg−1 intraperitoneally), placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes were drilled in the skull using stereotaxic coordinates, as described previously.1 The working electrode was placed in the striatum (1.2 mm AP, 3.0 mm ML, −5.0 mm DV relative to

160

bregma), and the Ag/AgCl reference electrode was placed in the contralateral forebrain. The animal’s body temperature was maintained at 37 °C by a heating pad.

One animal was treated with reserpine (5 mg/kg) and α-MPT (250 mg/kg) daily for 3 days.

The animal was urethane anesthetized (4 g kg−1 intraperitoneally), decapitated, and adrenal glands were quickly removed, trimmed of fat tissue, and embedded in 3% agarose gel in bicarbonate- buffered saline (BBS). The gel blocks were placed in cold BBS and 400 μm thick slices were cut with a vibratome (World Precision Instruments, Sarasota, FL). The slices were allowed to rest in

BBS for 1 h, then transferred to a recording chamber (Warner instruments, Hamden, CT) and superfused with BBS maintained at 34 °C for at least 30 min prior to electrical stimulation. Carbon- fiber microelectrodes were positioned in tissue near twin tungsten stimulating electrodes.

Electrode placements were made with the aid of a microscope (Nikon Instruments, Inc., Melville,

NY) and micromanipulators (Scientifica, ltd., United Kingdom). Electrical stimulations consisted of trains of 16 biphasic, 500 μA pulses at a frequency of 60 Hz with a pulse width of 2 ms.

A chronic electrode preparation was investigated in one animal. Surgery was performed as previously described.30 Briefly, the animal was anesthetized with 2.5%−4% isoflurane, placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes and screws were drilled in the skull using stereotaxic coordinates. The working electrode was placed in the

NAc (+1.2 mm AP, +1.4 mm ML, −7.0 mm DV relative to bregma), the stimulation electrode in the ventral tegmental area (−5.2 mm AP, + 1.0 mm ML, −8.2 mm DV relative to bregma), and a

Ag/AgCl reference electrode was placed in the contralateral forebrain. The components were permanently affixed with dental cement. The animal’s body temperature was maintained at 37 °C using a heating pad. The animal was allowed to recover for 2 weeks prior to administering

161

electrical stimulation. Electrical stimulations consisted of trains of 12 biphasic, 175 μA pulses at a frequency of 60 Hz with a pulse width of 2 ms, performed every other day.

6.2.7 Statistics

Data are presented as the mean ± standard deviation, unless otherwise noted. Paired t tests were employed to compare two groups of data. One-way ANOVA with a Tukey’s post hoc test and two-way ANOVA were used to compare significance between multiple groups. ANCOVA

(with Tukey’s post hoc test) was used to compare linear regression models. Significance was determined at p < 0.05. Statistical analyses and graphical depictions of the data were carried out using GraphPad Prism 6 (GraphPad Software, Inc., La Jolla, CA) and MATLAB R2012B

(Mathworks, Natick, MA).

6.3 Results and Discussion

6.3.1 Combining Fast-Scan Cyclic Voltammetry and Principal Component Regression

The detection of catecholamines using FSCV is commonly performed with a triangular waveform ramping from a −0.4 V holding potential to +1.3 V and back.1,2 Herein, this waveform

(Figure 3.2A) is applied to a CFME with a scan rate of 400 V s−1, an application frequency of 10

Hz, and a sampling frequency of 200 kHz (except for data present in Figure 3.3 which were sampled at 100 kHz). When applied to a CFME, it generates a large but stable background signal composed predominately of nonfaradaic capacitive (charging) current,37 with smaller faradaic contributions from electroactive species that are statically present in the extracellular space1 and redox groups inherent to the electrode surface.38 When the concentration of an electroactive analyte fluctuates in the immediate vicinity of the electrode, the background current (Figure 3.2B, black dashed line) can be subtracted to reveal the faradaic current generated in response to the

162

concentration change. The resulting voltammogram serves to identify the analyte and can also be used for quantification (Figure 3.2C). This effectively removes contributions to the current that result from processes inherent to the electrode, and from species that exist at relatively constant concentrations. The vertical dashed lines delineate the oxidation potentials of quinone-like surface

38–40 functionalities and surface confined quinone species, (Ep,QH, Figure 3.2B), and the oxidation potential of dopamine (Ep,DA, Figure 3.2C). The current at Ep,DA (ordinate) can be plotted versus analyte concentration (abscissa) and the slope of a fitted regression line determines electrode sensitivity (Figure 3.2D).1,2

Figure 6.2. Overview of fast-scan cyclic voltammetry. (A) A triangular waveform from −0.4 to +1.3 V is applied at 400 V/s and repeated at 10 Hz. (B) Representative background (black) and dopamine (red) CVs. (C) Representative background-subtracted CV for 1 μM dopamine. In both (B) and (C), potential and current are plotted as the abscissa and ordinate, with dashed lines indicating positions of Ep,QH and Ep,DA, respectively. (D) Calibration curve for dopamine using a single CFME. Dopamine concentration and current at Ep,DA are plotted on the abscissa and ordinate, respectively. Data are presented as mean ± standard deviation (n=3 replicate injections).

A comprehensive training set consists of voltammograms for all analytes expected to be present in the experimental data, with concentrations spanning the physiological range.27,34

Voltammetric signals collected in vivo are highly variable in terms of peak width, position, and

163

relative intensity. Additional variation results from instrumentation, inherent differences in carbon-fiber preparations and surface structure,39,41 reference electrode performance,42 and differences in recording microenvironments. In fact, oxidation potentials can dynamically shift during an in vivo electrochemical experiment. When oxidation potentials are intentionally offset from those used in the training set, systematic misestimation of analyte concentrations and/or failures in residual analysis for model validation arise.27,34 Thus, training a model for PCR is not straightforward, and the selection of an inappropriate training set could result in unreliable assignment of analyte concentrations.27,34 Realization of this unresolved issue initiated investigation into how electrode surface states and recording microenvironments influence oxidation potentials.

6.3.2 The Carbon-Fiber Microelectrode Surface Influences Oxidation Potential

Conditioning of the CFME changes the chemical nature of the carbon surface,43 and as such is expected to influence analyte oxidation potentials. To verify this, three surface conditions were investigated using a flow-injection system. Dopamine was detected in a TRIS buffered saline

(1) before implantation in the striatum of an anesthetized rat for 30 min, (2) after removal from the striatum, and (3) following postimplantation cleaning in isopropyl alcohol (IPA) overnight.

Representative CVs are displayed in Figure 3.3A,B. Potential is plotted on the abscissa and normalized current (I/Ip) on the ordinate. The dashed lines indicate nominal values for the oxidation potentials, Ep,QH and Ep,DA. The difference between the two potentials (ΔEp,DA-QH) is displayed in Figure 3.3C. Data are presented as the mean ± standard deviation (n=12 electrodes).

One-way ANOVA with Tukey’s post hoc test was employed for multiple comparisons of Ep,QH,

Ep,DA and ΔEp,DA-QH resulting from different electrode states. There was a significant effect of the surface state on Ep,QH and Ep,DA; F(2,33) = 6.895, **p = 0.0031; and F(2,33) = 16.83, ****p <

164

0.0001, respectively. A Tukey’s post hoc test revealed that Ep,QH and Ep,DA are shifted to more positive potentials following implantation in striatal tissue. This is attributed to fouling of the electrode surface by adsorbed biological material. Cleaning the electrode in IPA reverses the effects of fouling, returning the oxidation potentials to approximately their original values, suggesting removal of adsorbed material. Perhaps the most interesting result is that ΔEp,DA-QH is conserved for each electrode condition; F(2,33) = 0.8282, p > 0.05. Thus, the oxidation potentials are similarly affected by changes to the electrode surface, suggesting that the position of Ep,QH could serve as a predictor for the position of Ep,DA. To investigate this, a series of circuits of varying impedance was used to controllably shift oxidation potential positions in vitro.

Figure 6.3. Electrode surface influences oxidation potential. Normalized (A) background CVs and (B) background-subtracted dopamine CVs collected on an electrode: (1) before tissue implantation (black), (2) postimplantation (red), and (3) postimplantation and cleaning in IPA (blue). Potential and normalized current are plotted as the abscissa and ordinate, respectively. Dashed lines indicate nominal values of the oxidation potentials. (C) Summary of Ep,QH (top), Ep,DA (middle), and ΔEp,DA-QH (bottom) as a function of the electrode state. Data represent mean ± standard deviation (n=12 electrodes). Significance was determined via a one-way ANOVA with Tukey’s post-test.

165

6.3.3 Model Circuits to Simulate Impedance Contributions Due to Electrode Fouling

The nonspecific adsorption of material to the electrode surface perturbs the electrochemical system. This can be treated as a change in impedance and modeled using equivalent circuits.37

Generally, an electrode−electrolyte interface can be modeled with a Randles equivalent circuit consisting of solution resistance (Rs) in series with both double-layer capacitance (CDL) and

37,44–46 charge-transfer resistance (RCT); such that CDL is in parallel with RCT (Figure 3.4). Warburg impedance is a constant phase element that describes impedance due to diffusing redox species. It is observed at low frequencies (<100 Hz) not relevant to this work, so it is not included.47 The simplified circuit model represents a two electrode electrochemical system in tissue. It consists of a potentiostat connected to two Randles equivalent circuits representing the working and reference electrodes (RE is the resistance inherent to the electrode). A variable impedance element is included to represent impedance contributions due to fouling (of both working and reference electrodes), and the system is completed by including solution resistance (Rs). Work characterizing the impedance properties of single carbon fiber microelectrodes in vivo is minimal. However, fouling contributions to and silicon microelectrode impedance have been studied extensively using electrochemical impedance spectroscopy (EIS).48–52 These studies rely on the assumption that the electrode state does not change significantly over the duration of the experiment. However, impedance is expected to vary due to the dynamic nature of the tissue and slower changes in the microenvironment (inflammation, gliosis, etc.) that occur over the course of an extended electrode implant. The data presented in Figure 3.3C define the approximate range of Ep,QH and Ep,DA that must be achievable by controlling impedance.

166

Figure 6.4. Equivalent circuit diagram depicting an in vivo electrochemical system. Both the working and reference electrodes are represented by Randles equivalent circuits with an impedance contribution due to adsorbed material (fouling) that is simulated by adding impedance (Z) between the potentiostat and working electrode.

Figure 3.5A (inset) displays a schematic of the Randles circuit used to mimic shifts in impedance due to biological fouling, and a table of the values for resistance (R) and capacitance

(C) used. These values represent a myriad of possible contributing factors such as bulk tissue resistance and tissue response to the implanted electrodes by way of matrix proteins, microglia, and astrocytes (the response is dynamic across the duration of the implant).48,50,53 Changes in capacitance can be attributed to microglia/astrocyte cell membrane capacitances as well as the adsorption of proteins (which affects surface area), and changes in extracellular ionic composition.48,50,51 A representative Nyquist plot resulting from an EIS simulation of a model

Randles circuit is also shown (Figure 3.5A), in which the real and imaginary components of impedance are plotted on the abscissa and ordinate, respectively. Each point making up the Nyquist plot represents impedance due to different alternating current frequencies, decreasing from left to right. A standard voltammetric sweep is completed in ∼8.5 ms when using common waveform parameters in FSCV (−0.4 to 1.3 V applied at 400 V/s). If this triangle were constantly repeated, the time of the applied triangular waveform (8.5 ms) could be thought of as the “period” of the

167

waveform, approximated by that of a sine wave of ∼118 Hz (dotted line, Figure 3.5A inset). Thus, this frequency was sampled for a series of 27 circuits, including one with no added impedance, to study how impedance modulates the position of Ep,QH and Ep,DA. Figure 3.5B displays a contour plot describing the relationship between capacitance (x-axis), resistance (y-axis), and total impedance (z-axis) of the circuits at 118 Hz. Total impedance is also depicted in color. It is evident that impedance increases as resistance increases, and it decreases as capacitance increases. The effects of capacitance are much more evident at high versus low resistance. However, the plot clearly indicates that these relationships are not linear. Rather, there is a complex relationship that spans a range of impedance values.

Figure 6.5. Impedance characterization of model Randles circuits. (A) Representative Nyquist plot for a Randles circuit (top right), as well as a table of the resistance (R) and capacitance (C) values employed. Triangular and sine waveforms are overlaid (bottom right) to depict the period used in determining an appropriate impedance frequency (118 Hz). (B) Total impedance at 118 Hz as a function of resistance and capacitance. Capacitance, resistance, and impedance are plotted on the x, y, and z axes, respectively; color also denotes impedance. Simulated data points are indicated by the black dots.

168

6.3.4 Employing Model Circuits to Investigate How Impedance Shifts Redox Potential

Following characterization, the circuits were coupled with CFMEs in either phosphate- buffered saline (PBS) or TRIS-buffered saline (TRIS). Both buffers are commonly used in electrochemical experiments at physiological pH; however, TRIS contains amines (present in many biomolecules) that competitively bind to the electrode,54 creating a situation that is thought to more closely resemble conditions in the brain. The model circuits were implemented between the potentiostat and the working electrode; the reference electrode was unaltered. Impedance is additive across the electrochemical cell. As such, altering the impedance of one electrode should serve to shift impedance of the entire circuit. Background CVs and background-subtracted dopamine CVs were used to determine Ep,QH and Ep,DA values. Triplicate standards were averaged across five electrodes for each impedance preparation in PBS and TRIS.

Figure 3.6A displays representative background CVs (left) and background-subtracted dopamine CVs (1 μM, right) for impedance preparations in which the total resistance was systematically increased (60, 110, 160, 210, 310 kΩ) while the capacitance was held constant (3.19 nF) in PBS (top) or TRIS (bottom) buffer. The black traces were collected with no added impedance (control), and simulated impedance values are described in the legend. The dashed lines indicate the mean potentials (n = 5 electrodes) corresponding to Ep,QH and Ep,DA in the control circuit, with no impedance added. The data clearly demonstrate that adding impedance impacts the nature of the electrochemical system, as both Ep,QH and Ep,DA shift to more positive potentials as impedance is added. A positive shift in these oxidation potentials is also observed in the presence of TRIS, an amine that competitively binds to the electrode surface, as compared to PBS solution

(with no impedance added: Ep,QH, t(4)=5.74, **p=0.0023; Ep,DA, t(4)=3.675, *p=0.0107, one-tailed t test, data not shown). Increased impedance also results in decreased current amplitude (Figure

169

3.6B; PBS, F(5,24)=15.05, ****p<0.0001; TRIS, F(5,24)=21.07, ****p<0.0001), as well as decreased electrode sensitivity (Figure 3.6C, F(2,30)=56.52, ****p<0.0001). Working in TRIS buffer also reduces the amplitude of the signal, but this is largely due to the presence of the amines that occupy dopamine adsorption sites on the sensor surface.54 Taken together, these data underscore the importance of improving calibration strategies, and the need for an accurate in situ approach.21

Figure 3.7A plots the position of Ep,QH (blue) and Ep,DA (red) as a function of added impedance in both PBS (left) and TRIS (right) buffers. At 118 Hz, it is apparent that this relationship is linear (R2≥0.87, n=5). Figure 3.7B compares the slopes of the regression lines fit to these data. Slope (mV/kΩ) is plotted on the ordinate, with the various peak potentials identified on the abscissa; columns are color coordinated according to buffer. A two-way ANOVA determined that the source of variation in the slopes is due to the oxidation potential monitored,

F(1,100)=9.034, **p=0.0033, and not due to the buffer, F(1,100)=0.2363, p>0.05. Thus, Ep,QH is modulated differently (per unit impedance) than Ep,DA. These results also demonstrate that the effects of impedance are consistent across these buffers. Figure 3.7C displays the relationship between Ep,QH (abscissa) and Ep,DA (ordinate) in both PBS (black) and TRIS (gray) buffers. Each point represents the mean (n=5) value for a given impedance preparation, with the lowest values

(bottom left) collected with no added impedance and the highest values (top right) with 310 kΩ of added impedance. Ep,DA,PBS values are plotted on the left ordinate and Ep,DA,TRIS values on the right ordinate. The linear relationship demonstrates that Ep,QH can be used as a reliable predictor of Ep,DA.

This is important, because it shows that this model can be used to identify training set parameters that are most likely to be appropriate for PCR and data analysis, and to thus improve interpretation of in vivo data. ANCOVA was used to compare regression models in both PBS and TRIS, and no

170

differences were found (F(1,50)=0.154, p>0.05). The more physiologically relevant TRIS model was employed for further ex vivo validation.

Figure 6.6. Impedance impacts the nature of the electrochemical system. (A) Representative background CVs (left) and background-subtracted dopamine CVs (1 μM, right) resulting from systematic increases in impedance using the model circuits. Values in the legend are simulated impedance at 118 Hz for each circuit. In each case the capacitance was held constant (3.93 nF) and the resistance varied (60, 110, 160, 210, 310 kΩ). Data was collected in PBS (top) and TRIS (bottom). Mean values of Ep,QH and Ep,DA (n=5) for the preparation with no added impedance are indicated with dashed lines. (B) Added impedance significantly decreases peak current (at Ep,DA) for dopamine oxidation when working in either PBS (left) or TRIS buffer (right). One-way ANOVA with Tukey’s post hoc test was used to compare each preparation with added impedance to that with no added impedance, for both PBS and TRIS (n=5). (C) Sensitivity to dopamine is significantly decreased with additional impedance. Data was collected in PBS with no added impedance (black), 60 kΩ (red), and 310 kΩ (blue) of added impedance (n=3). A summary of the sensitivities is provided in the table. Slopes were compared via ANCOVA with Tukey’s post hoc test. All data are presented as mean ± standard deviation.

171

Figure 6.7. Establishing a relationship between Ep.QH and Ep,DA. (A) Ep.QH (blue) and Ep,DA (red) as a function of added impedance (abscissa) in both PBS (left) and TRIS (right). (B) Comparison of the slopes (mV/kΩ) of the regression model for both Ep,QH and Ep,DA as a function of added impedance in PBS (black) and TRIS (gray). Significance was determined via a two-way ANOVA. (C) Plot of Ep,QH (abscissa) versus Ep,DA (ordinate) resulting from all impedance preparations in PBS (black, left ordinate) and TRIS (gray, right ordinate). ANCOVA was used to compare the two regression models. In all cases error bars represent the mean ± standard deviation (n=5).

6.3.5 Validation of the Impedance Model

Electrically evoked catecholamine (CA) release was detected in CA-depleted adrenal tissue slices as well as in the nucleus accumbens (NAc) of a freely moving animal to validate a predictive model for Ep,DA using Ep,QH. The adrenal medulla contains significant quantities of epinephrine and norepinephrine, and lower levels of dopamine, even when the animal has been treated with reserpine and α-MPT to decrease CA levels.55,56 The oxidation potentials of these CAs are indistinguishable from one another using the waveform employed and the resulting CVs are quite similar.41,57 Adrenal tissue provides an effective model for validation because it is significantly impeding, as evident by the fact that catecholamine CVs have Ep,QH and Ep,DA values higher than in the NAc (Figure 3.8A). This may be due to high levels of CA fouling the electrode via electopolymerization.58 Electrodes (n=5) were positioned in the medulla and allowed to cycle for

172

30−60 min prior to electrical stimulation every 5 min for 30 min. To further evaluate the model, electrically evoked catecholamine release was also detected at a CFME chronically implanted in the NAc over the course of 3 weeks. Background CVs and background-subtracted catecholamine

CVs were used to determine Ep,QH and Ep,DA, respectively.

Figure 3.8A displays representative color plots and CVs from the two tissue preparations.

Figure 3.8B displays the relationship between the critical oxidation potentials for individual electrical stimulation events (black symbols) as they relate to the TRIS regression model (gray).

Measured values for Ep,QH and Ep,DA are plotted on the abscissa and ordinate, respectively (black).

It is apparent that the relationship between Ep,QH and Ep,DA approximates the relationship established in vitro, confirming that the range of impedances used to create the model was appropriate. The individual data points should be normally distributed about the regression model; however, the tissue data is offset from the regression line. In tissue, ongoing background redox processes are present that are not accounted for in the flow-cell. To investigate the role of background redox processes on the Ep,QH/Ep,DA relationship, physiological concentrations of electroactive species present in the NAc were added to the buffer. Ascorbic acid (AA, 50 μM) and

3,4-dihydroxyphenylacetic acid (DOPAC, 5 μM) were used to create a TRIS-AA-DOPAC buffer

(Figure 3.8B, blue),59,60 and a TRIS-DA buffer (Figure 3.8B, red) contained 5 μM dopamine.

ANCOVA determined no significant difference in the slopes of TRIS, TRIS-AA-DOPAC, and

TRIS-DA and of a regression line fit to the live tissue data (F(3,119)=0.84, p=0.4766). ANCOVA did reveal that the y-intercepts were significantly different (F(3,122)=13.2, ****p<0.0001).

Tukey’s post hoc test determined that TRIS-AA-DOPAC, TRIS-DA, and the live tissue regression lines had intercepts that were not different (p≥0.37), but that were all different from the TRIS model (****p<0.0001). Therefore, the presence of background redox processes is critical when

173

predicting Ep,DA in tissue. Figure 3.8C plots the relationship between the measured Ep,DA (abscissa) and Ep,DA predicted (ordinate) using the linear model for the TRIS (gray), TRIS-AA-DOPAC

(blue), and TRIS-DA (red) buffers. If the models reliably predict Ep,DA, then all the resulting points should fall on, or be normally distributed about, the dashed unity line, as in the models that incorporate background redox processes.

