gawk" g, A .1... .2 x ,me..an « v1. 5 . agau. 35.2 :1”. .45 :u., » FL. .v 5;. I :5 ..,?x.r..za......r. :t. .9. A fl. ‘7 Z .. V -‘.p..-¢..n~san .. : , m . ‘ u . , : ‘ This is to certify that the dissertation entitled UVN IS AND IR SPECTROELECTROCHEMICAL INVESTIGATIONS OF SMALL MOLECULES AND REDOX-ACT IVE BIOMOLECULES USING OPTICALLY TRANSPARENT DIAMOND ELECTRODES presented by Shannon {Marie Haymonzf has been accepted towards fulfillment of the requirements for PhD. degree in Chemistry Jug/14% V a Major pr'ofessor Date 13 November 2002 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 W; MiChiQan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIFiC/DateDuepssosz -w-HW UVNIS AND IR SPECTROELECTROCHEMICAL INVESTIGATIONS OF SMALL MOLECULES AND REDOX-ACTIVE BIOMOLECULES USING OPTICALLY TRANSPARENT DIAMOND ELECTRODES By Shannon Marie Haymond A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2002 ABSTRACT UVNIS AND IR SPECTROELECTROCHEMICAL INVESTIGATIONS OF SMALL MOLECULES AND REDOX-ACTIVE BIOMOLECULES USING OPTICALLY TRANSPARENT DIAMOND ELECTRODES By Shannon Marie Haymond Spectroelectrochemistry combines two very different techniques— spectroscopy and electrochemistry—providing more information about the redox mechanism of an electroactive analyte than can be acquired using traditional electrochemical methods alone. The oxidation state of an analyte is changed electrochemically by adding or removing electrons at an optically transparent electrode (OTE) while the spectroscopic response of the species in solution is simultaneously measured. Spectroelectrochemistry is useful for obtaining spectra and redox potentials for electrogenerated species as well as important information in the investigation of electrode reaction mechanisms, in electroanalytical applications, and, as demonstrated herein, in elucidating structural information about redox-active protein active sites. The use of electrically conductive diamond as an OTE is a new area of research. Boron-doped diamond (BDD) possesses attractive qualities as both an electrode and an optically transparent material, making it an obvious choice for utilization as an OTE in transmission spectroelectrochemical measurements. Diamond OTES offer several advantages over other materials including: (i) the possibility of optical measurements from the near-UV into the far-IR (0.25— 100 pm), (ii) a low background current,‘(iii) a wide working potential window, (iv) stability in aqueous and nonaqueous solution environments during both cathodic and anodic polarization, and (v) resistance to fouling. An added advantage of BDD OTES is the versatility of fabrication, as BDD thin-films can be grown on a variety of substrates (e.g., quartz, undoped Si, and white diamond grown by chemical vapor deposition), and can also exist as freestanding disks. Each type of OTE possesses characteristics that can be exploited for a specific application. This work is a product of our long-term goal to develop BDD as an OTE for electroanalytical applications and the study of biological electron transfer mechanisms. Development of the OTE comprised the design and testing of thin-layer spectroelectrochemical cells and the electrochemical and spectroelectrochemical characterization of two systems: (i) a small molecule, ferrocene, in nonaqueous solvent/electrolyte systems, and (ii) a redox-active protein, cytochrome 0. Besides the basic electrochemical and spectroelectrochemical characterizations of these systems, this work focused on determining the deposition conditions that produce an optimized balance between electrical conductivity and optical transparency in this material. to Jerry ACKNOWLEDGEMENTS This journey was in no way a lonely one, and for their various contributions to my scientific progress or social well being, I would like to acknowledge the following people: To the past and present members of the Art’s Softball Team for many fun times and amusing memories on, and off of the field at a fine establishment where peaches are served with beer, and people always win at the skill crane and sing loudly with the jukebox. To my co-hostesses and co-counsel, the notorious Asphalt Angel Hudy and Sweet Cheeks Lamp: I consider our transformation of 5077 Wardcliff into the “Hog Wyld Roadhouse” and subsequent, victorious legal battle to be among my greatest accomplishments! You two have been great friends and roommates and I will miss you both. To the current members of the Babcock group for displaying the guts and composure to make the best of a bad situation and for their input and assistance throughout my time here. And I can’t forget my former labmates and friends that have gone on to further greatness, Kristi, Heather, and Jose, although you guys are out of sight, you will never be out of mind. I think I speak for the entire Babcock group when I give my heartfelt thanks to our personal grammar lady, Vada O’Donnell, for keeping us true to Jerry’s values of pristinely correct grammar usage, and for always bringing out the best in us. Additionally, I’d like to thank the members of the Swain group for their help and support. I am fortunate to have gone through this dissertation writing and defending process with my dear friends Maggie and Matt. For their scientific collaboration, I am indebted to following people, as their contributions were significant and integral to the work presented herein: Professor Jerzy Zak for sharing with me his expertise in spectroelectrochemical cell design and thin-layer electrochemistry, to Zuzana Cvackova for continuing the voltammetric study of cytochrome c at boron-doped diamond electrodes, to Professor Shelagh Ferguson-Miller, Dr. Denise Mills, and Bryan Schmidt for demonstrating interest in my projects and, thus, providing me with high-caliber protein samples, and to Glenn Wesley for his ability to turn my crude designs into functioning pieces. To Dr. Warwick Hillier for his guidance, encouragement, and contributions to my scientific development. Thanks for answering all of my mundane questions and teaching me that sometimes turning the “do not touch” knob is allowed. I am fortunate that Jerry hired you “for my amusement”. Good luck on your bowling game, Pip. To Professor Jerry Babcock for instilling in me a desire to be great at what I enjoy, scientific and otherwise. I have strived to carry on as you wished, and for that, I leave here more daring and resourceful. To my other great advisor, Professor Greg Swain, for providing stability and support during a difficult transition. I would also like to thank you for the guidance and encouragement you provided as you took over the presidential role of the SH fan club. To my family for their pride in me, for all that I choose to do. Without you all, I would have less to my credit, personally and otherwise. Thank you for your vi unending support. To my love, Ryan, for indulging my stubbornness and Independence while always pulling me back up when I called from the basement of the chemistry building, literally, in the pits of despair. Thanks for always making me laugh and for showing continuous faith in my pursuit of this degree. vii TABLE OF CONTENTS LIST OF TABLES ................................................................................................. xi LIST OF FIGURES ............................................................................................. xiii CHAPTER 1 1 Introduction 1.1 Insight into Metalloprotein Mechanisms .................................................... 1 1.2 Spectroelectrochemical Methods .............................................................. 3 1.3 Optically Transparent Electrodes ............................................................ 10 1.3.1 Conventional Types .......................................................................... 10 1.3.2 A New Material: Boron-Doped Diamond ........................................... 11 1.4 Scientific Relevance ................................................................................ 14 1.5 Dissertation Outline ................................................................................. 16 1.6 References .............................................................................................. 19 CHAPTER 2 2 Materials and Methods 2.1 Boron-Doped Diamond Growth Conditions ............................................. 34 2.1.1 Microcrystalline Boron-Doped Diamond ............................................ 34 2.1.2 Nanocrystalline Boron-Doped Diamond ............................................ 38 2.1.3 Freestanding Boron-Doped Diamond OTE ....................................... 38 2.1.4 Boron-Doped Diamond IR OTE ........................................................ 40 2.2 Sample Preparation ................................................................................ 40 2.2.1 Ferrocene .......................................................................................... 40 2.2.2 Cytochrome c .................................................................................... 41 2.2.3 Cytochrome c Oxidase ...................................................................... 41 2.3 Cyclic Voltammetric Experiments ............................................................ 42 2.4 Spectroelectrochemistry .......................................................................... 44 2.4.1 000 Experiments - Au Grid OTTLE Cell ........................................... 44 2.4.2 BDD UV/vis OTTLE Cell ................................................................... 48 2.4.3 BDD IR OTTLE Cell I ........................................................................ 50 2.4.4 800 IR OTTLE Cell ll ....................................................................... 53 2.5 Temperature-Control Set Up ................................................................... 54 2.6 References .............................................................................................. 58 viii CHAPTER 3 3 Ferrocene Electron Transfer Kinetics Measured at Boron-Doped Diamond Electrodes 3.1 Introduction ............................................................................................. 60 3.2 Results .................................................................................................... 