Figure 6.8 - Model validation in live tissue. (A) Representative color plots (top), background CVs (middle), and background-subtracted catecholamine CVs (bottom) collected in adrenal and brain (NAc) tissue. In the color plots, time is the ordinate, potential is the abscissa, and current is plotted in false color with the scale bar below the color plot. Time of electrical stimulation is denoted by red arrows. (B) Measured values for Ep,QH and Ep,DA from individual electrical stimulations are plotted on the abscissa and ordinate, respectively. One group of data points was obtained from five CFMEs acutely implanted in adrenal tissue. The other group of data points was obtained over 3 weeks using a chronically implanted CFME. The four lines display the linear regressions for the TRIS (gray), TRIS-DA (red, n=3 electrodes), and TRIS-AA-DOPAC (blue, n=3 electrodes) models, and for the live tissue data (black). ANCOVA with Tukey’s post hoc test was used to compare the slopes and intercepts of the regression models. A table summarizing the regression statistics is provided below. (C) Measured Ep,DA (abscissa) and predicted Ep,DA (ordinate) obtained using Ep,QH in the TRIS (gray), TRIS-DA (red), and TRIS-AA-DOPAC (blue) models. The dashed line has a slope equal to unity.

174

6.4 Conclusion

Precise and reliable quantification of neurochemical concentration is critical in order to accurately interpret the molecular mechanisms underlying any aspect of brain function. Electrode calibration is necessary for accurate quantification; however, the nature of in vivo measurements can make conventional postcalibration difficult, or even impossible. Even when calibration is feasible, the analyte-specific voltammograms and scaling factors that lie at the core of PCR can shift or fluctuate in vivo, largely due to changes in impedance brought about by long-term biological processes. Thus, an in situ method for accurate electrode calibration is imperative. We provide a step toward this goal by describing a model that uses the oxidation potential of quinone- like moieties in the background signal as an accurate predictor for the oxidation potential of dopamine at any time point in an electrochemical experiment. Importantly, this model does not assume that current−potential relationships are fixed across the entirety of an experiment. One example of its utility would be when the experimental signal-tonoise ratio is insufficient to resolve peak oxidation potential. The nature of the background signal could bring attention to inconsistencies between the training set and the experimental data, so that the training set could be adjusted as required. Overall, this work characterizes the effects of impedance on FSCV measurements, and provides a tool for researchers to make informed decisions when constructing appropriate PCR models. It promises to simplify data collection protocols for in vivo applications, and to facilitate widespread use of FSCV by allowing more freedom in experimental design, particularly when combined with a previously described approach to predict electrode sensitivity in situ.21 Thus, it represents a significant step toward more automated analysis of in vivo data.

175

6.5 References

1. Roberts, J. G.; Lugo-Morales, L. Z.; Loziuk, P. L.; Sombers, L. A. Real-Time Chemical Measurements of Dopamine Release in the Brain. In Dopamine; Springer, 2013; pp 275–294.

2. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, 2006.

3. Stamford, J. A.; Kruk, Z. L.; Millar, J. Regional Differences in Extracellular Ascorbic Acid Levels in the Rat Brain Determined by High Speed Cyclic Voltammetry. Brain Res. 1984, 299 (2), 289–295. https://doi.org/10.1016/0006-8993(84)90710-8.

4. Stamford, J. A.; Kruk, Z. L.; Millar, J.; Wightman, R. M. Striatal Dopamine Uptake in the Rat: In Vivo Analysis by Fast Cyclic Voltammetry. Neurosci. Lett. 1984, 51 (1), 133–138. https://doi.org/10.1016/0304-3940(84)90274-X.

5. Kile, B. M.; Walsh, P. L.; McElligott, Z. A.; Bucher, E. S.; Guillot, T. S.; Salahpour, A.; Caron, M. G.; Wightman, R. M. Optimizing the Temporal Resolution of Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2012, 3 (4), 285–292. https://doi.org/10.1021/cn200119u.

6. Takmakov, P.; McKinney, C. J.; Carelli, R. M.; Wightman, R. M. Instrumentation for Fast- Scan Cyclic Voltammetry Combined with Electrophysiology for Behavioral Experiments in Freely Moving Animals. Rev. Sci. Instrum. 2011, 82 (7), 074302. https://doi.org/10.1063/1.3610651.

7. Wightman, R. M.; Heien, M. L. A. V.; Wassum, K. M.; Sombers, L. A.; Aragona, B. J.; Khan, A. S.; Ariansen, J. L.; Cheer, J. F.; Phillips, P. E. M.; Carelli, R. M. Dopamine Release Is Heterogeneous within Microenvironments of the Rat Nucleus Accumbens. Eur. J. Neurosci. 2007, 26 (7), 2046–2054. https://doi.org/10.1111/j.1460-9568.2007.05772.x.

8. Keithley, R. B.; Takmakov, P.; Bucher, E. S.; Belle, A. M.; Owesson-White, C. A.; Park, J.; Wightman, R. M. Higher Sensitivity Dopamine Measurements with Faster-Scan Cyclic Voltammetry. Anal. Chem. 2011, 83 (9), 3563–3571. https://doi.org/10.1021/ac200143v.

9. Hermans, A.; Keithley, R. B.; Kita, J. M.; Sombers, L. A.; Wightman, R. M. Dopamine Detection with Fast-Scan Cyclic Voltammetry Used with Analog Background Subtraction. Anal. Chem. 2008, 80 (11), 4040–4048. https://doi.org/10.1021/ac800108j.

10. Heien, M. L. A. V.; Johnson, M. A.; Wightman, R. M. Resolving Neurotransmitters Detected by Fast-Scan Cyclic Voltammetry. Anal. Chem. 2004, 76 (19), 5697–5704. https://doi.org/10.1021/ac0491509.

11. Roitman, M. F.; Stuber, G. D.; Phillips, P. E. M.; Wightman, R. M.; Carelli, R. M. Dopamine Operates as a Subsecond Modulator of Food Seeking. J. Neurosci. 2004, 24 (6), 1265. https://doi.org/10.1523/JNEUROSCI.3823-03.2004.

12. Gan, J. O.; Walton, M. E.; Phillips, P. E. M. Dissociable Cost and Benefit Encoding of Future Rewards by Mesolimbic Dopamine. Nat. Neurosci. 2010, 13 (1), 25–27. https://doi.org/10.1038/nn.2460.

176

13. Day, J. J.; Roitman, M. F.; Wightman, R. M.; Carelli, R. M. Associative Learning Mediates Dynamic Shifts in Dopamine Signaling in the Nucleus Accumbens. Nat. Neurosci. 2007, 10 (8), 1020–1028. https://doi.org/10.1038/nn1923.

14. Phillips, P. E. M.; Stuber, G. D.; Heien, M. L. A. V.; Wightman, R. M.; Carelli, R. M. Subsecond Dopamine Release Promotes Cocaine Seeking. Nature 2003, 422 (6932), 614–618. https://doi.org/10.1038/nature01476.

15. Howe, M. W.; Tierney, P. L.; Sandberg, S. G.; Phillips, P. E. M.; Graybiel, A. M. Prolonged Dopamine Signalling in Striatum Signals Proximity and Value of Distant Rewards. Nature 2013, 500 (7464), 575–579. https://doi.org/10.1038/nature12475.

16. Stouffer, M. A.; Woods, C. A.; Patel, J. C.; Lee, C. R.; Witkovsky, P.; Bao, L.; Machold, R. P.; Jones, K. T.; de Vaca, S. C.; Reith, M. E. A.; et al. Insulin Enhances Striatal Dopamine Release by Activating Cholinergic Interneurons and Thereby Signals Reward. Nat. Commun. 2015, 6 (1), 8543. https://doi.org/10.1038/ncomms9543.

17. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2013, 4 (5), 782–789. https://doi.org/10.1021/cn4000499.

18. Hashemi, P.; Dankoski, E. C.; Wood, K. M.; Ambrose, R. E.; Wightman, R. M. In Vivo Electrochemical Evidence for Simultaneous 5-HT and Histamine Release in the Rat Substantia Nigra Pars Reticulata Following Medial Forebrain Bundle Stimulation. J. Neurochem. 2011, 118 (5), 749–759. https://doi.org/10.1111/j.1471-4159.2011.07352.x.

19. Wagner, A. K.; Drewencki, L. L.; Chen, X.; Santos, F. R.; Khan, A. S.; Harun, R.; Torres, G. E.; Michael, A. C.; Dixon, C. E. Chronic Methylphenidate Treatment Enhances Striatal Dopamine Neurotransmission after Experimental Traumatic Brain Injury. J. Neurochem. 2009, 108 (4), 986–997. https://doi.org/10.1111/j.1471-4159.2008.05840.x.

20. Logman, M. J.; Budygin, E. A.; Gainetdinov, R. R.; Wightman, R. M. Quantitation of in Vivo Measurements with Carbon Fiber Microelectrodes. J. Neurosci. Methods 2000, 95 (2), 95–102. https://doi.org/10.1016/S0165-0270(99)00155-7.

21. Roberts, J. G.; Toups, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode Calibration Strategy for Voltammetric Measurements In Vivo. Anal. Chem. 2013, 85 (23), 11568–11575. https://doi.org/10.1021/ac402884n.

22. Harris, D. C. Quantitative Chemical Analysis; Macmillan, 2010.

23. Kramer, R. Chemometric Techniques for Quantitative Analysis; CRC Press, 1998.

24. Michael L. A. V. Heien; Khan, A. S.; Ariansen, J. L.; Cheer, J. F.; Paul E. M. Phillips; Wassum, K. M.; Wightman, R. M.; Murray, R. W. Real-Time Measurement of Dopamine Fluctuations after Cocaine in the Brain of Behaving Rats. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (29), 10023–10028.

177

25. Keithley, R. B.; Mark Wightman, R.; Heien, M. L. Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis. TrAC Trends Anal. Chem. 2009, 28 (9), 1127–1136. https://doi.org/10.1016/j.trac.2009.07.002.

26. Keithley, R. B.; Wightman, R. M.; Heien, M. L. Erratum to “Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis” [Trends Anal. Chem. 28 (2009) 1127–1136]. TrAC Trends Anal. Chem. 2010, 29 (1), 110. https://doi.org/10.1016/S0165-9936(09)00282-9.

27. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2016, 7 (3), 349–359. https://doi.org/10.1021/acschemneuro.5b00302.

28. Lugo-Morales, L. Z.; Loziuk, P. L.; Corder, A. K.; Toups, J. V.; Roberts, J. G.; McCaffrey, K. A.; Sombers, L. A. Enzyme-Modified Carbon-Fiber Microelectrode for the Quantification of Dynamic Fluctuations of Nonelectroactive Analytes Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2013, 85 (18), 8780–8786. https://doi.org/10.1021/ac4017852.

29. Peters, J. L.; Miner, L. H.; Michael, A. C.; Sesack, S. R. Ultrastructure at Carbon Fiber Microelectrode Implantation Sites after Acute Voltammetric Measurements in the Striatum of Anesthetized Rats. J. Neurosci. Methods 2004, 137 (1), 9–23. https://doi.org/10.1016/j.jneumeth.2004.02.006.

30. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V.; et al. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat. Methods 2010, 7 (2), 126–129. https://doi.org/10.1038/nmeth.1412.

31. Flagel, S. B.; Clark, J. J.; Robinson, T. E.; Mayo, L.; Czuj, A.; Willuhn, I.; Akers, C. A.; Clinton, S. M.; Phillips, P. E. M.; Akil, H. A Selective Role for Dopamine in Stimulus–Reward Learning. Nature 2011, 469 (7328), 53–57. https://doi.org/10.1038/nature09588.

32. Clark, J. J.; Collins, A. L.; Sanford, C. A.; Phillips, P. E. M. Dopamine Encoding of Pavlovian Incentive Stimuli Diminishes with Extended Training. J. Neurosci. 2013, 33 (8), 3526. https://doi.org/10.1523/JNEUROSCI.5119-12.2013.

33. Hart, A. S.; Rutledge, R. B.; Glimcher, P. W.; Phillips, P. E. M. Phasic Dopamine Release in the Rat Nucleus Accumbens Symmetrically Encodes a Reward Prediction Error Term. J. Neurosci. 2014, 34 (3), 698. https://doi.org/10.1523/JNEUROSCI.2489-13.2014.

34. Rodeberg, N. T.; Johnson, J. A.; Cameron, C. M.; Saddoris, M. P.; Carelli, R. M.; Wightman, R. M. Construction of Training Sets for Valid Calibration of in Vivo Cyclic Voltammetric Data by Principal Component Analysis. Anal. Chem. 2015, 87 (22), 11484–11491. https://doi.org/10.1021/acs.analchem.5b03222.

35. Skoog, D. A.; West, D. M.; Holler, F. J.; Crouch, S. R. Fundamentals of Analytical Chemistry; Nelson Education, 2013.

178

36. Bondarenko, A. S.; Ragoisha, G. A. Progress in Chemometrics Research. Nova Sci. Publ. N. Y. 2005, 89–102.

37. Bard, A. J.; Faulkner, L. R. Fundamentals and Applications. Electrochem. Methods 2001, 2, 482.

38. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108 (7), 2646–2687. https://doi.org/10.1021/cr068076m.

39. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Bucher, E. S.; McCarty, G. S.; Wightman, R. M. Characterization of Local PH Changes in Brain Using Fast-Scan Cyclic Voltammetry with Carbon Microelectrodes. Anal. Chem. 2010, 82 (23), 9892–9900. https://doi.org/10.1021/ac102399n.

40. DuVall, S. H.; McCreery, R. L. Self-Catalysis by Catechols and Quinones during Heterogeneous Electron Transfer at Carbon Electrodes. J. Am. Chem. Soc. 2000, 122 (28), 6759–6764. https://doi.org/10.1021/ja000227u.

41. Heien, M. L. A. V.; Phillips, P. E. M.; Stuber, G. D.; Seipel, A. T.; Wightman, R. M. Overoxidation of Carbon-Fiber Microelectrodes Enhances Dopamine Adsorption and Increases Sensitivity. Analyst 2003, 128 (12), 1413–1419. https://doi.org/10.1039/B307024G.

42. Moussy, Francis.; Harrison, D. Jed. Prevention of the Rapid Degradation of Subcutaneously Implanted Ag/AgCl Reference Electrodes Using Polymer Coatings. Anal. Chem. 1994, 66 (5), 674–679. https://doi.org/10.1021/ac00077a015.

43. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir 2010, 26 (11), 9116– 9122. https://doi.org/10.1021/la9048924.

44. Chang, B.-Y.; Park, S.-M. Electrochemical Impedance Spectroscopy. Annu. Rev. Anal. Chem. 2010, 3 (1), 207–229. https://doi.org/10.1146/annurev.anchem.012809.102211.

45. Lisdat, F.; Schäfer, D. The Use of Electrochemical Impedance Spectroscopy for Biosensing. Anal. Bioanal. Chem. 2008, 391 (5), 1555.

46. W. Franks; I. Schenker; P. Schmutz; A. Hierlemann. Impedance Characterization and Modeling of Electrodes for Biomedical Applications. IEEE Trans. Biomed. Eng. 2005, 52 (7), 1295–1302. https://doi.org/10.1109/TBME.2005.847523.

47. Pollak, V. An Equivalent Diagram for the Interface Impedance of Metal Needle Electrodes. Med. Biol. Eng. 1974, 12 (4), 454–459.

48. Williams, J. C.; Hippensteel, J. A.; Dilgen, J.; Shain, W.; Kipke, D. R. Complex Impedance Spectroscopy for Monitoring Tissue Responses to Inserted Neural Implants. J. Neural Eng. 2007, 4 (4), 410–423. https://doi.org/10.1088/1741-2560/4/4/007.

179

49. Xiaoqiu Huang; Duc Nguyen; D. W. Greve; M. M. Domach. Simulation of Microelectrode Impedance Changes Due to Cell Growth. IEEE Sens. J. 2004, 4 (5), 576–583. https://doi.org/10.1109/JSEN.2004.831302.

50. A. Mercanzini*; P. Colin; J. Bensadoun; A. Bertsch; P. Renaud. In Vivo Electrical Impedance Spectroscopy of Tissue Reaction to Microelectrode Arrays. IEEE Trans. Biomed. Eng. 2009, 56 (7), 1909–1918. https://doi.org/10.1109/TBME.2009.2018457.

51. Lempka, S. F.; Miocinovic, S.; Johnson, M. D.; Vitek, J. L.; McIntyre, C. C. In Vivoimpedance Spectroscopy of Deep Brain Stimulation Electrodes. J. Neural Eng. 2009, 6 (4), 046001. https://doi.org/10.1088/1741-2560/6/4/046001.

52. McAdams, E. T.; Jossinet, J. Tissue Impedance: A Historical Overview. Physiol. Meas. 1995, 16 (3A), A1–A13. https://doi.org/10.1088/0967-3334/16/3a/001.

53. Szarowski, D. H.; Andersen, M. D.; Retterer, S.; Spence, A. J.; Isaacson, M.; Craighead, H. G.; Turner, J. N.; Shain, W. Brain Responses to Micro-Machined Silicon Devices. Brain Res. 2003, 983 (1), 23–35. https://doi.org/10.1016/S0006-8993(03)03023-3.

54. Bath, B. D.; Michael, D. J.; Trafton, B. J.; Joseph, J. D.; Runnels, P. L.; Wightman, R. M. Subsecond Adsorption and Desorption of Dopamine at Carbon-Fiber Microelectrodes. Anal. Chem. 2000, 72 (24), 5994–6002. https://doi.org/10.1021/ac000849y.

55. Martínez-Olivares, R.; Villanueva, I.; Racotta, R.; Piñón, M. Depletion and Recovery of Catecholamines in Several Organs of Rats Treated with Reserpine. Auton. Neurosci. 2006, 128 (1), 64–69. https://doi.org/10.1016/j.autneu.2006.04.004.

56. Toneff, T.; Funkelstein, L.; Mosier, C.; Abagyan, A.; Ziegler, M.; Hook, V. Beta-Amyloid Peptides Undergo Regulated Co-Secretion with Neuropeptide and Catecholamine Neurotransmitters. Peptides 2013, 46, 126–135. https://doi.org/10.1016/j.peptides.2013.04.020.

57. Mundorf, M. L.; Joseph, J. D.; Austin, C. M.; Caron, M. G.; Wightman, R. M. Catecholamine Release and Uptake in the Mouse Prefrontal Cortex. J. Neurochem. 2001, 79 (1), 130–142. https://doi.org/10.1046/j.1471-4159.2001.00554.x.

58. Harreither, W.; Trouillon, R.; Poulin, P.; Neri, W.; Ewing, A. G.; Safina, G. Carbon Nanotube Fiber Microelectrodes Show a Higher Resistance to Dopamine Fouling. Anal. Chem. 2013, 85 (15), 7447–7453. https://doi.org/10.1021/ac401399s.

59. Zetterström, T.; Sharp, T.; Marsden, C. A.; Ungerstedt, U. In Vivo Measurement of Dopamine and Its Metabolites by Intracerebral Dialysis: Changes After d-Amphetamine. J. Neurochem. 1983, 41 (6), 1769–1773. https://doi.org/10.1111/j.1471-4159.1983.tb00893.x.

60. Parsons, L. H.; Justice Jr., J. B. Extracellular Concentration and In Vivo Recovery of Dopamine in the Nucleus Accumbens Using Microdialysis. J. Neurochem. 1992, 58 (1), 212– 218. https://doi.org/10.1111/j.1471-4159.1992.tb09298.x.

180

CHAPTER 7

Interpreting Dynamic Interfacial Changes at Carbon Fiber Microelectrodes Using

Electrochemical Impedance Spectroscopy

The work was completed in collaboration with J. Dylan Denison, Gregory S. McCarty, and Leslie

A. Sombers, and has been submitted to the journal Langmuir.

7.1 Introduction

Carbon-fiber ultramicroelectrodes are an invaluable tool for monitoring rapid molecular fluctuations with high spatiotemporal resolution due to their low cost, ease of fabrication, and resistance to biofouling.1,2 These electrodes are commonly paired with fast-scan cyclic voltammetry (FSCV) to monitor real-time neurochemical fluctuations in freely moving animals.

This approach has allowed careful investigation of dopaminergic signaling mechanisms, especially in the rat striatum, and has provided remarkable insight into how this signal modulates specific basal ganglia circuits that encode goal-directed behavior and associative learning.3–6 Additionally,

CFMEs have been used for catecholamine measurements at single cells, and in live tissue slices, to investigate a wide range of biophysical and biochemical aspects of neurotransmitter release.7–10

At its most fundamental level, electrochemical detection involves the heterogeneous transfer of electrons across the electrode/solution interface. As such, electrochemical performance is dictated by the physical properties of the system. The chemical nature of a given type of carbon fiber affects electrochemical performance,11–13 and oxidative pretreatments (i.e. flame-etching, light activation, or application of oxidative potentials) are routinely incorporated into experimental design to improve performance by altering the chemical nature of the electrode surface.13–15 Even

181

in a simple benchtop system, handmade electrodes are known to exhibit variable performance.

Yet, the effect of the recording environment on the electrochemical signal is less well characterized. Several reports have demonstrated that tissue exposure and the chemical and physical composition of the recording environment affect electrochemical performance.16–19 It has been shown on the bench top that the composition of the buffer employed for in vitro investigations can directly impact the observed response,19 and substantial chemical diversity is observed in tissue, in terms of both the spatial and the temporal regime. However, in vivo calibration is not possible, and electrochemical performance is estimated using a calibration factor determined ex vivo. As such, accurate quantification of neurochemical fluctuations remains a challenge.17

EIS is a powerful frequency-domain technique capable of providing remarkable insight into physical phenomena occurring at the electrochemical interface. It is routinely used to model and evaluate electrode performance in studies involving corrosion and energy storage.20,21

Although a few studies have investigated tissue fouling at metal and silicon microelectrodes using

EIS,22–26 these have relied on the assumption that the electrode state does not significantly change over the duration of the experiment.16,27,28 However, relatively little work has been done to investigate the micro- (or nano-) scale electrode/solution interface, especially at carbon-based electrodes. Further, microscopy-based studies demonstrate that CFME implantation elicits a biological response over time. Thus, understanding and mapping system-level impedance information onto the voltammetric response should be useful in compensating for real-world issues.