64 3.3 Conclusions ............................................................................................ 72 3.4 References .............................................................................................. 73 CHAPTER 4 4 Direct Voltammetry of Cytochrome c at Boron-Doped Diamond Electrodes 4.1 Introduction ............................................................................................. 76 4.2 Results .................................................................................................... 81 4.2.1 Microcrystalline BDD Thin-Film Electrodes ....................................... 82 4.2.2 Nanocrystalline BDD Thin-Film Electrodes ....................................... 94 4.3 Discussion ............................................................................................. 102 4.4 Conclusions .......................................................................................... 1 1 1 4.5 References ............................................................................................ 1 12 CHAPTER 5 5 Low-Frequency Electrochemical Difference FTIR Spectroscopy of Metalloproteins: Challenges and Solutions 5.1 Introduction ........................................................................................... 1 17 5.2 Structural lnforrnation Possible (Far-IR Region) ................................... 119 5.3 FTIR of Aqueous Samples .................................................................... 122 5.4 Detectors ............................................................................................... 132 5.5 Window Materials .................................................................................. 136 5.6 Temperature Control ............................................................................. 138 5.7 Conclusions .......................................................................................... 140 5.8 References ............................................................................................ 143 CHAPTER 6 6 Electrochemical Difference FTIR Spectroscopy of Cytochrome c Oxidase 6.1 Introduction ........................................................................................... 147 6.2 Results .................................................................................................. 152 6.2.1 Electrochemical Difference Spectra ................................................ 152 6.2.2 Comparison of Bacterial and Mammalian Forms of CcO ................ 159 6.2.3 IsotopicaIIy-Labeled Spectra ........................................................... 166 6.3 Discussion ............................................................................................. 173 6.3.1 Carbonyl Region ............................................................................. 173 6.3.2 Histidine Modes .............................................................................. 176 6.4 Implications ........................................................................................... 185 6.5 Conclusions .......................................................................................... 187 6.6 References ............................................................................................ 190 CHAPTER 7 7 Spectroelectrochemical Studies Using a Novel Optically Transparent Electrode: Boron-Doped Diamond 7.1 Introduction ........................................................................................... 198 7.2 Results - UV/vis Spectroelectrochemistry ............................................. 206 7.2.1 Optical Properties of BDD OTES — UV/vis ....................................... 206 7.2.2 Electrochemical Characterization of the UV/vis OTTLE Cell ........... 211 7.2.3 Ferrocene ........................................................................................ 212 7.2.4 Cytochrome c .................................................................................. 219 7.3 Results - IR Spectroelectrochemistry .................................................... 223 7.3.1 Optical Properties of BDD OTES — IR ............................................. 223 7.3.2 Ferrocene ........................................................................................ 228 7.3.3 Electrochemical Characterization of the IR OTTLE Cell .................. 230 7.3.4 Ferrocyanide ................................................................................... 233 7.3.5 Cytochrome c .................................................................................. 234 7.4 Conclusions .......................................................................................... 237 7.5 References ............................................................................................ 240 Table 2—1 Table 3—1 Table 4-1 Table 4—2 Table 4—3 Table 5—1 Table 5—2 Table 5-3 Table 6—1 Table 6—2 LIST OF TABLES Mediators Used in Spectroelectrochemical Measurements of Cytochrome c Oxidase .................................................................. 47 Apparent Heterogeneous Electron Transfer Rate Constants for Ferrocene at Untreated Boron-Doped Diamond and Polished Glassy Carbon Electrodes ............................................................. 69 Effect of Cytochrome 0 Exposure on Test Analyte Cyclic Voltammetric Peak Splitting for Acid-Washed Rehydrogenated 10 ppm Microcrystalline Boron-Doped Diamond Electrodes ......... 93 Effect of Cytochrome 0 Exposure on Test Analyte Cyclic Voltammetric Peak Splitting for Acid-Washed Rehydrogenated 10 ppm Nanocrystalline Boron-Doped Diamond Electrodes ........ 102 Apparent Heterogeneous Electron Transfer Rate Constants Measured from the Direct Voltammetry of Cytochrome c at Solid Electrodes ................................................................................... 105 Low-Frequency Vibrational Modes of Metalloenzymes ............... 121 Infrared Properties of H20 and 020 ............................................ 123 Properties of Infrared Transmissive Materials ............................. 138 Discrepant Modes and Tentative Assignments for FTIR Electrochemical Difference Spectra of Cytochrome c Oxidases.. 162 15N Sensitive Peaks in Rhodobacter Sphaeroides Cytochrome c Oxidase FTIR Spectra ................................................................. 169 xi Table 6—3 Table 6—4 Table 6-5 Table 6—6 Table 6—7 Table 7—1 CSN1 Stretching Modes for Histidine and 4-Melm ...................... 181 CSN1 Stretching Modes for Metal-Bound Histidine and 4-Melm .181 C4C5 Stretching Modes for N8 Protonated Histidine and 4-Melm .................................................................................................... 182 C4C5 Stretching Modes for Metal-Bound N8 Protonated Histidine and 4-Melm ................................................................................. 182 Metal-Histidine Stretching Modes ................................................ 186 Effect of Selected Growth Parameters on the Optical and Electronic Properties of Boron-Doped Diamond Optically Transparent Electrodes ................................................................................... 201 xii Figure 1-1 Figure 2-1 Figure 2-2 Figure 2-3 Figure 2-4 Figure 2-5 Figure 2-6 Figure 2-7 Figure 3-1 Figure 3-2 Figure 4-1 Figure 4-2 LIST OF FIGURES Spectroelectrochemical Techniques Using Absorption Spectroscopies ................................................................................ 9 Conventional 3-Electrode Glass Electrochemical Cell ................... 43 Optically Transparent Thin-Layer Electrochemical Cell Design Employing a Au Grid Optically Transparent Electrode ................... 45 UV/vis Spectroelectrochemical Cell Employing a Freestanding Boron-Doped Diamond Optically Transparent Electrode ............... 49 IR Spectroelectrochemical Cell Employing a Boron-Doped Diamond Optically Transparent Electrode -— Model I ..................................... 51 IR Spectroelectrochemical Cell Employing a Boron-Doped Diamond Optically Transparent Electrode - Model II .................................... 53 Temperature-Control Set Up ......................................................... 55 Cryostat Box Diagrams .................................................................. 56 Simulated Cyclic Voltammogram — Diffusion-Limited Case ........... 63 Ferrocene Cyclic Voltammetric Response at a Boron-Doped Diamond Electrode ........................................................................ 66 Cytochrome c Heme Redox Center ............................................... 78 Cyclic Voltammetry of Cytochrome c at “As Deposited” 1 ppm and 10 ppm Microcrystalline Boron-Doped Diamond Electrodes ......... 83 xiii Figure 4-3 Figure 4-4 Figure 4-5 Figure 4-6 Figure 4-7 Figure 4-8 Figure 4-9 Figure 4-10 Figure 4-11 Figure 4-12 Figure 4-13 Cyclic Voltammetry of Cytochrome c at “As Deposited” Microcrystalline Boron-Doped Diamond Electrodes ...................... 84 Cyclic Voltammetry of Cytochrome c at “Acid-Washed, Rehydrogenated” Microcrystalline Boron-Doped Diamond Electrodes ..................................................................................... 85 Demonstration of the Cyclic Voltammetric Response Stability for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode ....................................................................................... 86 Cyclic Voltammetric Response for Cytochrome 0 Measured at a Microcrystalline Boron-Doped Diamond Electrode ........................ 88 Peak Current Dependence on the Square Root of the Scan Rate Measured for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode ........................................................................ 90 Peak Current Dependence on the Concentration Measured for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode ....................................................................................... 91 Cyclic Voltammetric Background Curves Measured for a Boron- Doped Diamond Electrode after Exposure to Cytochrome c ......... 92 Cyclic Voltammetry of Cytochrome c at “As Deposited" 1 ppm and 10 ppm Nanocrystalline Boron-Doped Diamond Electrodes .......... 94 Cyclic Voltammetry of Cytochrome c at “As Deposited” Nanocrystalline Boron-Doped Diamond Electrodes ...................... 