Herein, we utilized EIS to validate equivalent circuit models for two styles of CFME regularly used for biological investigations. The first style, which is typically used for acute measurements, involves heat pulling a borosilicate capillary to form an insulating glass seal around

182

the fiber (G-CFME).29 A more robust design for chronic implantation in an animal model uses a fused-silica capillary to insulate the carbon fiber, sealed with epoxy (C-CFME).30 These electrodes have been demonstrated to behave comparably,31 but likely deviate in impedance characteristics due to differences in insulation.32 To investigate this, both electrode subtypes were modeled using a combination of resistor and constant phase elements (CPE). The models were used to determine double-layer capacitance values for sensors of equal geometric surface area. The equivalent circuits were then used to investigate differences to the electrochemical system resulting from electrochemical conditioning of the CFME, the use of pitch-based vs PAN-based carbon fibers, changes in ionic strength, and the effects of tissue exposure. This work represents the first EIS study relating the impedance properties of CFMEs to voltammetric performance. The information gained can be used to improve accuracy when determining the identity and concentration of a target species in less-than-ideal, real-world conditions, especially when routine sensor calibration is challenging.

7.2 Experimental Section

7.2.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. In vitro electrochemical experiments were performed in phosphate-buffered saline (PBS; 10 mM Na2HPO4, 138 mM NaCl, and 2.7 mM KCl). Buffer was adjusted to pH 7.4 with 1 M NaOH and 1 M HCl. All aqueous solutions were made using ultrapure 18.2 MΩ water.

183

7.2.2 Electrode Fabrication

Cylindrical CFMEs were created using T-650 (7-µm diameter) or P-55 (10-µm diameter) carbon fibers (Cytec Industries, West Patterson, NJ), as previously described.29 Briefly, a single carbon fiber was aspirated into a borosilicate glass capillary (1.0mm x 0.5 mm, A-M Systems,

Carlsborg, WA). A P-22 micropipette puller (Narishige, Tokyo, Japan) was used to create a tapered glass seal. A subset of CFMEs was insulated with polyimide-coated, fused-silica tubing

(Polymicro Technologies, Phoenix, AZ; 164.7-μm outer diameter/98.6-μm inner diameter), as previously described.30 Briefly, the tubing was cut to 4 cm in length and placed in an aqueous 70% isopropyl alcohol solution. A T-650 carbon fiber was inserted and allowed to dry. A hemispherical epoxy seal (V.O. Baker Company, Mentor, OH) was created at one end, and electrodes were cured at 100 °C for 20 min. The protruding fiber was cut to a desired length, and electrical connection was established using conductive silver epoxy and a wire lead.

Disk microelectrodes were fabricated with P-55 fibers in a similar fashion as described for

CFMEs insulated in borosilicate glass, but were cut at the glass seal. Electrodes were dipped in an epoxy (Epoxy Technology Inc., Billerica, MA) for 20 s and cured overnight at 100 oC. Electrodes were polished at a 30o angle to create a planar, elliptical disk surface. An electrical connection was established by backfilling with a high ionic strength solution (4 M CH3CO2K and 150 mM KCl) and inserting a wire lead.

7.2.3 Data Acquisition

Unless otherwise specified, all electrodes were electrochemically conditioned in PBS using a triangular waveform that ranged from -0.4 to +1.3/1.4 V, with a scan rate of 400 V s-1 and a waveform application frequency of 60 Hz, for at least 15 min. The conditioning was performed using High Definition Cyclic Voltammetry software (HDCV, University of North Carolina at

184

Chapel Hill) in conjunction with data acquisition cards (National Instruments, Austin TX) to control and record data from a custom-built potentiostat (University of North Carolina at Chapel

Hill, Department of Chemistry, Electronics Facility). EIS measurements were collected using

Aftermath software (Pine Research Instrumentation, Durham, NC) in conjunction with a

WaveDriver 200 Bipoteniostat/Galvanostat (Pine Research Instrumentation, Durham, NC). EIS was performed using a series of 25 mV amplitude sinusoidal waveforms from 1 Hz to 1 MHz. All electrochemical measurements were performed inside a faraday cage using a Ag/AgCl counter/reference electrode.

7.2.4 Animal Experiments

Animal care and use was in accordance with North Carolina State University Institutional

Animal Care and Use Committee (IACUC) guidelines. A male Sprague−Dawley rat (250−300 g;

Charles River Laboratories, Raleigh, NC) was individually housed on a 12:12 h light/dark cycle with free access to food and water and allowed to acclimate to the facility for several days. The animal was isoflurane anesthetized (4 % induction maintained at ~1.5% during the experiment), placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes were drilled in the skull using stereotaxic coordinates.33 The working electrodes (n=5) were placed in the striatum (+1.2 mm AP, + 2.0 mm ML, −4.0 to -6.0 mm DV relative to bregma) and the

Ag/AgCl reference electrode was placed in the contralateral forebrain. Each subsequent CFME was lowered to a position more ventral than the previous to ensure that the carbon sensing surface was not positioned in the damaged tract. The animal’s body temperature was maintained at 37 °C by a heating pad.

185

7.2.5 Data Processing and Analysis

HDCV was used to evaluate voltammetric recordings and assess electrode viability.

Aftermath software was used to assess, process, and export EIS recordings, as well as to perform equivalent circuit modeling. Model optimization was performed using a modified Levenburg-

Marquardt algorithm with parametric weighting. A chi-square goodness of fit test was used to assess equivalent circuits describing the data. All statistical analyses and graphical depiction of the data were carried out using GraphPad Prism 6 (GraphPad Software, Inc., La Jolla, CA). t-tests were used to compare the means of two different groups. One-way ANOVA with a Dunnett’s post- test was used to compare changes from initial impedance properties during and following tissue exposure. Significance was determined at p<0.05. All error is reported as standard error of the mean (SEM), unless otherwise stated.

7.3 Results and Discussion

7.3.1 Overview of the Electrochemical System and EIS

The carbon-based microelectrodes typically employed in background-subtracted FSCV studies can be comprised of different carbon materials (fibers, nanotubes, doped-diamond, etc.) and fabricated in a variety of electrode geometries. These are used for measurements in a wide variety of recording applications; single-cells, tissue slice, flow-system, freely moving animal, bench-top separations, etc.2,12,34 Cylindrical CFMEs are most commonly used to monitor neurochemicals in live brain tissue, including but not limited to catecholamines, indolamines, reactive oxygen species, and peptides.35–39 Figure 1A displays a typical FSCV waveform, as well as typical background current responses collected at a potential-naïve electrode, and at the same electrode after electrochemical conditioning. Following conditioning, a redox couple is observed

186

(~0.1 V) due to the redox activity of ‘quinone-like’ moieties generated on the carbon surface during conditioning.40 In addition, the background is larger. This is indicative of increased capacitance, and can serve as a predictor of electrode sensitivity.41,42 A background-subtracted voltammogram for dopamine is also shown, with faradaic current resulting from the oxidation of dopamine to dopamine-σ-quinone (~0.5 V) and subsequent reduction back to dopamine. The overall shape of the voltammograms, the position of redox potentials, and analyte sensitivity are all influenced by the physical nature of the CFME/recording environment interface, which can be quantitatively characterized by monitoring impedance.19

Figure 7.1. Introduction to EIS and FSCV. A) Typical FSCV waveform (left). Background CVs (middle) are shown for a potential-naïve and an electrochemically conditioned CFME in the absence and presence of dopamine. A background-subtracted CV for dopamine is also shown (right). B) 10, 100, and 1000 Hz EIS potential waveforms (left) and current responses (right). C) An overlay of a 100 Hz potential waveform (left axis) and generated current (right axis) displaying the phase shift.

Electrochemical impedance spectroscopy (EIS) is a frequency-domain electroanalysis technique that typically employs a low-amplitude sinusoidal potential (potentiostatic) or current

(galvanostatic) waveform to characterize the state of a CFME in its performance environment.43 187

Herein, potentiostatic EIS was used in a two-electrode setup to make impedance measurements at

CFMEs over a range of 1 Hz to 1 MHz. With this approach, low currents were generated that did not significantly polarize the reference electrode (Ag/AgCl), and no significant non-linearity was observed in the Lissajous plots (not shown). Figure 1B displays representative potential waveforms

(25 mV amplitude) and current responses for 10, 100, and 1000 Hz measurements at a G-CFME.

The majority of electrode systems used in chemical sensing are dominated by capacitance, thus current precedes the applied voltage. The temporal lag between these curves (phase shift, Φ) and total impedance (Ohm’s law) can be determined by fitting both the applied voltage and resultant current to a sinusoidal function (Figure 1C). In this research, Fourier transform-based methods were used to determine impedance properties over a range of frequencies. These data were used to validate equivalent circuit models containing individual elements that were representative of the physical properties of the system. Thus, any changes to the overall system were reflected in one or more of the elements in the equivalent circuit(s).

7.3.2 Determining Equivalent Circuits for CFMEs

The most commonly used CFME constructions involve insulating carbon fibers using either: 1) a heat-pulled borosilicate glass capillary (G-CFME) or 2) a fused silica capillary with one end sealed using an epoxy resin (C-CFME). The graphical representation in Figure 2A depicts physical differences between these electrode types. Notably, there are differences in the geometry of the seal about the exposed carbon fiber, as well as the location where electrical connection is established; either near to, or far from, the exposed carbon. Figure 2B displays average Bode plots generated for these electrode subtypes and Figure 2C displays an average Nyquist plot (controlling for the length of exposed carbon fiber). In all plots, distinct differences are evident between the electrode subtypes. When CFME are insulated with a silica capillary, electrical connectivity is

188

established at the top of the fused silica, requiring current to transmit through a long segment of carbon. This is evident in the magnitude of the impedance response at high frequencies. The phase angle of the frequency response also deviates between these electrode subtypes at the high frequencies. There are two observable peaks in the phase Bode plot for glass-insulated CFME, indicating an additional capacitive contributor to this system. This may be due to two possible phenomena: 1) an imperfect seal or porous glass that allows some small amount of solution to accumulate between the carbon fiber and glass seal, or 2) the glass insulation is sufficiently thin that a double layer is induced across some portion of the seal when potential is applied.32 Glass electrode seals can vary in taper length and thickness, and can visually appear to be ‘perfect’ even though gaps may exist between the carbon fiber and glass. As such, the contribution of this feature to the impedance data is variable. Figure 2C displays an average Nyquist plot for each electrode subtype. Figure 2D plots the capacitive reactance versus frequency for the two different CFME subtypes. Overall, these data were used to determine capacitance (C)

푋 = √푍2 − 푍2 푋 (푓) = 1 푐 푡 푟 푐 2휋푓퐶

where Xc is capacitive reactance, Zt is total impedance, Zr is real impedance, and f is frequency; frequency dependent reactance was fit using nonlinear regression to determine C.43 Capacitance values of 2.32 nF and 1.83 nF were determined for glass-insulated and silica-insulated electrodes, respectively, consistent with an additional capacitance contribution inherent to the glass-insulated electrode/solution interface. Overall, these impedance data demonstrate that the method of insulation impacts CFME performance. This can be modeled using equivalent circuits to gain additional insight into the system.

189

Figure 7.2. EIS performed from 1 Hz to 1 MHz at CFMEs insulated in glass (red; n=16) and silica (black; n=11). A) Graphic depicting physical differences between these electrode subtypes. B) Average impedance (left) and phase (right) Bode plots. C) Average Nyquist plot. Inset is an expanded view from 100 Hz to 1 MHz (dashed box). D) Plot of capacitive reactance versus frequency, as well as relevant equations for determining capacitance.

When fitting impedance data with equivalent circuits, multiple combinations of resistance

(R), constant phase elements (Q), and Warburg elements can be used to adequately fit the data. A constant phase element is used to model a non-ideal capacitor and is composed of two properties: capacitance and a modifier, alpha (α); when α=1, a constant phase element behaves as an ideal capacitor. Non-ideal capacitor behavior is common in real electrochemical systems, and is caused by a myriad of phenomena generally related to heterogeneity in the system, such as surface roughness and defects.44–47 A Warburg element models diffusional processes at low frequencies; however, a Warburg element is unnecessary here because no portion of the Nyquist plot displays a 45° line when the imaginary and real axes are scaled similarly.44,48 Thus, four circuit

190

combinations composed of resistors (R) and constant phase elements (Q) could be used to fit the impedance data shown in Figure 2: 1) R1+Q1, 2) R1+Q1/(R2+Q2), 3) R1+(Q1/R2)+(Q2/R3), and 4)

R1+Q1/(R2+(Q2/R3)). Due to variability associated with measurements at the highest and lowest frequencies, the impedance data were fit from 5 Hz to 500 kHz. The most appropriate equivalent circuits were determined by attributing each circuit element to a physical property of the system,

2 2 and a chi-square goodness-of-fits (X

Additional circuit elements were added only when incorporation into the system served to decrease

X2 by at least an order of magnitude.43,49 The normalized residual error of the fit to both the imaginary and real impedance components were also examined. Data collected at the silica- insulated electrodes could be fit with any of the above circuits to generate reasonable X2 values.

Only circuits 2,3, and 4 could reasonably fit the glass-insulated electrodes. However, there was no rationale for inclusion of three resistors in the circuit, as the additional elements did not substantially decrease X2. Thus, circuits 1 and 2 were selected for further analyses.

Figure 3A displays the two equivalent circuit models used to fit the data, as well as the physical characteristic described by each element. R1 represents a series addition of the uncompensated resistance attributed to the electrode and solution. Q1 represents the double-layer capacitance associated with the exposed carbon fiber. R2 represents the resistance associated with leakage current at the electrode seal; a higher value indicates a lower leakage current. Q2 represents the capacitance associated with the insulating seal about the carbon fiber. Figure 3B displays the average impedance data for both electrode constructs, and the fits to those data generated using these equivalent circuits. Both circuits adequately fit the data collected using the silica-insulated electrodes; however, circuit 2 has a negligible Q2 value. As such, R2 has no discernable effect on the fit, and can be set infinitely high or low with no effect. The impedance data for the glass-

191

insulated electrodes are not effectively modeled with circuit 1. This model fits the double-layer capacitance contribution well, but the seal capacitance, Q2, is not accounted for unless circuit 2 is used to fit the data.

The individual circuit elements determined by these models are quantitatively compared in

Figure 3C. The Q1 and α1 values are similar across electrode subtypes, because the nature of the carbon sensing material is equivalent in both fabrications. The modulator, α1, was determined to be ~0.9 which is in good agreement with previous studies at oxidized graphitic carbon.50,51 The deviation from ideality has been attributed to surface roughness, and it is well established that

52,53 carbon fiber materials possess various defects and striations. By contrast, Q2 exhibits nearly ideal capacitor behavior (α2 ~0.98), because the glass surface is a more homogenous surface than that of the carbon fiber. As expected, the model for the silica-insulated electrodes generates a higher R1 value than that for the glass-insulated electrodes, due to the longer length of carbon fiber between the electrical connection and the exposed sensing surface (Figure 2A). Finally, the Q1 and

-2 R1 values were used to determine specific capacitance values (µF cm ) of 87.9 and 90.5, and RC time constants (µs) of 24.1 and 51.7, for glass- and silica-insulated CFME, respectively. These values are quantitatively compared in Figure 3D. The RC time constants are generally in agreement with previously reported values for cylindrical CFMEs (39.2 µs).54 The specific capacitance values are higher than previously reported for basal (<10 µF cm-2) and edge-plane carbon (~60 µF cm-2) substrates.12,51 However, they are in agreement with values reported for measurements at electrochemically conditioned carbon-fiber mircroelectrodes (~40-120 µF cm-2).54,55

192

Figure 7.3. Equivalent model circuits used to fit EIS data. A) Model circuits and a description of the physical significance of each resistance (R) and constant phase (Q) element. B) Average impedance Bode (left), phase Bode (middle), and Nyquist (right) plots for the glass-insulated (top; n=16) and silica-insulated (bottom; n=11) CFMEs overlaid with fits performed using both equivalent circuits. C) Quantitative comparison of equivalent circuit values generated in fitting impedance data. D) Specific capacitance values (left) and RC time constants (right) calculated using R1 and Q1 fit values.

193

7.3.3 CFME Equivalent Circuits Inform on the Nature of the Electrode/Solution Interface

In order to evaluate the utility of a given model in describing the electrochemical system, the ionic strength of the electrolyte, the length of exposed carbon, and the nature of the carbon surface were systematically varied. Increasing the ionic strength of the electrolyte serves to decrease the diffuse layer thickness and increase double-layer capacitance, as described by the

56 Guoy-Chapman model of the electrical double layer. The Q1 values determined by fitting EIS data for glass-insulated, T-650 CFMEs are consistent with theory (Figure 4A and 4B). Debye length was estimated by substituting ionic strength in place of electrolyte concentration and charge

57 for a single species. It should be noted that the α1 values did not significantly vary in response to this manipulation, supporting the idea that deviations from an ideal capacitor are due to surface inhomogeneities rather than environmental factors. As expected, R1 was inversely related to ionic strength.

The mechanical and electrical properties of a carbon fiber dictate electrochemical performance.11,52,58 Pitch- and polyacrylonitrile (PAN)-based carbon fibers are both routinely used to construct CFMEs. Briefly, pitch-based carbon fibers have a high carbon content and an ordered graphitic structure that is oriented parallel to the main axis of the fiber. PAN-based fibers can vary in carbon content, and exhibit a more turbostriatic structure. As a result, pitch-based fibers are generally more brittle, exhibit greater conductivity, and result in a more homogenous and smooth surface than PAN-based carbon fibers.52,53 Cylindrical CFMEs are most often created from T-650 carbon fibers (PAN-based, ~7-µm diameter), whereas disk CFMEs are generally created using P-

55 fibers (pitch-based, ~10-µm diameter). Figure 4B displays a linear relationship between the geometric surface area and the modeled Q1 values for both types of electrodes (insulated in glass).

194

Regardless of the base carbon material, it is common to perform some pretreatment to activate the carbon electrode surface prior to use. EIS recordings of pristine and electrochemically conditioned (+1.3 V anodic limit), glass-insulated CFMEs were fit with equivalent circuit 2 to determine circuit element values. Figure 4C displays a comparison of the double-layer capacitor terms, Q1 (specific capacitance; left) and α1 (middle), as well as the combined solution and carbon resistance, R1 (right). Electrochemical conditioning of the T-650 carbon fibers resulted in a significant increase in specific capacitance values (Q1 (left), t(6)paired=4.95, **p=0.0026), supporting previous work documenting surface renewal via etching. A direct comparison of T-650 and P-55 CFMEs that were electrochemically conditioned in an identical manner (exposed carbon surfaces of equal length) reveals a greater specific capacitance associated with the PAN-based material (t(10)unpaired=2.971, *p=0.014). These measurements support imaging studies that have demonstrated a greater surface roughness associated with PAN-based fibers.52,53 It should be noted that, in this experiment, electrodes were conditioned using a +1.3 V waveform limit; whereas the data displayed in Figure 2 were collected for electrodes that were conditioned using a +1.4 V waveform limit. This likely accounts for the difference in specific capacitance evident across these data sets, as well as differences in sensitivity achieved through more oxidative conditioning. A comparison of the α1 values also supports this assertion. While electrochemical conditioning did not significantly impact α1 (t(6)paired=1.491, p=0.19, n.s.), the more structurally ordered P-55 material generated a larger α1 value (t(10)unpaired=2.97, *p=0.014), indicating more ideal capacitor behavior. The values of α1 determined herein are consistent with studies performed on other

50,51 graphitic materials. Additionally, a comparison of R1 demonstrates that the P-55 material is more conductive than the less ordered T-650 material (t(10)unpaired=7.75, ****p<0.0001), which is confirmed by manufacturer characterization. Taken together, these measurements demonstrate that

195

the equivalent circuit model is capable of accounting for controlled manipulations to the electrochemical system, in a manner that is consistent with theory and literature.

Figure 7.4. The equivalent circuit model can be used to accurately report on controlled manipulations to the electrochemical system. A) Constant phase element and resistance values for glass-insulated CFMEs as a function of ionic strength (left and right), and the approximated Debye length (middle); all x-axes plotted on a log10 scale. B) Constant phase element value as a function of geometric surface area. Black and red data points generated using glass-insulated cylindrical T- 650 and disc P-55 CFMEs, respectively. C) Fit values for pristine (black) and electrochemically conditioned (gray) T-650 CFMEs, and for conditioned P-55 CFMEs (white) of equal length.

7.3.4 Tissue Exposure Shifts Impedance

Although carbon-fiber microelectrodes are relatively resistant to biofouling, it is well established that electrochemical performance is impacted by the recording environment.18,19,41 A number of studies have investigated the impedance properties of metal and silicon microelectrodes following tissue exposure;22–25 however, there is a lack of comparable literature exploring carbon

196

microelectrode impedance. Thus, EIS recordings were performed at silica-insulated CFMEs prior to, during, and following tissue implantation. Electrodes were positioned in the striatum of an anesthetized rat and allowed to rest for approximately 15-20 min, after which impedance was measured. The electrodes were removed from tissue and impedance was immediately assessed in

PBS, and again following benchtop electrochemical conditioning for 10 min. Figure 5A displays impedance, phase, and capacitance Bode plots, as well as Nyquist plots at each stage of the experiment. It is evident that tissue exposure substantially changes the impedance response. For example, examination of the high-frequency range of the impedance Bode plot and the entirety of the Nyquist plot reveals a significant increase in electrode/solution resistance and real impedance in tissue, and immediately following explant. Interestingly, the system is reversible. Following electrochemical conditioning, the impedance characteristics revert to a state resembling that of the pre-implant condition. The duration of the FSCV waveform is ~10 ms, which corresponds to an

EIS waveform frequency of 100 Hz. Figure 5B quantitatively compares the impedance and capacitance information collected at 100 Hz. Exposure to tissue induces a significant increase in impedance (F(3,12)=17.04, ****p<0.0001) and capacitive reactance (F(3,12)=15.96,

***p=0.0002), which ultimately normalize following additional electrochemical conditioning on the benchtop (repeated-measures one-way ANOVA with Dunnett’s post-test comparing all groups to the pre-implant values).

These impedance data were fit to equivalent circuit 1 in order to gain more information regarding the nature of the observed impedance changes. Figure 5C quantitatively compares Q1,

α1, and R1, all of which are significantly affected by the recording environment, as determined via a repeated-measures one-way ANOVA with Dunnett’s post-test comparing all groups to the pre- implant values; Q1: F(3,12)=14.3, ***p=0.0003; α1: F(3,12)=36.24, ****p<0.0001; R1:

197

F(3,12)=13.1, ***p=0.0004. The combined carbon/solution resistance increases during implantation. This is likely due to increased tortuosity inherent to the tissue environment (as compared to buffer), as well as to physical barriers to complete the electrochemical system (the reference/counter electrode is placed in the opposite brain hemisphere).22,24 Surprisingly, a decrease in Q1 was not observed upon implantation. The ionic strength inherent to the biological recording environment may be higher than that in buffer, due to the presence of myriad charged neurotransmitters, peptides, proteins, etc. However, immediately following removal from tissue, the measurement in PBS solution reveals a decrease in capacitance, suggesting that implantation decreased effective surface area. This is further supported by the measured decrease in α1 upon implantation and immediately following explant. Deviation from ideal capacitance increased in the heterogeneous biological recording environment, likely due to biofouling. Q1 and α1 returned to pre-implant values following electrochemical conditioning, as this serves to at least partially etch and renew the carbon surface.59 These Q-values are consistent with calculations of capacitance generated by fitting the capacitive reactance data displayed in Figure 5A; 1.37 nF (pre- implant), 1.56 nF (in tissue), 1.05 nF (post-implant), and 1.51 nF (post-implant following electrochemical conditioning). Overall, the data and results presented in Figure 5 demonstrate significant changes to the electrochemical system with tissue exposure, which undoubtedly affect electrochemical performance. The equivalent circuit model fits provide a quantitative description of the physical phenomena that lead to these changes. Importantly, electrochemical conditioning, which occurs continually during FSCV recordings, mitigates these effects.

198

Figure 7.5. Tissue exposure impacts impedance properties. A) Impedance (top-left) and phase (top-right) Bode plots, as well as Nyquist (bottom-left) and capacitive reactance versus frequency (bottom-right) plots for silica-insulated CFME (n=5). Data were collected in buffer prior to implantation in the striatum, while implanted in tissue, and in buffer following removal from tissue before and after additional electrochemical conditioning. B) 100 Hz impedance (top) and capacitive reactance (bottom) values extracted from the data displayed in A). C) Equivalent circuit fit values to individual electrode EIS data.

7.4 Conclusion

The properties of the microelectrode/recording environment interface, where the heterogeneous electron transfer occurs, directly impact electrochemical performance (sensitivity and selectivity).17–19,41 Indeed, system impedance and double-layer capacitance can serve as vital predictors of electrochemical performance for FSCV.19,41,42,58 Herein, EIS measurements were

199

used to develop equivalent circuit models describing the physical nature of the CFME/solution interface, at any stage of an in vivo experiment. Systematic manipulation of the CFME and the recording environment validated the circuit models, and also demonstrated the utility of EIS to provide information relevant to electrochemical sensing using FSCV. The data show that the carbon fiber material, the manner in which it is insulated, and the physical nature of the recording environment impact impedance properties in a manner consistent with theory. Implantation results in a significant increase in impedance and decrease in capacitance, as well significant deviation from ideal capacitance behavior at the electrical double-layer, presumably due to increased surface heterogeneity. This in-depth characterization of CFME impedance properties before, during, and after implantation represents the first study employing EIS to investigate the effect of tissue implantation on CFME electrical properties. The findings lay the framework for accurate determination of performance factors for sensitivity (signal amplitude) and selectivity

(voltammogram shape), at any point in the course of a long-term experiment.

200

7.5 References

1. Adams, R. N. In Vivo Electrochemical Measurements in the CNS. Prog. Neurobiol. 1990, 35 (4), 297–311. https://doi.org/10.1016/0301-0082(90)90014-8.

2. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, Boca Raton, FL, 2006.

3. Day, J. J.; Roitman, M. F.; Wightman, R. M.; Carelli, R. M. Associative Learning Mediates Dynamic Shifts in Dopamine Signaling in the Nucleus Accumbens. Nat. Neurosci. 2007, 10 (8), 1020–1028. https://doi.org/10.1038/nn1923.

4. Gan, J. O.; Walton, M. E.; Phillips, P. E. M. Dissociable Cost and Benefit Encoding of Future Rewards by Mesolimbic Dopamine. Nat. Neurosci. 2010, 13 (1), 25–27. https://doi.org/10.1038/nn.2460.

5. Syed, E. C. J.; Grima, L. L.; Magill, P. J.; Bogacz, R.; Brown, P.; Walton, M. E. Action Initiation Shapes Mesolimbic Dopamine Encoding of Future Rewards. Nat. Neurosci. 2016, 19 (1), 34–36. https://doi.org/10.1038/nn.4187.

6. Howe, M. W.; Tierney, P. L.; Sandberg, S. G.; Phillips, P. E. M.; Graybiel, A. M. Prolonged Dopamine Signalling in Striatum Signals Proximity and Value of Distant Rewards. Nature. 2013, 500 (7464), 575–579. https://doi.org/10.1038/nature12475.

7. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2013, 4 (5), 782–789. https://doi.org/10.1021/cn4000499.

8. Mellander, L. J.; Trouillon, R.; Svensson, M. I.; Ewing, A. G. Amperometric Post Spike Feet Reveal Most Exocytosis Is via Extended Kiss-and-Run Fusion. Sci. Rep. 2012, 2 (1), 907. https://doi.org/10.1038/srep00907.

9. Miles, P. R.; Mundorf, M. L.; Wightman, R. M. Release and Uptake of Catecholamines in the Bed Nucleus of the Stria Terminalis Measured in the Mouse Brain Slice. Synapse. 2002, 44 (3), 188–197. https://doi.org/10.1002/syn.10069.

10. Li, Y.-T.; Zhang, S.-H.; Wang, L.; Xiao, R.-R.; Liu, W.; Zhang, X.-W.; Zhou, Z.; Amatore, C.; Huang, W.-H. Nanoelectrode for Amperometric Monitoring of Individual Vesicular Exocytosis Inside Single Synapses. Angew. Chem. Int. Ed. 2014, 53 (46), 12456–12460. https://doi.org/10.1002/anie.201404744.

11. Huffman, M. L.; Venton, B. J. Electrochemical Properties of Different Carbon-Fiber Microelectrodes Using Fast-Scan Cyclic Voltammetry. Electroanalysis 2008, 20 (22), 2422– 2428. https://doi.org/10.1002/elan.200804343.

12. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108 (7), 2646–2687. https://doi.org/10.1021/cr068076m.

201

13. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir. 2010, 26 (11), 9116– 9122. https://doi.org/10.1021/la9048924.

14. Strein, T. G.; Ewing, A. G. In Situ Laser Activation of Carbon Fiber Microdisk Electrodes. Anal. Chem. 1991, 63 (3), 194–198. https://doi.org/10.1021/ac00003a001.

15. Strand, A. M.; Venton, B. J. Flame Etching Enhances the Sensitivity of Carbon-Fiber Microelectrodes. Anal. Chem. 2008, 80 (10), 3708–3715. https://doi.org/10.1021/ac8001275.

16. Kozai, T. D. Y.; Jaquins-Gerstl, A. S.; Vazquez, A. L.; Michael, A. C.; Cui, X. T. Brain Tissue Responses to Neural Implants Impact Signal Sensitivity and Intervention Strategies. ACS Chem. Neurosci. 2015, 6 (1), 48–67. https://doi.org/10.1021/cn500256e.

17. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2016, 7 (3), 349–359. https://doi.org/10.1021/acschemneuro.5b00302.

18. Logman, M. J.; Budygin, E. A.; Gainetdinov, R. R.; Wightman, R. M. Quantitation of in Vivo Measurements with Carbon Fiber Microelectrodes. J. Neurosci. Methods. 2000, 95 (2), 95– 102. https://doi.org/10.1016/S0165-0270(99)00155-7.

19. Meunier, C. J.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Background Signal as an in Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2017, 8 (2), 411–419. https://doi.org/10.1021/acschemneuro.6b00325.

20. Lasia, A. Electrochemical Impedance Spectroscopy and Its Applications. In Modern aspects of electrochemistry; Springer: New York, 2002; pp 143–248.

21. Chang, B.-Y.; Park, S.-M. Electrochemical Impedance Spectroscopy. Annu. Rev. Anal. Chem. 2010, 3 (1), 207–229. https://doi.org/10.1146/annurev.anchem.012809.102211.

22. Williams, J. C.; Hippensteel, J. A.; Dilgen, J.; Shain, W.; Kipke, D. R. Complex Impedance Spectroscopy for Monitoring Tissue Responses to Inserted Neural Implants. J. Neural Eng. 2007, 4 (4), 410–423. https://doi.org/10.1088/1741-2560/4/4/007.

23. Xiaoqiu Huang; Duc Nguyen; D. W. Greve; M. M. Domach. Simulation of Microelectrode Impedance Changes Due to Cell Growth. IEEE Sens. J. 2004, 4 (5), 576–583. https://doi.org/10.1109/JSEN.2004.831302.

24. A. Mercanzini*; P. Colin; J. Bensadoun; A. Bertsch; P. Renaud. In Vivo Electrical Impedance Spectroscopy of Tissue Reaction to Microelectrode Arrays. IEEE Trans. Biomed. Eng. 2009, 56 (7), 1909–1918. https://doi.org/10.1109/TBME.2009.2018457.

202

25. Lempka, S. F.; Miocinovic, S.; Johnson, M. D.; Vitek, J. L.; McIntyre, C. C. In Vivoimpedance Spectroscopy of Deep Brain Stimulation Electrodes. J. Neural Eng. 2009, 6 (4), 046001. https://doi.org/10.1088/1741-2560/6/4/046001.

26. McAdams, E. T.; Jossinet, J. Tissue Impedance: A Historical Overview. Physiol. Meas. 1995, 16 (3A), A1–A13. https://doi.org/10.1088/0967-3334/16/3a/001.

27. Peters, J. L.; Miner, L. H.; Michael, A. C.; Sesack, S. R. Ultrastructure at Carbon Fiber Microelectrode Implantation Sites after Acute Voltammetric Measurements in the Striatum of Anesthetized Rats. J. Neurosci. Methods 2004, 137 (1), 9–23. https://doi.org/10.1016/j.jneumeth.2004.02.006.

28. Schwerdt, H. N.; Shimazu, H.; Amemori, K.; Amemori, S.; Tierney, P. L.; Gibson, D. J.; Hong, S.; Yoshida, T.; Langer, R.; Cima, M. J.; et al. Long-Term Dopamine Neurochemical Monitoring in Primates. Proc. Natl. Acad. Sci. 2017, 114 (50), 13260. https://doi.org/10.1073/pnas.1713756114.

29. Roberts, J. G.; Lugo-Morales, L. Z.; Loziuk, P. L.; Sombers, L. A. Real-Time Chemical Measurements of Dopamine Release in the Brain. In Dopamine - Methods and Protocols; Kabbani, N, Ed.; Humana Press: Totowa, NJ, 2013; p 275-294. https://doi.org/10.1007/978-1- 62703-251-3_16.

30. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V.; et al. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat. Methods. 2010, 7 (2), 126–129. https://doi.org/10.1038/nmeth.1412.

31. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E. M.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234. https://doi.org/10.1021/acschemneuro.6b00393.

32. Wipf, D. O. Ohmic Drop Compensation in Voltammetry: Iterative Correction of the Applied Potential. Anal. Chem. 1996, 68 (11), 1871–1876. https://doi.org/10.1021/ac951209b.

33. Paxinos, G.; Watson, C. The Rat Brain in Stereotaxic Coordinates: Hard Cover Edition; Elsevier, 2006.

34. Yang, C.; Denno, M. E.; Pyakurel, P.; Venton, B. J. Recent Trends in Carbon Nanomaterial- Based Electrochemical Sensors for Biomolecules: A Review. Anal. Chim. Acta 2015, 887, 17– 37. https://doi.org/10.1016/j.aca.2015.05.049.

35. Calhoun, S. E.; Meunier, C. J.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin. ACS Chem. Neurosci. 2019, 10 (4), 2022–2032. https://doi.org/10.1021/acschemneuro.8b00351.

203

36. Hensley, A. L.; Colley, A. R.; Ross, A. E. Real-Time Detection of Melatonin Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2018, 90 (14), 8642–8650. https://doi.org/10.1021/acs.analchem.8b01976.

37. Abdalla, A.; Atcherley, C. W.; Pathirathna, P.; Samaranayake, S.; Qiang, B.; Peña, E.; Morgan, S. L.; Heien, M. L.; Hashemi, P. In Vivo Ambient Serotonin Measurements at Carbon-Fiber Microelectrodes. Anal. Chem. 2017, 89 (18), 9703–9711. https://doi.org/10.1021/acs.analchem.7b01257.

38. Samaranayake, S.; Abdalla, A.; Robke, R.; Wood, K. M.; Zeqja, A.; Hashemi, P. In Vivo Histamine Voltammetry in the Mouse Premammillary Nucleus. Analyst 2015, 140 (11), 3759– 3765.

39. Lee, S. T.; Venton, B. J. Regional Variations of Spontaneous, Transient Adenosine Release in Brain Slices. ACS Chem. Neurosci. 2018, 9 (3), 505–513. https://doi.org/10.1021/acschemneuro.7b00280.

40. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Bucher, E. S.; McCarty, G. S.; Wightman, R. M. Characterization of Local PH Changes in Brain Using Fast-Scan Cyclic Voltammetry with Carbon Microelectrodes. Anal. Chem. 2010, 82 (23), 9892–9900. https://doi.org/10.1021/ac102399n.

41. Roberts, J. G.; Tou; p ps, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode -Calibration Strategy for Voltammetric Measurements In Vivo. Anal. Chem. 2013, 85 (23), 11568–11575. https://doi.org/10.1021/ac402884n.

42. Schuweiler, D. R.; Howard, C. D.; Ramsson, E. S.; Garris, P. A. Improving in Situ Electrode Calibration with Principal Component Regression for Fast-Scan Cyclic Voltammetry. Anal. Chem. 2018, 90 (22), 13434–13442. https://doi.org/10.1021/acs.analchem.8b03241.

43. Orazem, M. E.; Tribollet, B. Electrochemical Impedance Spectroscopy; John Wiley & Sons: Hoboken, NJ, 2017.

44. Gamry Instruments. Basics of Electrochemical Impedance Spectroscopy. https://www.gamry.com/application-notes/EIS/basics-of-electrochemical- impedancespectroscopy/, accessed 12-1-19.

45. Córdoba-Torres, P.; Mesquita, T. J.; Devos, O.; Tribollet, B.; Roche, V.; Nogueira, R. P. On the Intrinsic Coupling between Constant-Phase Element Parameters α and Q in Electrochemical Impedance Spectroscopy. Electrochimica Acta 2012, 72, 172–178. https://doi.org/10.1016/j.electacta.2012.04.020.

46. Alexander, C. L.; Tribollet, B.; Orazem, M. E. Contribution of Surface Distributions to Constant-Phase-Element (CPE) Behavior: 1. Influence of Roughness. Electrochimica Acta 2015, 173, 416–424. https://doi.org/10.1016/j.electacta.2015.05.010.

204

47. Alexander, C. L.; Tribollet, B.; Orazem, M. E. Contribution of Surface Distributions to Constant-Phase-Element (CPE) Behavior: 2. Capacitance. Electrochimica Acta 2016, 188, 566–573. https://doi.org/10.1016/j.electacta.2015.11.135.

48. Taylor, S. R.; Gileadi, E. Physical Interpretation of the Warburg Impedance. CORROSION 1995, 51 (9), 664–671. https://doi.org/10.5006/1.3293628.

49. Orazem, M. E.; Agarwal, P.; Garcia-Rubio, L. H. Critical Issues Associated with Interpretation of Impedance Spectra. J. Electroanal. Chem. 1994, 378 (1), 51–62. https://doi.org/10.1016/0022-0728(94)87056-X.

50. Sarac, A. S.; Ates, M.; Kilic, B. Electrochemical Impedance Spectroscopic Study of Polyaniline on Platinum, Glassy Carbon and Carbon Fiber Microelectrodes. Int J Electrochem Sci 2008, 3 (7), 777–786.

51. Kim, C.-H.; Pyun, S.-I.; Kim, J.-H. An Investigation of the Capacitance Dispersion on the Fractal Carbon Electrode with Edge and Basal Orientations. Electrochimica Acta 2003, 48 (23), 3455–3463. https://doi.org/10.1016/S0013-4686(03)00464-X.

52. Edie, D. D. The Effect of Processing on the Structure and Properties of Carbon Fibers. Carbon 1998, 36 (4), 345–362. https://doi.org/10.1016/S0008-6223(97)00185-1.

53. Hoffman, W. P.; Hurley, W. C.; Liu, P. M.; Owens, T. W. The Surface Topography of Non- Shear Treated Pitch and PAN Carbon Fibers as Viewed by the STM. J. Mater. Res. 1991, 6 (8), 1685–1694. https://doi.org/10.1557/JMR.1991.1685.

54. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal. Chem. 2017, 89 (11), 6166–6174. https://doi.org/10.1021/acs.analchem.7b01005.

55. Feng, J. Xing.; Brazell, Michael.; Renner, Kenneth.; Kasser, Richard.; Adams, R. N. Electrochemical Pretreatment of Carbon Fibers for in Vivo Electrochemistry: Effects on Sensitivity and Response Time. Anal. Chem. 1987, 59 (14), 1863–1867. https://doi.org/10.1021/ac00141a028.

56. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; Wiley: New York, 2000.

57. Israelachvili, J. N. Intermolecular and Surface Forces, 3rd ed.; Academic press, 2011.

58. Pandolfo, A. G.; Hollenkamp, A. F. Carbon Properties and Their Role in Supercapacitors. J. Power Sources 2006, 157 (1), 11–27. https://doi.org/10.1016/j.jpowsour.2006.02.065.

59. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal. Chem. 2010, 82 (5), 2020–2028. https://doi.org/10.1021/ac902753x.

205

CHAPTER 8

Tracking Carbon Microelectrode Impedance Time-Locked with Fast-Scan Cyclic

Voltammetry

The work was completed in collaboration with Gregory S. McCarty and Leslie A. Sombers, and is in preparation for submission to the journal ACS Sensors.

8.1 Introduction

Background-subtracted fast-scan cyclic voltammetry (FSCV) employing carbon-fiber microelectrodes (CFME) is used predominately in neuroscience to monitor subsecond fluctuations of molecular species in freely moving animals and other biological preparations with high spatial resolution.1–3 When combined with behavioral and pharmacological paradigms, FSCV has provided remarkable information about the molecular mechanisms underlying specific aspects of goal-directed behavior and associative learning.4–8 Difficulties associated with the use of FSCV and CFMEs have inhibited reliable quantification of long-term measurements. Yet, the ability to make reliable long-term (weeks to >1 year) voltammetric recordings at chronically implanted electrodes in freely moving animals will be necessary for a myriad of studies within neuroscience and behavioral psychology.6,9–13

CFME performance is dependent on physical characteristics inherent to the handmade electrodes (surface area, carbon-fiber type, surface functionalities),14–18 and the recording environment (composition, fouling, encapsulation).14,19–24 Furthermore, advances in electrodes and waveform development have resulted in FSCV recordings that often contain signal contributions from multiple analytes or interferents. As a result, electrode-specific multivariate calibration is

206

required to accurately relate the observed current to analyte concentration.25–30 In many early works involving FSCV, measurements were made with acutely implanted (<8 hours) electrodes, after which the electrode was removed for post-calibration in a flow-injection system. Although this calibration method does not account for the involvement of spectator species and changes in surface state, calibrations were electrode-specific, and proved to be effective. Long-term measurements involve permanently implanted electrodes, and thus electrode post-calibration becomes extremely difficult. As a result, it has become common to perform analyses using calibration factors (sensitivities) and training sets (cyclic voltammograms) from other electrodes and recording environments.7,10,30,31 Advances in multivariate calibration validation methods have proved this manner of calibration generally effective. However, this method is labor and time intensive, can result in significant misestimation of analyte concentrations, and assumes electrode performance is constant throughout the duration of the experiment.

It has been demonstrated that implantation results in electrode fouling (working and reference),22,24,32 tissue encapsulation due to biological response to the implantation,21,33,34 and changes to cyclic voltammogram (CV) signatures for an analyte.31,35 Electrochemical impedance spectroscopy (EIS) is a powerful frequency domain electrochemical technique used in applications ranging from corrosion studies, testing energy storage devices, biosensors, studying interfacial phenomena, and characterizing resistive and capacitive properties of materials.36–38 Impedance based studies at CFMEs are very limited, and inferences regarding the impact of biological response and fouling on an electrode’s electrical properties are based on studies involving metal and silicon microelectrodes.13,21,34 Recently, a convolution-based method employing a small potential-step was used to probe CFME impedance state.39 This was shown to be effective in removing FSCV background-subtraction artifacts resulting from changes in local ionic

207

composition impacting charging of the electrical double-layer; albeit computationally complex.

Thus, the ability to make simple impedance measurement to track CFME electrical properties would provide feedback on background-subtraction artifacts and variable electrochemical performance.

Recent work from our group has worked towards this goal by developing in situ methods utilizing the CFME background signal, generally discarded following subtraction, for predicting performance factors of CFMEs.22,35 Herein, we developed a multiplexed measurement paradigm for performing time-locked FSCV and EIS measurements. To accomplish this, FSCV hardware performance for EIS was evaluated over a range of frequencies and demonstrated to provide valuable feedback on CFME performance. Next, a controllable analog switch was used to combine multimodal measurements into a single potential scan without sacrificing temporal resolution commonly employed in FSCV experiments. These measurements permitted evaluation of the effects of electrochemical preconditioning and implantation into live tissue on the impedance properties of the electrochemical system. Notably, it is demonstrated that implantation results in a significant impedance and reactance increase, and capacitance decrease. These changes are mitigated when an electrode is continuously conditioned electrochemically. These data demonstrate that impedimetric measurements provide feedback regarding changes in an electrochemical systems performance, which will be valuable for interpreting long-term recordings and would be supplemental to developing real-time calibration strategies.

208

8.2 Experimental Section

8.2.1 Data Acquisition

Electrochemical impedance spectroscopic were recorded from 1 Hz-1 MHz with a 25 mV amplitude sinusoidal waveform with a DC-offset of 0.0 V using a WaveDriver200 potentiostat

(Pine Research Instrumentation, Inc.; Durham, NC) in a two-electrode setup; counter and reference electrode leads connected to a Ag/AgCl pellet reference. Pine Aftermath software was used for instrument control.

FSCV and EIS measurements were performed using two different FSCV potentiostat systems. High Definition Cyclic Voltammetry software (HDCV, University of North Carolina at

Chapel Hill) was used for waveform output, signal processing, and data analysis, in conjunction with data acquisition cards (National Instruments, Austin TX) used for measuring current and synchronizing the electrochemical cell with the flow-injection system. All measurements were recorded using a sampling frequency of 10, 25, or 100 kHz. FSCV measurements were made with a triangular waveform scanning from -0.4 V to either a +1.0, +1.2, +1.3, or +1.4 V switching potential, and back to -0.4 V with a scan rate of 400 V s-1, and an application frequency of 10 Hz.

The potential was held at -0.4 V between measurements.

Non-multiplexed EIS measurements were performed using a one second waveform consisting of a 0.0 V DC potential superimposed with a 25 mV amplitude sinusoidal waveform, an application frequency of 1 Hz, and a sampling frequency of 100 kHz. The waveform frequency was controlled by changing the number of periods applied in a second. Potential application and current transduction for these measurements were accomplished using a WaveNeuro potentiostat

(Pine Research Instrumentation, Inc.; Durham, NC). All FSCV and EIS data displayed from this setup employed no explicit filtering.

209

Multiplexed FSCV-EIS measurements were performed using a custom-built, multi- channel instrument for potential application and current transduction (Universal Electrochemistry

Instrument, University of North Carolina at Chapel Hill, Department of Chemistry, Electronics

Facility). One output channel supplied the FSCV waveform. A second output supplied the EIS waveform, which consisted of a -0.4 V DC potential superimposed with a 70 ms, 25 mV amplitude sinusoidal waveform that was delayed 20 ms with respect to the beginning of the FSCV waveform.

EIS frequency was controlled by changing the number of periods applied. A single, combined waveform output was accomplished using an ADG1633 CMOS analog switch (Analog Devices,

Norwood, MA) to alternate between the FSCV and EIS waveform channels. The combined waveform output was applied with an application frequency of 10 Hz and analog filtered using a

2-pole Sallen-Key low-pass filter at 50 kHz; highest cut-off frequency available. No additional explicit filtering from this setup was employed.

8.2.2 Animal Experiments

Male Sprague-Dawley rats (250-300 g, Charles River Laboratories, Raleigh, NC) were used in all animal experiments. Animal care and use was in accordance with North Carolina State

University Institutional Animal Care and Use Committee (IACUC) guidelines. One animal was used as an acute electrode preparation. The animal was isoflurane anesthetized (4% induction, maintained at 1.5%), placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes were drilled in the skull using stereotaxic coordinates.40 The working electrode was placed in the striatum (1.2 mm AP, 3.0 mm ML, -5.0 mm DV relative to bregma), and the

Ag/AgCl reference electrode was placed in the contralateral forebrain. The animal’s body temperature was maintained at 37 °C by a heating pad.

210

Four animals were used in adrenal gland slice experiments. One animal was treated with reserpine (5 mg/kg) and α-methyl-DL-tyrosine methyl ester hydrochloride (250 mg/kg) on the day of, and two days prior to decapitation. Animals were isoflurane anesthetized (4% induction), decapitated and adrenal glands were quickly removed, trimmed of fat tissue, and embedded in 3% agarose gel in bicarbonate-buffered saline (BBS). The gel blocks were placed in cold BBS and 400

µm thick slices were cut with a vibratome (World Precision Instruments, Sarasota, FL). The slices were allowed to rest in oxygenated BBS for one hour, then transferred to a recording chamber

(Warner instruments, Hamden, CT) and superfused with oxygenated BBS maintained at 34ºC for at least 30 min prior to placing electrodes. Carbon-fiber microelectrodes were positioned in tissue near twin tungsten stimulating electrodes. Electrode placements were made with the aid of a microscope (Nikon Instruments, Inc., Melville, NY) and micromanipulators (Scientifica, ltd.,

United Kingdom). Electrical stimulations consisted of trains of 16 biphasic, 500 µA pulses at a frequency of 60 Hz with a pulse width of 2 ms.

One animal was used as a chronic electrode preparation. Surgery was performed as previously described. Briefly, the animal was isoflurane anesthetized (4% induction maintained at

1.5%), placed in a stereotaxic frame (David Kopf Instruments, Tujunga, CA), and holes for electrodes and screws were drilled in the skull using stereotaxic coordinates. The working electrode was placed in the nucleus accumbens (+1.2 mm AP, +1.4 mm ML, -7.0 mm DV relative to bregma), the stimulation electrode in the ventral tegmental area (-5.2 mm AP, +1.0 mm ML, -

8.2 mm DV relative to bregma) and the Ag/AgCl reference electrode was placed in the contralateral forebrain. The components were permanently affixed with dental cement. The animal’s body temperature was maintained at 37 °C by a heating pad. The animal was allowed to recover for two weeks prior to recordings and electrical stimulation. Electrical stimulations

211

consisted of trains of 12 biphasic, 175 μA pulses at a frequency of 60 Hz with a pulse width of 2 ms.

8.2.3 Data Processing and Statistics

Pine Aftermath software was used to processes and analyze data recorded from the

WaveDriver200. All data recorded from FSCV hardware were exported using HDCV and imported into MATLAB R2018a (Mathworks, Natick, MA). Briefly, if necessary single and multi- frequency impedance current were separated from FSCV current, and traces were plotted versus time and fit to a sum of sines model; correlation between the raw data and the fit were R2>0.95.

Coefficients from the fit were used for determination of impedance, phase angle, and reactance.

Capacitance was determined by fitting reactance data over multiplex frequencies using nonlinear regression. A more detailed description of data processing is described in the Supplementary

Information.

Data are presented as the mean ± standard error, unless otherwise noted. Repeated measures one-way ANOVA was used to compare significance between multiple groups.

Bonferroni’s post-hoc test determined significance between specific groups for data comparing electrode sensitivity and impedance between electrochemical conditionings. Dunnett’s post-hoc test determined significance of all impedance time-point measurements to either the initial of final value, and for comparisons of multiplex EIS phase angles to those recorded on a commercial instrument. Significance between two groups was determined via a two-tailed t-test. Significance was determined at p<0.05. Statistical analyses and graphical depictions of the data were carried out using GraphPad Prism 6 (GraphPad Software, Inc., La Jolla, CA) and MATLAB.

212

8.3 Results and Discussion

8.3.1 Surface State/Microenvironment Impact FSCV Performance

Background-subtracted fast-scan cyclic voltammetry (FSCV) is commonly utilized to detect chemical dynamics of catecholamine species such as dopamine (DA) and norepinephrine with a triangular waveform ramping from a -0.4 V holding potential to +1.3 V and back, using scan rates >300 V s-1 and a waveform application frequency of 10 Hz. Figure 1A displays a typical waveform which when applied to a carbon fiber microelectrode (CFME) generates a large, predominately nonfaradaic background cyclic voltammogram (CV). The background can be subtracted to reveal faradaic contributions from fluctuating analytes as shown in Figure 1B for a

DA signal. Signal amplitude and position are utilized for quantification and identification of an analyte(s). Advances in sensor design and waveform development have expanded FSCV detection to a range of species vital for nervous system function such as serotonin, histamine, melatonin, adenosine, glucose, lactate, hydrogen peroxide, and peptides.41–47

Increasing adoption of this technique within neuroscience and behavioral psychology have led to its use for long-term measurements (weeks to months) made in freely behaving animals in a variety of tissues. It has been demonstrated that tissue exposure influences electrochemical performance.22,24,31,35 To demonstrate this effect, a flow-injection system was used to introduce

DA to CFMEs: (1) before implantation, (2) immediately following implantation in the striatum of an anesthetized rat for ~30 min, and (3) following post-implantation cleaning in isopropyl alcohol

(IPA) overnight. Representative background and DA CVs are displayed in Figure 1C. Following implantation, shifts in both CVs are observed, namely the oxidation potential for DA shifts to a more positive potential and the signal amplitude decreases for the same concentration of DA. This is attributed to fouling of the electrode surface by adsorbed biological material. Following cleaning

213

with IPA, the signal more closely resembles that of the pre-implant DA response. There is also a difference in observed catecholamine (CA) response based on the tissue type, as well as the chemical composition of the recording microenvironment; CVs for the catecholamines DA, norepinephrine, and epinephrine are similar with the waveform used.48 To display this effect, electrically stimulated CA was monitored in different tissue preparations: (1) chronically implanted CFME in the nucleus accumbens of a freely-moving rat, and acutely implanted CFME in the medulla of an adrenal gland slice (2) treated with reserpine and α-MPT to deplete basal CA levels and (3) an untreated animal.49,50 The nucleus accumbens contains a relatively high density of dopaminergic neuron terminals that release DA upon stimulation of the cell bodies in the ventral tegmental area. The adrenal medulla contains chromaffin cells that are electrically excitable cells that excrete primarily norepinephrine and epinephrine. Basal CA levels in the adrenal medulla are clearly observable in the background CV (not shown); background response in glands from the

CA-depleted animals do not display as pronounced a CA signature. Figure 1D displays normalized

CVs recorded in the different preparations. It is evident that the shape of the CA response is significantly impacted by both the locale and the chemical composition of the recording environment, with those recorded in the untreated (high basal catecholamine) adrenal medulla displaying the most distortion from CVs recorded in vitro. Previous work indicated that these observed distortions of CA responses could be mimicked using RC-circuits incorporated in vitro to modulate electrochemical system impedance.35 Figure 1E displays the change in FSCV response to both DA and hydrogen peroxide resulting from increased impedance. Both signals change significantly, and the DA response following addition of impedance resembles the distortion observed following tissue exposure and in the adrenal medulla. If not accounted these shifts can result in significant misestimation of analyte(s) concentrations.31 Thus, a method to monitor

214

impedance properties in situ and in real-time with FSCV measurements would supply researchers feedback on changing electrochemical performance.

Figure 8.1. Fast-Scan Cyclic Voltammetry and Surface State. A) Conventional FSCV waveform. B) Current arising from waveform application with (blue; left axis) and without (black; left axis) electroactive species present, and the resulting background-subtracted CV of DA (red; right axis). C) In vitro background (left) and 250 nM DA CVs (right): prior to (black) and following implantation in the striatum of an anesthetized rat (red), as well as after post-implant cleaning in isopropyl alcohol (IPA; blue). D) Normalized, electrically stimulated catecholamine CVs recorded from a chronically implanted electrode in the rat striatum (black), and from an acutely implanted electrode in an adrenal slice of a native (red) and catecholamine-depleted rat (blue). E) In vitro 1 µM DA (black) and 200 µM H2O2 (red) with (dashed) and without impedance added in series with the CFME.

8.3.2 Performing Impedance Measurements Using FSCV Specific Hardware

EIS measurements are made using low amplitude excitations to minimize perturbation to an electrochemical system and to transform relevant mathematical calculations to linear forms.38

Commercially available EIS systems are designed to minimize intrinsic filtering from hardware

215

components to enable accurate frequency measurements over many orders of magnitude (mHz-

MHz). FSCV instrumentation is designed for use with ultramicro- and nanolectrodes, in a two- electrode setup due to the low currents generated. Furthermore, the majority of FSCV instrumentation is custom-made.

To assess the viability of FSCV instrumentation for EIS measurements, a series of RC- circuits of known impedances were used to assess performance over a range of frequencies on multiple FSCV instruments; a commercially available and a custom-built instrument. A series of

25 mV amplitude sinusoidal waveforms (DC=0.0 V) were generated from 1 Hz-10 kHz and each applied for 1 s with a 1 Hz waveform application frequency to each of the RC-circuits. Figure 2A displays representative waveform and current versus time traces for a 100 Hz measurement. These data were fit to a sum of sines (R2>0.95; 1 s fit) and resulting amplitude coefficients were used to calculate impedance based on Ohm’s law. Figure 2B displays a plot of added vs calculated impedance for the 100 Hz data; a slope close to unity (linear regression) indicates an acceptable measurement of impedance. Figure 2C displays the slope of the added vs calculated impedances

(all R2=0.99), as in Figure 2B, from 1 Hz-10 kHz. It is evident that at frequencies >1 kHz, significant systematic error and misestimation of the circuit impedance is introduced. Thus, only frequencies ≤1 kHz are used in further investigations. Next, a comparison of EIS data recorded at a CFME was made. Figure 2D displays a representative Bode (left) and Nyquist (right) plots for a typical CFME. The two peaks in the phase Bode plot result from double-layer and shunt capacitance; shunt capacitance refers to induced double-layer across the insulation of the electrode.51 Figure 2E displays a comparison of 100 Hz Lissajous curves for data recorded on commercially available EIS (black) and FSCV (red) systems, which are in good agreement. In

Figure 2, only measurements recorded on the commercially available FSCV system, without

216

explicit analog filtering on the applied waveform and current response, are displayed. On the custom-built instrument, an analog low-pass filter on the applied waveform could be selected between 1-50 kHz; no explicit current filter. By minimizing filtering (50 kHz), results are consistent with those made on the commercial instrument; increased filtering led to increased error that is more pronounced in higher frequency EIS measurements. Typical FSCV waveform frequencies are less than 1000 Hz; 8.5 ms total waveform time (Figure 1A) is approximately 118

Hz. Thus, if careful consideration to instrument settings and capabilities are made, informative measurements of impedance properties can be made using FSCV instrumentation at frequencies relevant to FSCV waveforms.

Figure 8.2. Validation of impedance spectroscopic measurements on FSCV hardware. A) 25 mV, 100 Hz sinusoidal potential waveform (left-axis, black) and resulting current (right-axis, red) when applied to an RC-circuit. B) 100 Hz impedance values calculated from current responses for a set of RC-circuits (n=29 circuits), compared to known circuit impedances. C) Slope of known versus calculated impedances (as in B) from 1 Hz-10 kHz. D) Representative Bode (left) plot for impedance (red; left axis) and phase (blue; right axis), and Nyquist (right) plots for a carbon-fiber microelectrode; 25 mV amplitude, 1 Hz-1 MHz. Inset is an expanded view of the Nyquist plot from 1 MHz-100 Hz. E) 100 Hz Lissajous curves for EIS data recorded on commercial EIS hardware (black) and commercial FSCV hardware (red) for a CFME.

217

8.3.3 CFME Impedance Properties are Paired with FSCV Performance

CFMEs can be subjected to pretreatments that enhance sensitivity to analytes.

Conventionally, this is achieved through electrochemical conditioning via application of waveforms with potentials exceeding 1.0 V, which have been demonstrated to result in surface etching, alter carbon microstructure, and introduce beneficial surface functionalities.15,17,52,53 This pretreatement has been demonstrated to increase the magnitude of the FSCV background

(charging) current suggesting an increase in capacitance and surface area, presumably due to etching.15,22 To demonstrate the utility of impedance measurements to inform about CFME performance, impedance measurements on FSCV instrumentation will be made, as described in the previous section, following sequentially more oxidative electrochemical conditioning. This was done by rapidly applying a FSCV waveform (60 Hz application frequency), similar to that in

Figure 1A, with positive limits increased from +1.0 to +1.2 V, and then to +1.4 V for at least 30 min before each set of FSCV and EIS measurements. Figure 3A displays three-point DA calibrations following each conditioning; using the conditioning waveform with a waveform application frequency of 10 Hz. As expected, a significant increase in sensitivity (slope;

F(2,5)=83.41, p<0.0001) to dopamine is observed with more anodic conditioning; one-way

ANOVA with Bonferroni post-hoc. Figure 3B (left) displays the impedance Bode plots for 10,

100, and 1000 Hz measurements following each conditioning step. Figure 3B displays impedance and capacitive reactance Bode plots, as well as comparisons of 100 Hz impedance and capacitance for recordings following electrochemical conditioning; capacitance determined by fitting the capacitive reactance data to the inset equation. Anodic conditioning of the CFME reduces impedance (F(2,7)=42.53, p<0.0001 and increases capacitance (F(2,7)=317.8, p<0.0001); one- way ANOVA with Bonferroni post-hoc.

218

Figure 8.3. EIS measurements inform about FSCV performance. A) DA calibrations (n=6 electrodes) and B) Impedance (left) and capacitive reactance (left-middle) Bode plots (n=8 electrodes) determined following progressive electrochemical conditioning using more positive waveform limits. 100 Hz impedance (right-middle) and capacitance determined from the reactance data fit (right) are compared to determine significance. C) 100 Hz impedance (left), 100 Hz reactance (middle), capacitance (right) and, D) electrode drift as a function of electrochemical conditioning time. Measurements (n=3 electrodes) were made every 5 min. with application of FSCV waveform between recordings. E) 100 Hz impedance (left), 100 Hz reactance (middle) and capacitance (right) recorded before, during, and after implantation in an adrenal tissue slice. Measurements (n=5 electrodes) were made every 5 min. with or without the application of an FSCV waveform between recordings.

In a similar experiment, pristine CFMEs were electrochemically conditioned using a triangular waveform with a +1.4 V limit and a waveform application frequency of 10 Hz. Every 5 minutes impedance measurements at 10, 100, and 1000 Hz were obtained to investigate the time- course of interfacial changes resulting from conditioning. An FSCV recording obtained during the

219

last 30 s of each 5 minute interval was used to assess electrochemical drift, or the extent to which the background signal changes over a duration of time.39,54 Figure 3C displays results for changing impedance and capacitive reactance at 100 Hz, as well as capacitance determined from fitting reactance data, over 60 min of electrochemical conditioning. The 500 and 1000 Hz measurements display similar results only with lower impedance and reactance (not shown). Figure 3D displays

FSCV electrode drift over the 60 min. Electrode drift was determined by performing background- subtraction using the average of the first 10 background CVs (1 s) for each 30 s recording, and taking the sum of the absolute value of all current. Electrochemical conditioning takes minutes to complete, resulting in continual changes to electrode impedance (F(12,2)=323.5, p<0.0001), capacitive reactance (F(12,2)=308.1, p<0.0001), capacitance (F(12,2)=521.7, p<0.0001), and

FSCV electrode drift (F(12,2)=22.91, p<0.0001) that occur over time; one-way ANOVA with a

Dunnett’s post-hoc comparing all time points to the final measurement. From experiments described in Figures 3A-C, it is demonstrated that electrochemical conditioning significantly decreases impedance and increases capacitance over time, suggesting an increase in surface area that results in improved sensitivity. These experiments also lend evidence that changes at the carbon-electrolyte interface lead to differential impedance and capacitance that are significant contributors to electrode drift.

As described in an above section, once implanted in tissue, the CV for catecholamine can undergo distortions and sensitivity can decrease relative to measurements made in vitro.35 This effect can be mimicked using RC-circuits to modulate impedance in vitro. To explore changes in

CFME impedance that occur in tissue, measurements were made prior to, during, and after implantation into the medulla of an adrenal tissue slice, in ~5 min intervals; reference electrode remains in BBS buffer. Figure 3E displays the 100 Hz impedance and capacitive reactance, as well

220

as capacitance monitored over time. The first data point (black circle) is the initial impedance in the BBS buffer following waveform cycling. The electrode was then implanted, and an impedance measurement made immediately (blue). Next, the electrode was then allowed to rest in tissue without (red) and with (orange) FSCV waveform cycling with a 10 Hz application frequency between measurements. The FSCV waveform was then turned-off, the electrode moved back into the buffer (green), and finally electrochemically conditioned prior to the final measurement in the

BBS buffer (black star). It should be noted that capacitance depends on ionic strength; a higher ionic strength will decrease diffuse layer thickness and increase capacitance.55 The extracellular space likely has a higher ionic strength than the BBS buffer, due to the presence of charged biological molecules, which would result in elevated capacitance relative to measurements made in BBS. A one-way ANOVA with a Dunnett’s post-hoc was used to compare all time course data to that of the initial measurement in BBS buffer. From these data, it is evident that a significant increase in impedance (F(17,4)=28.74, p<0.0001) and capacitive reactance (F(17,4)=28.01, p<0.0001), as well as a decrease in capacitance (F(17,4)=36.61, p<0.0001) is observed following implantation. Furthermore, if left stagnant in tissue these values will continue to deviate from initial buffer measurements. This is attributed to increasing adsorption of biological material to the electrode surface. However, following sufficient conditioning the continual deviation from the pre- implant impedance and capacitance can be mitigated. Once removed from the tissue, the electrode remains fouled until sufficient waveform cycling has occurred, at which point the impedance and capacitance are no longer significantly different than the pre-implant values. These data suggest that significant adsorption of material continually occurs and that the active electrode surface area following implantation decreases.

221

8.3.4 Combined Impedimetric and Voltammetric Measurement with High Temporal Resolution

A significant advantage of FSCV lies in the ability to make in situ measurements with spatial resolution of individual brain nuclei dimensions and the time scale of neurochemical fluctuations. Up to this point, impedance measurements have been made by switching from an 8.5 ms FSCV waveform applied at 10 Hz, to a 1 s EIS waveform applied at 1 Hz. The need to switch between waveforms during a freely behaving or anesthetized animal experiment is cumbersome, as time is at a premium and many neurochemical events can occur seemingly at random. To circumvent the need to switch between waveforms, we aimed to construct a multiplexed measurement incorporating an impedimetric measurement in the time between FSCV waveform applications; in this case 91.5 ms. To accomplish this task, a multi-channel FSCV system (capable of individually addressing multiple electrodes) was utilized in conjunction with a logic controlled analog switch to create a time-division multiplexed waveform output delivered to an electrode as shown in Figure 4A for a 100 Hz impedance measurement. Through this, a combined FSCV:EIS waveform can be delivered to an electrode. An example multiplexed FSCV:EIS100Hz waveform and the resulting current response at a CFME are displayed in Figure 4B. Here, the analog switch

(~150-200 ns) changes to waveform channel two between 20-90 ms, resulting in a 70 ms impedance measurement. The EIS segment is set with a -0.4 V offset; a negative potential between waveform applications is important for preconcentrating CA to improve sensitivity. These data can then be exported, and measurement modalities separated for analysis. The EIS segment can then be fit to a sum of sines, and coefficients used in further calculations. Herein, this paradigm is used to perform impedance measurements at 100, 500, and 1000 Hz. Additionally, the use of a frequency-division multiplex to create a combined 100, 500, and 1000 Hz EIS segment is explored.

222

Figure 8.4. Time-division multiplexing of FSCV and EIS measurements. A) FSCV (top) and 100 Hz EIS (bottom) waveforms applied to separate waveform outputs. The final output to the electrode was controlled via a CMOS analog switch by application of ±5 V to select between the two channels. B) Multiplexed FSCV-EIS100Hz waveform output (top) and resulting current response (bottom). C) Comparison (n=15 electrodes) of 100 Hz impedance, phase angle, and capacitive reactance determined using a commercial EIS instrument (black) and using the multiplexed FSCV-EIS100Hz paradigm (red). Percent relative standard deviation (%RSD) and correlation between the two measurements (Rpairing) are displayed.

To validate measurements made using the multiplexed paradigm, a comparison of

FSCV:EIS100 Hz measurements made on a set of electrodes are compared to those made using the commercial EIS instrument (DC=0 V), and results are displayed in Figure 4C. Using the

FSCV:EIS100Hz waveform, impedance was underestimated (two-tailed t-test, t(14)=23.76, p<0.0001), phase angle was less negative (two-tailed t-test, t(14)=37.97, p<0.0001), and capacitive reactance underestimated (two-tailed t-test, t(14)=15.18, p<0.0001) relative to the commercial EIS instrument. However, the pairing between the measurements on the two instruments/paradigms was excellent (Rpairing≥0.98), suggesting that the discrepancy is not random. Furthermore, the

223

relative standard deviations between the two instruments/paradigms are also in agreement, suggesting that the precision of the measurements are equal. Thus, the FSCV:EIS paradigm should serve to provide valuable information on changing electrical properties at a temporal resolution conventionally used in FSCV experiments. Also important to note, the relative standard deviation indicates that CFME impedance and capacitive reactance (proportional to 1/capacitance) is more variable than phase angle; i.e. different sets of electrodes could have significantly different impedances and capacitances, but should have similar phase angles.

Next, in a similar experiment similar to that shown in Figure 3E, the FSCV:EIS waveform was utilized at a CFME implanted in the medulla of a rat adrenal slice. Figure 5 displays impedance and capacitive reactance recorded in tissue when the FSCV waveform was cycled (black) for 15 min, and then when the FSCV waveform was replaced by a DC hold at -0.4 V (red). As observed in Figure 3E, when the FSCV waveform is cycled the impedance and reactance remain relatively constant. Once the electrode was left stagnant, impedance and reactance rapidly increase suggesting a decrease in capacitance. These measurements demonstrate the utility of impedimetric measurements made in real-time with FSCV, as they serve to inform the user regarding the state of the electrode and the stability of an electrode to perform measurements. For example, a permanently implanted electrode may be left stagnant in the brain for days to weeks at a time, the impedance feedback would inform a user about the duration of waveform cycling needed to mitigate effects of fouling. In addition, tracking sensor impedance and capacitance time locked with FSCV during a long-term experiment could inform a user about potential significant changes in sensor performance.

224

Figure 8.5. FSCV:EIS100Hz measurements in the medulla of adrenal tissue slices (n=4 electrodes). Impedance (left) and capacitive reactance (right) measurements recorded (10 Hz waveform application frequency) over time during implantation, with (black) and without (red) the application of a FSCV waveform. Applied waveforms for each case are displayed above.

To gain additional impedance information within a single waveform and permit real-time determination of capacitance (from a fit to reactance data), a frequency-division multiplex incorporating 100, 500, and 1000 Hz measurements in the EIS segment of the FSCV:EIS waveform was investigated and compared to measurements made with individual frequencies.

Figure 6A displays the multi-frequency FSCV:EIS100,500,1000 Hz waveform (top) as well a closer view of the potential waveform (middle) and resulting current (bottom) of the EIS segment at a

CFME. The EIS segments can be fit to a sum of sines (R2>0.9) and coefficients used in further calculations. Figure 6B displays the phase angles determined using both 100, 500, and 1000 Hz individual frequency and multi-frequency FSCV:EIS waveforms for a set of electrodes (n=3) were compared to a set (n=15) recorded on the commercial EIS instrument, as the relative standard deviation in the phase angle of CFMEs is less than that of impedance and capacitance. A comparison between the data were made using a one-way ANOVA with a Dunnett’s post-hoc comparing the EIS:FSCV100,500,1000 Hz to the commercial EIS measurements. Figure 6C displays comparisons of impedance (top) and reactance/capacitance (bottom) measured using the individual

225

frequency and multi-frequency FSCV:EIS waveforms. From these data, it is evident that the phase angle measured using FSCV instrumentation is again slightly mismatched to those recorded on the commercial EIS instrument; F(6,26)=7.655100 Hz, 87.01500 Hz, 517.21000 Hz, p<0.0001). However, the data are consistent with impedance data recorded on the commercial instrument; i.e. the phase angle becomes less negative with increased frequency, and the impedance and reactance both decrease with an increase in frequency, as expected for a system dominated by capacitance.

Furthermore, the impedances, phase angles, reactances, and capacitances measured with the single- and multi-frequency FSCV:EIS waveforms are in agreement (Figure 6 and Supplemental

Figure 2). Therefore, multi-frequency FSCV-EIS is capable in making precise measurements comparable to those recorded using a single frequency waveform and the commercial EIS instrument; albeit with a slight deviation in accuracy with respect to the commercial EIS instrument.

8.3.5 Critical Considerations Using the Multiplexed Paradigm

During a typical FSCV experiment, high sampling rates (typically 100 kHz) are used to record only the current attributed to the 8.5 ms waveform. This leads to significant data density when making measurements over minutes to hours. Using this FSCV:EIS paradigm, the entirety of the 100 ms is recorded. With this in mind, as well as the development of wireless based measurement systems, the sampling rate for these measurements was reduced to either 10 kHz for

100 Hz impedance measurements, and 25 kHz for 500 and 1000 Hz impedance measurements to reduce data density.56–58

In all recordings using the FSCV:EIS paradigm, a delay or lag was introduced into the current response, as shown in Supplemental Figure 1, corresponding to ~100-120 µs. This delay was not observed in the nonmultiplexed, 1s EIS waveform measurements. This current delay does

226

not impact impedance calculations across all measurements, minimally impacts phase and reactance measurements made at 100 Hz, but introduces significant error into phase and reactance measurements made at 500 and 1000 Hz, leading to misestimation of capacitance (Supplemental

Figure 2). Thus, all FSCV:EIS calculations shown herein have had the EIS segment corrected for the delay. A description of how this delay was corrected, and its impact on measurements, is described in the Supplemental Information.

Figure 8.6. 100, 500, and 1000 Hz frequency-division multiplexing of EIS portion of FSCV:EIS waveform. A) Combined, multi-frequency FSCV:EIS100,500,1000 Hz waveform (top), expanded EIS portion of the waveform (middle), and resulting EIS current (bottom). B) Comparison of phase angles calculated (n=3 electrodes) using single-frequency (solid) and multi-frequency (checkered) FSCV:EIS from compared to those recorded for a set of electrodes (n=15) using a commercial EIS instrument (black striped). C) Comparison of impedance (top), capacitive reactance (bottom) and capacitance (inset) calculated (n=3 electrodes) using single-frequency (solid) and multi-frequency (checkered) FSCV:EIS.

227

8.4 Conclusions

As demonstrated herein, FSCV performance can be impacted by the physical and chemical composition of the recording microenvironment, as well as the surface state of the CFME. We developed a method to multiplex voltammetric and impedimetric techniques within the same waveform application, for precise measurements using FSCV instrumentation. This method allowed tracking of CFME impedance at multiple frequencies without a loss in temporal resolution commonly utilized in FSCV experiments. These measurements were used to investigate changes to the CFME surface via impedance and capacitance. Electrochemical pretreatment that increases sensitivity to a variety of analytes, decreases and increases impedance and capacitance, respectively; supporting the claim of surface etching. Also, the magnitude of electrode drift (slowly occurring change in background signal) is directly related to the rate of impedance or capacitance change of an electrode. Furthermore, measurements reported herein represent the first measurements of dynamic impedance properties during implantation in live tissue. It was observed that upon implantation a significant increase and decrease in impedance and capacitance occurs rapidly, that continues to change if the electrode rests unperturbed. These measurements may explain the changes to FSCV performance observed at implanted electrodes. Importantly, this

FSCV:EIS paradigm will inform users regarding observed changes in performance and provide feedback regarding electrochemical conditioning required between measurements (days to weeks) at permanently implanted CFMEs. Additionally, tracking impedance properties time locked with

FSCV measurements will be valuable in developing in situ calibration strategies.

228

8.5 References

1. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, 2006.

2. Ou, Y. G.; Buchanan, A. M.; Witt, C. E.; Hashemi, P. Frontiers in Electrochemical Sensors for Neurotransmitter Detection: Towards Measuring Neurotransmitters as Chemical Diagnostics for Brain Disorders. Anal. Methods 2019, 11 (21), 2738–2755.

3. Roberts, J. G.; Lugo-Morales, L. Z.; Loziuk, P. L.; Sombers, L. A. Real-Time Chemical Measurements of Dopamine Release in the Brain. In Dopamine; Springer, 2013; pp 275–294.

4. Howe, M. W.; Tierney, P. L.; Sandberg, S. G.; Phillips, P. E.; Graybiel, A. M. Prolonged Dopamine Signalling in Striatum Signals Proximity and Value of Distant Rewards. Nature 2013, 500 (7464), 575–579.

5. Phillips, P. E.; Stuber, G. D.; Heien, M. L.; Wightman, R. M.; Carelli, R. M. Subsecond Dopamine Release Promotes Cocaine Seeking. Nature 2003, 422 (6932), 614–618.

6. Day, J. J.; Roitman, M. F.; Wightman, R. M.; Carelli, R. M. Associative Learning Mediates Dynamic Shifts in Dopamine Signaling in the Nucleus Accumbens. Nat Neurosci 2007, 10 (8), 1020–1028.

7. Gan, J. O.; Walton, M. E.; Phillips, P. E. Dissociable Cost and Benefit Encoding of Future Rewards by Mesolimbic Dopamine. Nat Neurosci 2010, 13 (1), 25–27.

8. Roitman, M. F.; Stuber, G. D.; Phillips, P. E. M.; Wightman, R. M.; Carelli, R. M. Dopamine Operates as a Subsecond Modulator of Food Seeking. J Neurosci 2004, 24 (6), 1265–1271.

9. Clark, J. J.; Sandberg, S. G.; Wanat, M. J.; Gan, J. O.; Horne, E. A.; Hart, A. S.; Akers, C. A.; Parker, J. G.; Willuhn, I.; Martinez, V.; et al. Chronic Microsensors for Longitudinal, Subsecond Dopamine Detection in Behaving Animals. Nat Methods 2010, 7 (2), 126–129.

10. Covey, D. P.; Dantrassy, H. M.; Zlebnik, N. E.; Gildish, I.; Cheer, J. F. Compromised Dopaminergic Encoding of Reward Accompanying Suppressed Willingness to Overcome High Effort Costs Is a Prominent Prodromal Characteristic of the Q175 Mouse Model of Huntington’s Disease. J Neurosci 2016, 36 (18), 4993–5002.

11. Perez, X. A.; Khroyan, T. V.; McIntosh, J. M.; Quik, M. Varenicline Enhances Dopamine Release Facilitation More than Nicotine after Long-Term Nicotine Treatment and Withdrawal. Pharmacol Res Perspect 2015, 3 (1), e00105.

12. Schwerdt, H. N.; Shimazu, H.; Amemori, K. I.; Amemori, S.; Tierney, P. L.; Gibson, D. J.; Hong, S.; Yoshida, T.; Langer, R.; Cima, M. J.; et al. Long-Term Dopamine Neurochemical Monitoring in Primates. Proc Natl Acad Sci U A 2017, 114 (50), 13260–13265.

13. Williams, J. C.; Rennaker, R. L.; Kipke, D. R. Long-Term Neural Recording Characteristics of Wire Microelectrode Arrays Implanted in Cerebral Cortex. Brain Res Brain Res Protoc 1999, 4 (3), 303–313.

229

14. Hafizi, S.; Kruk, Z. L.; Stamford, J. A. Fast Cyclic Voltammetry: Improved Sensitivity to Dopamine with Extended Oxidation Scan Limits. J Neurosci Methods 1990, 33 (1), 41–49.

15. Heien, M. L.; Phillips, P. E.; Stuber, G. D.; Seipel, A. T.; Wightman, R. M. Overoxidation of Carbon-Fiber Microelectrodes Enhances Dopamine Adsorption and Increases Sensitivity. Analyst 2003, 128 (12), 1413–1419.

16. McCreery, R. L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem Rev 2008, 108 (7), 2646–2687.

17. Roberts, J. G.; Moody, B. P.; McCarty, G. S.; Sombers, L. A. Specific Oxygen-Containing Functional Groups on the Carbon Surface Underlie an Enhanced Sensitivity to Dopamine at Electrochemically Pretreated Carbon Fiber Microelectrodes. Langmuir 2010, 26 (11), 9116– 9122.

18. Oleinick, A.; Alvarez-Martos, I.; Svir, I.; Ferapontova, E. E.; Amatore, C. Surface Heterogeneities Matter in Fast Scan Cyclic Voltammetry Investigations of Catecholamines in Brain with Carbon Microelectrodes of High-Aspect Ratio: Dopamine Oxidation at Conical Carbon Microelectrodes. J. Electrochem. Soc. 2018, 165 (12), G3057–G3065.

19. Hashemi, P.; Dankoski, E. C.; Petrovic, J.; Keithley, R. B.; Wightman, R. M. Voltammetric Detection of 5-Hydroxytryptamine Release in the Rat Brain. Anal Chem 2009, 81 (22), 9462– 9471.

20. Harreither, W.; Trouillon, R.; Poulin, P.; Neri, W.; Ewing, A. G.; Safina, G. Carbon Nanotube Fiber Microelectrodes Show a Higher Resistance to Dopamine Fouling. Anal Chem 2013, 85 (15), 7447–7453.

21. Williams, J. C.; Hippensteel, J. A.; Dilgen, J.; Shain, W.; Kipke, D. R. Complex Impedance Spectroscopy for Monitoring Tissue Responses to Inserted Neural Implants. J Neural Eng 2007, 4 (4), 410–423.

22. Roberts, J. G.; Toups, J. V.; Eyualem, E.; McCarty, G. S.; Sombers, L. A. In Situ Electrode Calibration Strategy for Voltammetric Measurements in Vivo. Anal Chem 2013, 85 (23), 11568–11575.

23. Peltola, E.; Sainio, S.; Holt, K. B.; Palomaki, T.; Koskinen, J.; Laurila, T. Electrochemical Fouling of Dopamine and Recovery of Carbon Electrodes. Anal Chem 2018, 90 (2), 1408– 1416.

24. Logman, M. J.; Budygin, E. A.; Gainetdinov, R. R.; Wightman, R. M. Quantitation of in Vivo Measurements with Carbon Fiber Microelectrodes. J Neurosci Methods 2000, 95 (2), 95–102.

25. Rodeberg, N. T.; Sandberg, S. G.; Johnson, J. A.; Phillips, P. E.; Wightman, R. M. Hitchhiker’s Guide to Voltammetry: Acute and Chronic Electrodes for in Vivo Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 2017, 8 (2), 221–234.

230

26. Meunier, C. J.; Mitchell, E. C.; Roberts, J. G.; Toups, J. V.; McCarty, G. S.; Sombers, L. A. Electrochemical Selectivity Achieved Using a Double Voltammetric Waveform and Partial Least Squares Regression: Differentiating Endogenous Hydrogen Peroxide Fluctuations from Shifts in PH. Anal Chem 2018, 90 (3), 1767–1776.

27. Rodeberg, N. T.; Johnson, J. A.; Cameron, C. M.; Saddoris, M. P.; Carelli, R. M.; Wightman, R. M. Construction of Training Sets for Valid Calibration of in Vivo Cyclic Voltammetric Data by Principal Component Analysis. Anal Chem 2015, 87 (22), 11484–11491.

28. Keithley, R. B.; Heien, M. L.; Wightman, R. M. Multivariate Concentration Determination Using Principal Component Regression with Residual Analysis. Trac-Trends Anal. Chem. 2009, 28 (9), 1127–1136.

29. Johnson, J. A.; Gray, J. H.; Rodeberg, N. T.; Wightman, R. M. Multivariate Curve Resolution for Signal Isolation from Fast-Scan Cyclic Voltammetric Data. Anal Chem 2017, 89 (19), 10547–10555.

30. Kishida, K. T.; Saez, I.; Lohrenz, T.; Witcher, M. R.; Laxton, A. W.; Tatter, S. B.; White, J. P.; Ellis, T. L.; Phillips, P. E. M.; Montague, P. R. Subsecond Dopamine Fluctuations in Human Striatum Encode Superposed Error Signals about Actual and Counterfactual Reward. Proc. Natl. Acad. Sci. U. S. A. 2016, 113 (1), 200–205.

31. Johnson, J. A.; Rodeberg, N. T.; Wightman, R. M. Failure of Standard Training Sets in the Analysis of Fast-Scan Cyclic Voltammetry Data. ACS Chem Neurosci 2016, 7 (3), 349–359.

32. Moussy, F.; Harrison, D. J. Prevention of the Rapid Degradation of Subcutaneously Implanted Ag/Agcl Reference Electrodes Using Polymer-Coatings. Anal Chem 1994, 66 (5), 674–679.

33. Kozai, T. D. Y.; Langhals, N. B.; Patel, P. R.; Deng, X.; Zhang, H.; Smith, K. L.; Lahann, J.; Kotov, N. A.; Kipke, D. R. Ultrasmall Implantable Composite Microelectrodes with Bioactive Surfaces for Chronic Neural Interfaces. Nat. Mater. 2012, 11 (12), 1065.

34. Mercanzini, A.; Colin, P.; Bensadoun, J. C.; Bertsch, A.; Renaud, P. In Vivo Electrical Impedance Spectroscopy of Tissue Reaction to Microelectrode Arrays. IEEE Trans Biomed Eng 2009, 56 (7), 1909–1918.

35. Meunier, C. J.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Background Signal as an in Situ Predictor of Dopamine Oxidation Potential: Improving Interpretation of Fast-Scan Cyclic Voltammetry Data. ACS Chem. Neurosci. 2017, 8 (2), 411–419.

36. Chang, B. Y.; Park, S. M. Electrochemical Impedance Spectroscopy. Annu. Rev. Anal. Chem. 2010, 3, 207–229.

37. Lasia, A. Electrochemical Impedance Spectroscopy and Its Applications. In Modern aspects of electrochemistry; Springer, 2002; pp 143–248.

38. Orazem, M. E.; Tribollet, B. Electrochemical Impedance Spectroscopy; John Wiley & Sons, 2017.

231

39. Johnson, J. A.; Hobbs, C. N.; Wightman, R. M. Removal of Differential Capacitive Interferences in Fast-Scan Cyclic Voltammetry. Anal Chem 2017, 89 (11), 6166–6174.

40. Paxinos, G.; Watson, C. The Rat Brain in Stereotaxic Coordinates: Hard Cover Edition; Elsevier, 2006.

41. Calhoun, S. E.; Meunier, C. J.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin. ACS Chem Neurosci 2019, 10 (4), 2022–2032.

42. Hensley, A. L.; Colley, A. R.; Ross, A. E. Real-Time Detection of Melatonin Using Fast-Scan Cyclic Voltammetry. Anal Chem 2018, 90 (14), 8642–8650.

43. Samaranayake, S.; Abdalla, A.; Robke, R.; Wood, K. M.; Zeqja, A.; Hashemi, P. In Vivo Histamine Voltammetry in the Mouse Premammillary Nucleus. Analyst 2015, 140 (11), 3759– 3765.

44. Spanos, M.; Gras-Najjar, J.; Letchworth, J. M.; Sanford, A. L.; Toups, J. V.; Sombers, L. A. Quantitation of Hydrogen Peroxide Fluctuations and Their Modulation of Dopamine Dynamics in the Rat Dorsal Striatum Using Fast-Scan Cyclic Voltammetry. ACS Chem Neurosci 2013, 4 (5), 782–789.

45. Ross, A. E.; Venton, B. J. Sawhorse Waveform Voltammetry for Selective Detection of Adenosine, ATP, and Hydrogen Peroxide. Anal Chem 2014, 86 (15), 7486–7493.

46. Smith, S. K.; Gosrani, S. P.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Carbon-Fiber Microbiosensor for Monitoring Rapid Lactate Fluctuations in Brain Tissue Using Fast-Scan Cyclic Voltammetry. Anal Chem 2018, 90 (21), 12994–12999.

47. Smith, S. K.; Lee, C. A.; Dausch, M. E.; Horman, B. M.; Patisaul, H. B.; McCarty, G. S.; Sombers, L. A. Simultaneous Voltammetric Measurements of Glucose and Dopamine Demonstrate the Coupling of Glucose Availability with Increased Metabolic Demand in the Rat Striatum. ACS Chem Neurosci 2017, 8 (2), 272–280.

48. Pihel, K.; Schroeder, T. J.; Wightman, R. M. Rapid and Selective Cyclic Voltammetric Measurements of Epinephrine and Norepinephrine as a Method to Measure Secretion from Single Bovine Adrenal-Medullary Cells. Anal Chem 1994, 66 (24), 4532–4537.

49. Toneff, T.; Funkelstein, L.; Mosier, C.; Abagyan, A.; Ziegler, M.; Hook, V. Beta-Amyloid Peptides Undergo Regulated Co-Secretion with Neuropeptide and Catecholamine Neurotransmitters. Peptides 2013, 46, 126–135.

50. Martinez-Olivares, R.; Villanueva, I.; Racotta, R.; Pinon, M. Depletion and Recovery of Catecholamines in Several Organs of Rats Treated with Reserpine. Auton. Neurosci.-Basic Clin. 2006, 128 (1–2), 64–69.

51. Wipf, D. O. Ohmic Drop Compensation in Voltammetry: Iterative Correction of the Applied Potential. Anal Chem 1996, 68 (11), 1871–1876.

232

52. Mitchell, E. C.; Dunaway, L. E.; McCarty, G. S.; Sombers, L. A. Spectroelectrochemical Characterization of the Dynamic Carbon-Fiber Surface in Response to Electrochemical Conditioning. Langmuir 2017, 33 (32), 7838–7846.

53. Takmakov, P.; Zachek, M. K.; Keithley, R. B.; Walsh, P. L.; Donley, C.; McCarty, G. S.; Wightman, R. M. Carbon Microelectrodes with a Renewable Surface. Anal Chem 2010, 82 (5), 2020–2028.

54. Meunier, C. J.; McCarty, G. S.; Sombers, L. A. Drift Subtraction for FSCV Using Double- Waveform Partial-Least-Squares Regression. Anal. Chem. 2019.

55. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

56. Lin, N. S.; Wang, L.; Wang, M. X.; Xu, S. W.; Yu, W. D.; Cai, X. X. A Wearable Wireless Electrochemical Instrument Used for In-Vivo Neurotransmitter Detection. Chin. J. Anal. Chem. 2015, 43 (1), 93–97.

57. Zamani, H.; Bahrami, H. R.; Chalwadi, P.; Garris, P. A.; Mohseni, P. C-FSCV: Compressive Fast-Scan Cyclic Voltammetry for Brain Dopamine Recording. IEEE Trans Neural Syst Rehabil Eng 2018, 26 (1), 51–59.

58. Dorta-Quinones, C. I.; Wang, X. Y.; Dokania, R. K.; Gailey, A.; Lindau, M.; Apsel, A. B. A Wireless FSCV Monitoring IC With Analog Background Subtraction and UWB Telemetry. IEEE Trans Biomed Circuits Syst 2016, 10 (2), 289–299.

233

APPENDICES

234

APPENDIX A

Supplemental Information to Chapter 4:

Electrochemical Selectivity Achieved Using a Double Voltammetric Waveform and Partial

Least Squares Regression: Differentiating Endogenous Hydrogen Peroxide Fluctuations from

Shifts in pH

A.1 Summary

This appendix displays relevant statistical information in regards to the PLSR analyses performed in Chapter 4. The first five component loadings (includes principal components) from both the sWF-predictor and the lWF-response are displayed for both in vivo training sets in Figures

A.S-1 and A.S-2. This information qualitatively assists in determining which components may or may not contain contributions from multiple analytes. Furthermore, for brevity and transparency,

Figure A.S-2 also displays the selected in vivo training set containing ∆pH, dopamine, and blank

CVs (not shown), as well as the cross-validation and cumulative percent variance for analyses in

Figure 4.7.

A.2 Table of Contents

• Figure A.S-1: Principal component loadings for the five-components retained for the DW-

PLSR analysis shown in Figures 4.5 and 4.6. Also, displays residual currents for the ∆pH-only

in vivo training CVs following ∆pH-subtraction.

235

• Figure A.S-2: Displays training set CVs, principal component loadings, cross validation, and

cumulative variance data for the in vivo analysis in Figure 4.7.

A.3 Supplemental Figures

Figure A.1. Additional PLSR information for data in Figure 4.6. A) PC loadings for the sWF- predictor (left) and lWF-response (right) plotted versus potential for the first (top) and remaining (bottom) principal components retained in the DWPLSR model B) Residual current not captured by the first five components for the 30 training CVs.

236

Figure A.2. Additional PLSR information for data in Figure 4.7. A) PC loadings for the sWF- predictor (left) and lWF-response (right) plotted versus potential for analyses in Figure 7. B) Cumulative percent variance and cross-validation MSE are plotted on the left and right ordinates, and number of components retained is plotted on the abscissa. 5-fold cross validation was performed by randomly dividing up the training data.

237

APPENDIX B

Supplemental Information to Chapter 5:

Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double-Waveform Partial-Least-

Squares Regression

B.1 Summary

This document displays relevant information with respect to the PLSR analyses performed

in this study. Figures B.S-1, B.S-2, and B.S-3 display training set cyclic voltammograms, as well

as results for k-fold cross validation (used in principal component selection) and cumulative

percent variance (used to assess relative noise). Figure B.S-1 also displays FSCV colorplots

demonstrating effective drift-subtraction from calibration data recorded on the two other electrodes

not displayed in Figure 5.3. Additional experimental methods are also described.

B.2 Table of Contents

• Figure B.S-1: Displays training set CVs, k-fold cross validation, and cumulative percent

variance data for drift-subtraction from the calibration data displayed in Figure 5.3. Also

displays raw, predicted-drift, and drift-subtracted lWF colorplots not included in Figure 5.3.

• Figure B.S-2: Displays training set CVs, k-fold cross validation, and cumulative percent

variance data for drift-subtraction from the in vivo recordings displayed in Figures 5.4 and 5.5.

• Figure B.S-3: Displays a representative training set CVs from each electrode included in the

nonelectrode specific drift-subtraction model shown in Figure 5.6. Also includes k-fold cross

238

validation and cumulative percent variance data, as well as extracted CVs from recordings on

additional electrodes not displayed in Figure 5.6.

• Additional Methods: Description of electrode fabrication and the flow-cell apparatus used for

in vitro investigations

B.3 Supplemental Figures

239

Figure B.1. In vitro drift removal for other calibrations not displayed in Figure 5.3. (A) Drift CVs collected using the sWF (left) and lWF (right) were used to construct training sets for the three electrodes used in Figure 3. The top training set was used for data in Figures 3A-D. (B) Cross- validation mean-squared error (left) and cumulative percent variance (right) for the predictor (black) and response (red) for each of the three training sets. (C) Raw data (top), PLSR-predicted drift (middle), and drift-subtracted color plots for extended calibration recordings of the other two electrodes; colorbar displayed above. The middle (4 components) and bottom (2 components) training sets are paired with the left and right sets of color plots, as well as the top and bottom CVs in Figure 3F, respectively.

240

Figure B.2. Additional PLSR information for data in Figures 5.4 and 5.5.(A) Training set CVs collected using the sWF (left; +0.8 V) and lWF (right), and (B) cumulative percent variance and cross-validation mean-squared error plots for the DW-PLSR model used to analyze the in vivo data shown in Figure 4. (C) Training set CVs collected using the sWF (left; +0.3 V) and lWF (right), and (D) cumulative percent variance and cross-validation mean-squared error plots for the DW- PLSR model used in Figure 5.

241

Figure B.3. Additonal information for data in Figure 5.6. (A) Normalized CVs for ECD collected using the sWF (left) and lWF (right) that are representative of the data used to construct the ten- electrode training set (8 ECD CVs per electrode). Each row contains paired (by color) sWF and lWF responses for electrodes obtained during a flow-cell recording. (B) Cross-validation mean- squared error (left) and cumulative percent variance (right) data for the ten-electrode DW-PLSR model; seven components were retained. (C) CVs extracted at the time of the H2O2 injection (variable concentrations), as in Figure 6C, for three other electrodes with differing drift signals. CVs generated using adjacent background subtraction are also included for comparison (green).

B.4 Additional Methods

B.4.1 Chemicals

All chemicals (≥98%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing; H2O2 was purchased as a

30% solution. In vitro electrochemical experiments were performed in a phosphate buffered saline

242

(10 mM Na2HPO4, 138 mM NaCl, and 2.7 mM KCl). All aqueous solutions were made using ultrapure 18.2 MΩ water.

B.4.2 Microelectrode Fabrication

Cylindrical carbon-fiber microelectrodes were fabricated using T-650 carbon fibers (Cytec

Industries, Inc., Woodland Park, NJ). In short, a single carbon fiber was aspirated into a glass capillary, a glass seal was created via a micropipette puller (Narishige, Tokyo, Japan), and the fiber extending past the seal was cut to 100 μm in length with a scalpel. Electrical connection to the carbon fiber was established via a conductive silver epoxy (GC Electronics, Rockford, IL). A

Ag/AgCl pellet reference electrode (World Precision Instruments, Inc., Sarasota, FL) was used to complete the two-electrode electrochemical cell.

B.4.3 Flow-Injection System

Working electrodes were positioned in a custom electrochemical flow cell using a micromanipulator (World Precision Instruments, Sarasota, FL). A syringe pump (New Era Pump

Systems, Wantagh, NY) supplied a continuous buffer stream (1 mL min-1) across the working and reference electrodes (Ag/AgCl pellet, World Precision Instruments, Inc., Sarasota, FL). Two- second bolus injections of analyte were accomplished with a six-port HPLC valve mounted on a two-position actuator controlled by a digital pneumatic solenoid (Valco Instruments, Houston,

TX). The entire apparatus was enclosed in a custom-built, grounded Faraday cage.

243

APPENDIX C

Supplemental Information to Chapter 8:

Tracking Carbon Microelectrode Impedance Time-Locked with Fast-Scan Cyclic Voltammetry

C.1 Summary

This document contains additional information regarding equations used in impedance calculations from measurements made using FSCV hardware. Namely, it contains details on data processing to account for the delay introduced into the current response when using the multiplexed FSCV:EIS paradigm. It also contains additional methods information. Figure C.S-1 displays the raw current traces of the EIS segment from the FSCV:EIS measurement, which shows the 100-120 µs current delay. Figure C.S-2 shows a comparison of 100, 500, and 1000 Hz impedance, phase angle, and capacitance calculations for single- and multifrequency FSCV:EIS measurements made using the raw traces, as well as with 100 and 120 µs delay compensation.

C.2 Table of Contents

• Relevant Equations

• Additional Information on Data Processing

• Figure C.S-1: Displays a close view of the raw current traces for the EIS segment of the

individual 100, 500, and 1000 Hz, as well as the multifrequency measurements. This displays

the observed time-delay on the current response.

244

• Figure C.S-2: Displays a comparison of calculated phase angles, impedances, and capacitances

performed on the raw and delay-corrected current traces, demonstrating the need to account

for the current-delay.

• Additional Methods

C.3 Relevant Equations

Sum of Sines Fit (For single frequency measurements only one term is used):

푦 = 푎1 ∗ sin(푏1 ∗ 푡 + 푐1) + 푎2 ∗ sin(푏2 ∗ 푡 + 푐2) + 푎3 ∗ sin(푏3 ∗ 푡 + 푐1)

where: a=amplitude, b=angular frequency (rads/s), c=phase (radian), t=time (s)

Impedance and Capacitance (Using Sum of Sines Coefficients)

Potential Amplitude (volts) Impedance (Ohm) = Current Amplitude (amps)

(Current Phase (rad) − Potential Phase(rad) ∗ 180 Phase Angle (deg) = π

Real Impedance (Ohm) = Impedance (Ohm) ∗ cos(Phase Angle)

Imaginary Impedance (Ohm) = Impedance (Ohm) ∗ sin(Phase Angle)

Capacitive Reactance (Ohm) = √Impedance (Ohm)2 − Real Impedance (Ohm)2

1 Capacitive Reactance (Ohm) = rads Angular Frequency ( s ) ∗ Capacitance (farad)

245

C.4 Additional Data Processing

All processing was performed in MATLAB 2018a. For impedance measurements made using the various FSCV:EIS waveforms used, a time-delay was observed in the raw current traces as displayed in Figure C.S-1. With sampling rates of 10 kHz and 25 kHz, this corresponds to ~1 and ~2-3 data points, respectively. To compensate for the time-delay:

1. Raw waveform and current traces were resampled to 50 kHz using the ‘interp’ function, which

resulted in no observable difference from the raw traces; each data point now represents 20 µs.

2. The resampled potential and current were then fit to either a one (‘sin1’) or three component

(‘sin3’) sum of sines; upper and lower bounds for frequencies were supplied based on the

waveform.

3. Coefficients were utilized to:

a. determine impedance based on Ohm’s law.

b. reconstruct potential and current traces for individual frequencies.

4. The phase difference between the reconstructed potential and current was determined using

function obtained through MathWorks File Exchange.1 Briefly, this function applies a

rectangular window to the potential and current traces and performs a discrete Fourier

transform. This was performed on the ‘raw’ reconstructed traces.

5. Delay-correction was then performed by removing the data points associated with times

between 20-120 µs from the beginning and end of the reconstructed current and potential

traces, respectively, prior to determining phase differences as in #4. It was determined that for

these data, delay-correction of 100-120 µs provided phase differences most aligned with those

recorded on the commercial EIS instrument.

246

6. These delay-corrected phase differences were then used in determining real and imaginary

impedances, which were used to determine capacitive reactance and capacitance.

C.4.1 Notes

• For impedance measurements made using a 1s, 25 mV sinusoidal waveform with a 1 Hz

application frequency and 100 kHz sampling rate (Figures 8.2 and 8.3), no additional data

processing was performed prior to fitting to a sum of sines function; no time-delay observed

on the current response. Phase angles determined using the fit coefficients produced results

consistent with those measured using the commercial EIS instrument (slightly less negative at

each frequency) and with the delay-corrected FSCV:EIS data. This is likely due to improved

determination of phase by the fit over the course of 1 s versus 70 ms.

• Impedance calculations on the multifrequency EIS segment were also done using the power

spectrum amplitudes (V2 or I2) following a discrete Fourier transform, and are consistent with

values determined using the sum of sines fit. The data were shown using the sum of sines fit

due to the relative simplicity of the method.

247

C.5 Supplemental Figures

Figure C.1. Examples of time-delay introduced into the EIS current response using the multiplexed setup; current should precede potential. Raw potential (black; right axis) and current (blue; left axis) shown for the first 2 ms of the 70 ms impedance measurement. Labels for the EIS frequency are displayed above each plot. The time-delay is circled in red in each plot.

248

Figure C.2. Comparison of 100 Hz, 500 Hz, and 1000 Hz single- (solid) and multi-frequency (checkered) FSCV:EIS measurements from either raw (black), 100 µs delay-corrected (red), 120 µs delay-corrected (blue) data; expanded data set from data shown in Figure 8.6B-C. A) Comparison of phase angle (n=3 electrodes) to those recorded for a set of electrodes (n=15) using a commercial EIS instrument (orange striped). B) Comparison of impedance (n=3 electrodes). C) Comparison of capacitive reactance (only multi-frequency) and capacitance determined from a fit to capacitive reactance data. Differences in delay-corrected reactance measurements determined using a one-way ANOVA with Bonferroni post-test for each frequency; F(2,4)=75.25100Hz and 195.51000Hz, p=0.0007100Hz and 0.00011000Hz. Differences in delay-corrected capacitances determined from both single- and multi-frequency determined using a two-way ANOVA with Bonferroni post-test; single- versus multi-frequency: F(1,2)=7.513, p=0.11, and delay-correction amount F(2,4)=41.5, p=0.002.

249

C.6 Additional Methods

C.6.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. In vitro electrochemical experiments were performed in phosphate-buffered saline (10 mM Na2HPO4, 138 mM NaCl, and

2.7 mM KCl). Adrenal slice experiments used bicarbonate-buffered saline (125 mM NaCl, 26 mM

NaHCO3, 2.5 mM KCl, 2.0 mM CaCl2, 1.0 mM MgCl2, 1.3 mM NaH2PO4, 10 mM HEPES, and

10 mM glucose) saturated with 95% O2 and 5% CO2. All buffers were adjusted to pH 7.4 with 1

M NaOH and 1M HCl. All aqueous solutions were made using ultrapure 18.2 MΩ water. Reserpine

(10 mg mL-1) and α-methyl-DL-tyrosine methyl ester hydrochloride (α-MPT, 250 mg mL-1) solutions for animal administration were prepared in 50% Saline: 50% dimethyl sulfoxide and in saline, respectively.

C.6.2 Microelectrode Fabrication

Cylindrical carbon-fiber microelectrodes were fabricated using T-650 carbon fibers (Cytec

Industries, Inc., Woodland Park, NJ) as previously described. In short, a single carbon fiber was aspirated into a glass capillary, a glass seal was created via a micropipette puller (Narishige, Tokyo,

Japan), and the fiber extending past the seal was cut to 100 μm in length. Electrical connection to the carbon-fiber was established via a conductive silver epoxy. A Ag/AgCl pellet reference electrode (World Precision Instruments, Inc., Sarasota, FL) was used to complete the two-electrode electrochemical cell.

C.6.3 Model Circuit Fabrication and Characterization

Circuit component values were confirmed with a voltmeter. Five circuits were fabricated with a 10 kΩ resistor in series with either a 60, 110, 160, 210, or 310 kΩ resistor, which was in

250

parallel with an interchangeable 0, 3.93, 7.43, 10.19, or 13.82 nF capacitor. EIS Spectrum Analyzer software (available online at http://www.abc.chemistry.bsu.by/vi/analyser/) was used to determine impedance values.

C.6.4 Flow-Injection System

Working electrodes were positioned in a custom electrochemical flow-cell using a micromanipulator (World Precision Instruments, Sarasota, FL). A syringe pump (New Era Pump

Systems, Wantagh, NY) supplied a continuous buffer stream of 1.0 mL min-1 across the working and reference electrodes. Two-second bolus injections of analyte to the microelectrode surface were accomplished with a six-port HPLC valve mounted on a two-position actuator controlled by a digital pneumatic solenoid valve (Valco Instruments, Houston, TX). The entire apparatus was enclosed in a custom-built grounded Faraday cage.

251

C.7 References

1. Zhivomirov, H. Phase Difference Measurement with Matlab Implementation (version 1.3.0.0). Available at: https://www.mathworks.com/matlabcentral/fileexchange/48025-phase- difference-measurement-with-matlab-implementation

252

APPENDIX D

Poly-Dioxythiophene and Nafion Copolymer Membranes for Enhanced Voltammetric

Peptide Selectivity

This appendix contains a brief summary of a project that I have worked on, but that are not ready for publication. This work was done in collaboration with Dr. James Roberts.

D.1 Introduction

Selectivity is always of great concern when monitoring molecules electrochemically. With respect to fast-scan cyclic voltammetry (FSCV) detection of neuropeptides (Chapter 3), tyrosine and methionine residues are oxidizable and detectable at relevant concentration (low nM to low

μM).1,2 However, any fluctuating tyrosine or methionine containing peptide could elicit a similar volumetric response. Of particular interest is a family of endogenous opioid neuropeptides called enkephalins, that are small peptide (<10 residues) neurotransmitters implicated in reward, pain, addiction, and feeding behavior. A common approach to enhance selectivity is to incorporate a selective membrane on the electrode surface that generally function using electrostatic or size- exclusion mechanisms. One such electrode coating involves the electrodeposition of a copolymer containing PEDOT, or poly(3,4-ethylenedioxythiophene), and Nafion, a sulfonated tetrafluoroethylene polymer.3 Application of this coating on carbon fiber microelectrodes

(CFMEs) for FSCV was demonstrated to enhance selectivity (pH 7.4) for cationic neurotransmitter dopamine (pKa ~9) through charge-repulsion of anionic species such as ascorbic acid (pKa ~4), presumably through interactions with the Nafion sulfonate functional groups.

253

In this work, FSCV was performed using the modified sawhorse waveform (MSW; Chapter

3) on a series of relevant tyrosine and/or methionine containing neuropeptides differing in size: methionine enkephalin (M-ENK; 5 residues), oxytocin (OXY; 9 residues), substance P (SubP; 11 residues), neurotensin (NT; 13 residues), and neuropeptide Y (NPY; 36 residues). Then, a PEDOT and Nafion copolymer (PEDOT:Nafion) was electrodeposited (Figure D.1) onto CFMEs using cyclic voltammetry and demonstrated to enhance selectivity for smaller peptide species, namely

M-ENK. The ability to tailor the selectivity of the coating was demonstrated by exchanging

PEDOT for larger poly-dioxythiophenes ProDOT, poly(3,4-propylenedioxythiophene), and

ProDOT-Et2, or poly(3,4-(2′,2′-diethylpropylene)dioxythiophene).

Figure D.1. Schematic of the copolymerization of PEDOT and Nation that coats the CFME surface.

D.2 Methods

D.2.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. Peptides (≥95%, HPLC assay) were purchased from Cayman Chemical (Ann Arbor, MI). In vitro electrochemical

254

experiments were performed in either a phosphate-buffered saline (10 mM Na2HPO4, 138 mM

NaCl, and 2.7 mM KCl). All buffers were adjusted to pH 7.4 with 1 M NaOH and 1 M HCl.. LQ-

1105 Nafion a 5% wt Nafion solution in methanol, was purchased from Ion Power, Inc. (DE,

USA).

D.2.2 Data Acquisition

For FSCV recordings, a modified sawhorse waveform (MSW) scanning from -0.4 to +0.7

V at 100 V s-1, then to +1.2/1.3 V at 600 V s-1, which is held for 3 ms before returning to -0.4 V at

100 V s-1; waveform application frequency was 5 Hz and a sampling rate of 100 kHz was used for all measurements displayed. a custom-built instrument for potential application and current transduction (University of North Carolina at Chapel Hill, Department of Chemistry, Electronics

Facility). High Definition Cyclic Voltammetry software (University of North Carolina at Chapel

Hill) was used for waveform output, data analysis and signal processing in conjunction with data acquisition cards (National Instruments, Austin TX) used for measuring current and synchronizing the electrochemical cell with the flow-injection system. Analog filtering of the applied waveform was accomplished with a 2-pole Sallen-Key, low-pass filter at 50 kHz. There was no additional analog or digital filtering of the data.

D.2.3 Electropolymerization

Electropolymerization was performed using cyclic voltammetry with a triangular waveform scanning from -0.8 V to 1.5 V at 100 mV s-1 for 15 cycles. The deposition solution contained 100-200 μM of the dioxythiophene monomer and 0.05 wt% Nafion in acetonitrile which was stirred for at least 15 minutes prior to use and used within 48 hours. Following electrodeposition electrodes were rinsed with water and allowed to dry for at least 15 min prior to use in electrochemical experiments.

255

D.3 Results and Discussion

D.3.1 Peptide FSCV Before and After Electrodeposition of PEDOT:Nafion Coating

Previous work on peptide detection within our group focused on detection of enkephalin opioid neuropeptide transmitters, methionine and leucine enkephalin (M-ENK and L-ENK), using a modified sawhorse waveform (MSW). Selectivity was established between via the lack of current generated through oxidation of methionine residues.1,2 Figure D.2A displays the MSW, where tyrosine oxidizes within the high scan rate window and methionine during the 3 ms amperometric hold. However, there are a large number of neuropeptide transmitters implicated in an array of functions in the central nervous system and likely a larger number of unknown peptide species. To demonstrate the effectiveness of FSCV for detection of a number of important neuropeptides, and the need to enhance selectivity, recordings were made for OXY, SubP, NT, and NPY. These measurements were made prior to and following electrodeposition of a PEDOT:Nafion coating

(200 μM monomer in deposition solution). To increase confidence in coating viability, similarities between reported deposition voltammogram and those herein, as well as enhanced dopamine selectivity against ascorbic acid were confirmed.3

256

Figure D.2. Peptide FSCV before and after application of PEDOT:Nafion coating. A) Potential waveform used for peptide detection. B) Representative deposition voltammogram for the oxidation of dioxythiophene monomer to a polymer. C) Representative voltammograms for peptides (2 μM bolus) before and after electropolymerization. D) Normalized current (to precoating maximum) before and after coating application for M-ENK, NT, and NPY. Standard error is reported (n=4 electrodes).

257

Figure D.2 displays the MSW, representative electrodeposition CVs, and FSCV peptide recordings before and after application of the PEDOT:Nafion coating. The deposition CV displays a dominant peak at the switching potential (+1.5 V) associated with oxidation of 3,4- ethylenedioxythiophene to PEDOT; successive scans result in a slight decrease in the amplitude of this signal (not shown). In Figure D.2C, peptide detection was successful the peptides selected.

Note that NT does not contain methionine and SubP does not contain tyrosine. Following application of a PEDOT:Nafion coating, all peptide signals are attenuated, with the greatest attenuation occurring for the larger peptides. Figure D.2D summarizes the significant attenuation of M-ENK, NT, and NPY signals via application of the coating; paired two-tailed t-test: M-ENK t(3)=4.01, *p=0.03, NT t(3)=16.45, ***p=0.0005, NPY t(3)=8.921, **p=0.003. Quantification of current response is from the oxidation potential of tyrosine (~+1.0 V). These data suggest that application of a PEDOT:Nafion coating are beneficial for selective detection of small peptide fluctuations such as M-ENK.

The sources of varying responsivity are currently under investigation. In the case of NT and NPY, these peptides contain multiple tyrosine residues that may be accessible to be oxidized, which may explain enhanced responsivity. Additional investigations are targeted at understanding adsorption of various peptide residues and lengths on FSCV responsivity.

D.3.2 Tailoring the Copolymer Coating to Alter Peptide Selectivity

The PEDOT:Nafion coating would excessively attenuate the signals of peptides slightly larger than M-ENK. However, NPY detection would still benefit from the ability to selectively

‘filter out’ even larger peptides. To demonstrate the ability to tailor the selectivity of the copolymer coating, the dioxythiophene monomer in the deposition solution was altered to create

ProDOT:Nafion and ProDOT-Et2:Nafion coatings; PEDOT:Nafion coatings were also deposited

258

with a lower monomer concentration (100 μM). Figure D.3 displays the effect of deposition conditions on dopamine and ascorbic acid selectivity. Previous studies of PEDOT polymerized in the presence of sulfonates have demonstrated an excess of sulfonate groups following polymerization.4 This would result in a non-charge-balanced coating that is net negative. The data at PEDOT:Nafion (200 μM) coatings support this claim, as significantly signal enhancement and attenuation are observed for both cationic dopamine and anionic ascorbic compared to bare

CFMEs; one-way ANOVA with Bonferroni post-test comparing all groups, F(4,20)=18.69,

****p<0.0001. The other coatings appear to marginally enhance dopamine selectivity, but the differences were not found to be significant from the bare CFMEs. Thus, it is likely that the composition or incorporation of Nafion/sulfonate groups in the other copolymer preparations is altered.

Figure D.3. Dopamine (1 μM) and ascorbic acid (100 μM) selectivity for different copolymer coatings. Left) Cyclic voltammograms for dopamine (top) and ascorbic acid (bottom) before and after (red) deposition of a PEDOT:Nafion coating. Right) A summary of the max current ratio resulting from different monomer composition in the deposition solution. Standard error is reported (n=5 electrodes each).

259

To investigate the effect of different poly-dioxythiophenes on the resulting peptide selectivity imparted by the copolymer coating, FSCV detection of the smallest (M-ENK) and largest (NPY) peptides was evaluated following application of Nafion and PEDOT, ProDOT, or

ProDOT-Et2, with a monomer concentration of 200 μM; 100 μM PEDOT:Nafion coatings displayed negligible attenuation of all peptides used herein (not shown). Figure D.4 displays the results for peptide responses before and after electrodeposition. While all poly- dioxythiophene:Nafion coatings attenuate both M-ENK and NPY signals, it appears that increasing the size of the monomer lessens the attenuation of peptide signals, demonstrating the ability to tailor the coating. These data also suggest that while the PEDOT and ProDOT containing coatings maximized M-ENK:NPY selectivity, the ProDOT:Nafion coating offered superior sensitivity for M-ENK. Furthermore, by enlarging the poly-dioxythiophene in the copolymer, a coating allowing passage of larger peptides can be fabricated.

260

Figure D.4. Copolymer coating effect on peptide FSCV peptide selectivity. Representative voltammograms (left; 2 μM) and averaged maximum current response (right) for A) M-ENK and B) NPY at bare and coated CFMEs. C) Current ratio for M-ENK:NPY resulting from incorporation of poly-dioxythiophene in the copolymer. Standard error is reported (n=19 bare electrodes; n=5 coated electrodes).

D.4 Conclusion

FSCV is capable of monitoring neuropeptide fluctuations through the oxidation of tyrosine and methionine residues. However, selectivity is a concern as many tyrosine and methionine containing peptide signals are very similar, and the peptide composition in vivo is highly diverse.

To enhance selectivity against larger peptides, we demonstrated that electrodeposition of a poly- dioxythiophene:Nafion copolymer coating enhances selectivity for and minimally attenuates the signals for M-ENK, compared to larger peptides such as OXY, SubP, NT, and NPY. Furthermore, it was shown that the copolymer system can be altered to alter selectivity and enlarge the size of

261

detectable peptides. Thus, these copolymer coating should be beneficial in electrochemical detection of relatively small neuropeptide transmitters in biological studies. By excluding larger peptides from the CFME surface, confidence in the identity of FSCV monitored peptide fluctuations will be improved.

Continued work on this project will involve application of this coating in live tissue or in vivo for peptide detection. Previous work done with the PEDOT:Nafion copolymer demonstrate that this coating is biocompatible and reduces biological fouling, which may further enhance the ability to monitor peptide dynamics. Furthermore, it is likely that both charge and size-exclusion mechanisms are vital to the performance of these coatings. Thus, additional studies will be aimed at elucidating the extent of these mechanisms.

262

D.5 References

1. Schmidt, A. C.; Dunaway, L. E.; Roberts, J. G.; McCarty, G. S.; Sombers, L. A. Multiple Scan Rate Voltammetry for Selective Quantification of Real-Time Enkephalin Dynamics. Anal. Chem. 2014, 86 (15), 7806–7812. https://doi.org/10.1021/ac501725u.

2. Calhoun, S. E.; Meunier, C. J.; Lee, C. A.; McCarty, G. S.; Sombers, L. A. Characterization of a Multiple-Scan-Rate Voltammetric Waveform for Real-Time Detection of Met-Enkephalin. ACS Chem. Neurosci. 2019, 10 (4), 2022–2032. https://doi.org/10.1021/acschemneuro.8b00351.

3. Vreeland, R. F.; Atcherley, C. W.; Russell, W. S.; Xie, J. Y.; Lu, D.; Laude, N. D.; Porreca, F.; Heien, M. L. Biocompatible PEDOT:Nafion Composite Electrode Coatings for Selective Detection of Neurotransmitters in Vivo. Anal. Chem. 2015, 87 (5), 2600–2607. https://doi.org/10.1021/ac502165f.

4. Zotti, G.; Zecchin, S.; Schiavon, G.; Louwet, F.; Groenendaal, L.; Crispin, X.; Osikowicz, W.; Salaneck, W.; Fahlman, M. Electrochemical and XPS Studies toward the Role of Monomeric and Polymeric Sulfonate Counterions in the Synthesis, Composition, and Properties of Poly(3,4-Ethylenedioxythiophene). Macromolecules 2003, 36 (9), 3337–3344. https://doi.org/10.1021/ma021715k.

263

APPENDIX E

Fourier Analysis of FSCV Recordings to Identify Time-Points of Molecular Fluctuations

This appendix contains a brief summary of a promising project that I have initiated, but that is not ready for publication.

E.1 Introduction

Selectivity is always of great concern when monitoring molecules electrochemically. Fast- scan cyclic voltammetry (FSCV) done at a carbon fiber microelectrodes (CFMEs) for example, is quite effective for monitoring rapid molecular fluctuations in live brain tissue and has been invaluable in studies uncovering links between neurotransmission and various behavior and pharmacology.1 Motivation to explore new brain regions, as well as advances in electrode materials and sensing strategies have increased the detectable pool of analytes and enabled minimally invasive recording of multiple analytes simultaneously at a single recording site.2–5

However, electrochemical drift and difficulty elucidating mixed analyte/interferent signals render data analysis time-consuming and limit FSCV to monitoring only short molecular fluctuations (ms to s).6,7 A number of approaches exist to reduce or account for electrochemical drift as well as to elucidate mixed signals in FSCV recordings. However, these methods generally require careful construction of a multivariate model (ex., principal component regression and partial least-squares regression), or suffer from a significant reduction of temporal resolution.

One approach yet to be applied to FSCV is analyze recordings in the frequency domain to gain an additional dimension of selectivity. Herein, FSCV recordings were visualized in the frequency domain by applying a Fourier transformation to cyclic voltammograms (potential versus

264

current) converted to the time domain (potential or current versus time). By interpreting the current response in the frequency domain, the harmonic composition can be used to elucidate nonfaradaic from faradaic current contributions. In brief, nonfaradaic current, the main component of electrochemical drift, is predominately isolated into the primary harmonic(s) of the applied potential. Faradaic contributions can be spread throughout the harmonics. This occurs because nonfaradaic current (capacitance) response is approximately linear with applied potential, whereas faradaic current response (heterogeneous electron transfer) is nonlinear and generally follows the

Bulter-Volmer relationship; current density changes as a function of (over)potential.8 Thus, extended FSCV recordings can be quickly processed and fluctuations in higher order harmonics in cyclic voltammograms recorded over time can be used to identify time-points on which to focus analyses. Furthermore, for a given potential waveform, different analytes can have differing harmonic ‘fingerprints’ that may be useful in analyte identification as in infrared or nuclear magnetic resonance spectroscopy.9,10

E.2 Methods

E.2.1 Chemicals

All chemicals (≥95%, HPLC assay) were purchased from Sigma-Aldrich Co. (St. Louis,

MO), unless otherwise stated, and used without additional processing. In vitro electrochemical experiments were performed in either a phosphate-buffered saline (10 mM Na2HPO4, 138 mM

NaCl, and 2.7 mM KCl). All buffers were adjusted to pH 7.4 with 1 M NaOH and 1 M HCl.

E.2.2 Data Acquisition

A triangular waveform scanning from −0.4 V to +1.3 or 1.4 V using a scan rate of 400 V s−1 was applied with a frequency of 10 Hz, sampling at ∼37 kHz, using either a custom-built

265

instrument for potential application and current transduction (Universal Electrochemistry

Instrument, University of North Carolina at Chapel Hill, Department of Chemistry, Electronics

Facility). High-Definition Cyclic Voltammetry software (HDCV, University of North Carolina at

Chapel Hill) was used for waveform output, signal processing, and data analysis, in conjunction with data acquisition cards (National Instruments, Austin TX) used for measuring current and synchronizing the electrochemical cell with the flow-injection system. Analog filtering of the applied waveform was accomplished with a 2- pole Sallen-Key, low-pass filter at 2 kHz. There was no additional analog or digital filtering of the data.

E.2.3 Data Analysis

All data were examined manually in HDCV before exporting to MATLAB R2018a. The sampling rate was used to convert each data point to time in order to plot cyclic voltammogram current response in the time domain. The data were then repeated 5x to generate data for frequency domain analyses. A fast Fourier transform was performed on both the waveform and each current response and a power density spectrum (PDS) was generated for each voltammogram recorded over time. This enables the harmonic(s) magnitude to be monitored as a function of time.

E.3 Results and Discussion

E.3.1 Fourier Transform Fast-Scan Cyclic Voltammetry (FT-FSCV)

An FSCV waveform is triangular and composed of primarily of odd harmonic components.

In this work, FSCV waveforms of either 8.5 or 9 ms in length are applied corresponding to primary frequencies of ~118 Hz and ~111 Hz, respectively. In a flow cell, a CFME was used to record

FSCV during bolus injections of hydrogen peroxide (H2O2) and catecholamines dopamine, norepinephrine, and epinephrine. Figure E.1 displays representative visualized in both the potential

266

domain (voltammograms) and in the frequency domain. Note that the catecholamines have a very similar response (Figure E.1B) with only epinephrine differing due to an additional observable oxidation.11 Figures E.1C-H display various normalized power spectra (PS) generated following a fast Fourier transform was performed. As expected, the PS of the applied triangular potential waveform is composed predominately of odd harmonics (very small contributions), with its fundamental contribution centered at ~118 Hz. The PS for the predominately nonfaradaic current also follows a similar trend, with the major contributions occurring in the odd harmonics, as expected. Importantly, the PS for the faradaic current containing voltammograms displays variable harmonic composition spread across a wider range of frequencies. Furthermore, distinct differences between PDS of H2O2 and the catecholamines, and some minor differences between the catecholamines are observed. Thus, Fourier analysis of the FSCV recordings should provide information effective for critical interpretation of FSCV recordings.

267

Figure E.1. Visualization of representative FSCV data in both the potential and frequency domains. A) Normalized background and H2O2 cyclic voltammograms (top) and power spectra for the potential waveform as well as the background and H2O2. B) Normalized catecholamine cyclic voltammograms (top) and power spectra (bottom).

268

E.3.2 Identification of Important Time-Points in Extended Recordings Using FT-FSCV

To demonstrate one example of the utility of this type of analysis, extended FSCV recordings (>10 min) were made in an electrochemical flow-cell during which bolus injections of both dopamine and H2O2 were performed. Figure E.2A-B displays a background-subtracted

(time=0) color plot for a 12-minute recording as well as the total absolute current of each cyclic voltammogram over time; the absolute values of each current data point of a cyclic voltammogram were summed. A large amount of electrochemical drift is observed in this recording, as indicated by the forming color streaks in the color plot and the continual rise in total current over time.

Conventionally, an in vivo recording like this would be manually analyzed in 30-60 second segments by repositioning the reference point for background subtraction in search of current fluctuations indicative of a target analyte.

Figure E.2C displays a frequency domain view of the data displayed in the color plot, where the power spectra density amplitude from ~111 Hz (fundamental frequency) to ~666 Hz are displayed versus time. The odd harmonics amplitude gradually increases over time similar to

Figure E.2B, indicative of nonfaradaic electrochemical drift. Yet, the even harmonics amplitude remains relatively close to baseline throughout the duration of the recording and fluctuations occur time-locked with the triplicate bolus injections for dopamine (first half of recording) and H2O2

(second half). The amplitudes are even indicative of analyte concentration (largest concentrations first). Thus, this analysis strategy could be valuable in facilitating data analysis by indicating time- points of interest in extended FSCV recordings thereby enhancing throughput.

269

Figure E.2. Frequency domain analyses help discern electrochemical drift from analyte signals. A) Color plot of a flow-cell recording in which bolus injections of dopamine and H2O2 are performed in triplicate. B) Total absolute current of each voltammogram over over time. The red and blue bars indicate time during which dopamine and H2O2 injections occurred, respectively. C) Power spectral density amplitudes at various harmonics for each voltammogram versus time.

E.4 Conclusion

FSCV is a very useful analytical tool to monitor neurochemical fluctuations in vivo. Many studies involve continuous FSCV recording during behavioral tasks or pharmacological manipulation. However, analysis of these long recordings is generally done manually, in segments, due to electrochemical drift. Herein, a straightforward Fourier analysis was performed on extended

FSCV recordings. Differences in the frequency harmonic composition of nonfaradaic and faradaic

270

current contribution were then utilized to confirm time-points containing faradaic current fluctuations from redox processes of dopamine and H2O2. This analysis tool is easy to implement, can be applied to any FSCV recording, and provides a wealth of information that would be useful to researchers.

To further this work, effectiveness in identifying time-points of electrically evoked catecholamine release during in vivo recordings will need to be confirmed. Furthermore, the ability to use the frequency composition as a ‘fingerprint’ to elucidate analyte signals displays promise and would be an interesting pursuit. Importantly, another dimension of selectivity not mentioned herein would be to determine the phase of each frequency harmonic. The frequency and phase composition have been used to elucidate different peptides.10 Therefore, it is likely that reliable determination of phase would effectively enhance selectivity further.

271

E.5 References

1. Michael, A. C.; Borland, L. Electrochemical Methods for Neuroscience; CRC press, 2006.

2. Smith, S. K.; Lee, C. A.; Dausch, M. E.; Horman, B. M.; Patisaul, H. B.; McCarty, G. S.; Sombers, L. A. Simultaneous Voltammetric Measurements of Glucose and Dopamine Demonstrate the Coupling of Glucose Availability with Increased Metabolic Demand in the Rat Striatum. ACS Chem. Neurosci. 2017, 8 (2), 272–280. https://doi.org/10.1021/acschemneuro.6b00363.

3. Schmidt, A. C.; Wang, X.; Zhu, Y.; Sombers, L. A. Carbon Nanotube Yarn Electrodes for Enhanced Detection of Neurotransmitter Dynamics in Live Brain Tissue. ACS Nano 2013, 7 (9), 7864–7873. https://doi.org/10.1021/nn402857u.

4. Puthongkham, P.; Yang, C.; Venton, B. J. Carbon Nanohorn-Modified Carbon Fiber Microelectrodes for Dopamine Detection. Electroanalysis 2018, 30 (6), 1073–1081. https://doi.org/10.1002/elan.201700667.

5. Cryan, M. T.; Ross, A. E. Scalene Waveform for Codetection of Guanosine and Adenosine Using Fast-Scan Cyclic Voltammetry. Anal. Chem. 2019, 91 (9), 5987–5993.

6. Meunier, C. J.; McCarty, G. S.; Sombers, L. A. Drift Subtraction for Fast-Scan Cyclic Voltammetry Using Double-Waveform Partial-Least-Squares Regression. Anal. Chem. 2019, 91 (11), 7319–7327. https://doi.org/10.1021/acs.analchem.9b01083.

7. Rodeberg, N. T.; Johnson, J. A.; Cameron, C. M.; Saddoris, M. P.; Carelli, R. M.; Wightman, R. M. Construction of Training Sets for Valid Calibration of in Vivo Cyclic Voltammetric Data by Principal Component Analysis. Anal. Chem. 2015, 87 (22), 11484–11491.

8. Bard, A. J.; Faulkner, L. R.; Leddy, J.; Zoski, C. G. Electrochemical Methods: Fundamentals and Applications; wiley New York, 1980; Vol. 2.

9. Brazill, S. A.; Bender, S. E.; Hebert, N. E.; Cullison, J. K.; Kristensen, E. W.; Kuhr, W. G. Sinusoidal Voltammetry: A Frequency Based Electrochemical Detection Technique. J. Electroanal. Chem. 2002, 531 (2), 119–132. https://doi.org/10.1016/S0022-0728(02)01067- 7.

10. Brazill, S. A.; Singhal, P.; Kuhr, W. G. Detection of Native Amino Acids and Peptides Utilizing Sinusoidal Voltammetry. Anal. Chem. 2000, 72 (22), 5542–5548. https://doi.org/10.1021/ac000543r.

11. Pihel, Karin.; Schroeder, T. J.; Wightman, R. Mark. Rapid and Selective Cyclic Voltammetric Measurements of Epinephrine and Norepinephrine as a Method To Measure Secretion from Single Bovine Adrenal Medullary Cells. Anal. Chem. 1994, 66 (24), 4532– 4537. https://doi.org/10.1021/ac00096a021.

272