95 Cyclic Voltammetry of Cytochrome c at “Acid-Washed, Rehydrogenated" Nanocrystalline Boron-Doped Diamond Electrodes ..................................................................................... 96 Demonstration of the Cyclic Voltammetric Response Stability for Cytochrome c at a Nanocrystalline Boron-Doped Diamond Electrode ....................................................................................... 97 xiv Figure 4-14 Figure 4-15 Figure 4-16 Figure 5-1 Figure 5-2 Figure 5-3 Figure 5-4 Figure 6-1 Figure 6-2 Figure 6-3 Figure 6-4 Figure 6-5 Figure 6-6 Cyclic Voltammetric Response for Cytochrome 0 Measured at a Nanocrystalline Boron-Doped Diamond Electrode ........................ 98 Reduction Peak Current Dependence on the Square Root of the Scan Rate Measured for Cytochrome c at a Nanocrystalline Boron- Doped Diamond Electrode .......................................................... 100 Reduction Peak Current Dependence on the Concentration Measured for Cytochrome c at a Nanocrystalline Boron-Doped Diamond Electrode ...................................................................... 101 Infrared Absorption Spectra of H20 and D20 .............................. 123 Comparison of the H20 Absorption Spectrum and the Noise Spectrum for an Aqueous Sample ............................................... 124 Theoretical Construction of an Electrochemical Difference FTIR Spectrum ..................................................................................... 129 Effect of Temperature Drift on Baseline Linearity ........................ 139 Cytochrome c Oxidase from Rhodobacter Sphaeroides .............. 148 UV/vis Electrochemical Difference Spectra of Rhodobacter Sphaeroides Cytochrome c Oxidase ........................................... 153 FTIR Electrochemical Difference Spectra of Rhodobacter Sphaeroides Cytochrome c Oxidase ........................................... 156 Illustration of Characteristic Mid-IR Bands ................................... 158 Comparison of the FTIR Difference Spectra of Bacterial and Mammalian Cytochrome c Oxidases ........................................... 161 Comparison of the FTIR Difference Spectra of Cytochrome c Oxidase Measured in D20 and H20 ............................................ 167 XV Figure 6-7 Figure 6-8 Figure 6-9 Figure 6-10 Figure 6-11 Figure 7-1 Figure 7-2 Figure 7-3 Figure 7-4 Figure 7-5 Figure 7-6 Figure 7—7 Figure 7-8 Effect of Global 15N Labeling on Mid-IR FTIR Spectra of Cytochrome c Oxidase ................................................................ 168 Isotope Exchange Effects in the Amide II Region ........................ 170 Deuterium Exchange Effects in the Carbonyl Region .................. 174 Closer Examination of a Possible Histidine Mode ....................... 177 Structures of the Two Neutral Histidine Tautomers ..................... 179 Simulated Cyclic Voltammogram - Thin-Layer Case ................... 204 UV/vis Transmission Spectra of Freestanding Boron-Doped and CVD White Diamond Disks: Effect of Boron Doping Level .......... 208 UV/vis Transmission Spectra of Optically Transparent Electrodes: Boron-Doped Diamond compared to Indium-Doped Tin Oxide ...209 Ferrocene Cyclic Voltammetric Response Measured in the UV/vis Optically Transparent Thin-Layer Electrochemical Cell ............... 211 UV/vis Spectroelectrochemical Absorbance Spectra of Ferrocene Oxidation at a Freestanding Boron-Doped Diamond Optically Transparent Electrode ................................................................. 213 Nernst Plot for the UV/vis Spectroelectrochemistry of Ferrocene at a Freestanding Boron-Doped Diamond Optically Transparent Electrode ..................................................................................... 215 Absorbance-Potential (Voltabsorptometry) Profiles of Ferrocene UV/vis Spectroelectrochemistry at a Freestanding Boron-Doped Diamond Optically Transparent Electrode ................................... 216 Concentration Dependence for the Electrogenerated Ferricenium UV/vis Absorbance Peaks ........................................................... 218 xvi Figure 7-9 Figure 7-10 Figure 7-11 Figure 7-12 Figure 7-13 Figure 7-14 Figure 7-15 Figure 7-16 Figure 7-17 UV/vis Spectroelectrochemical Absorbance Spectra of Cytochrome c at a Freestanding Boron-Doped Diamond Optically Transparent Electrode ................................................................. 220 Direct Electrochemical Redox Titration of Cytochrome c at a Freestanding Boron-Doped Diamond Optically Transparent Electrode ..................................................................................... 222 The Effect of Boron Doping Level on the IR Transmission of Diamond Optically Transparent Electrodes ................................. 223 IR Transmission Spectra of Different Types of Boron-Doped Diamond Optically Transparent Electrodes ................................. 226 IR Spectroelectrochemical Absorbance Spectra of Ferrocene/Ferricenium Measured at a Boron-Doped Diamond Optically Transparent Electrode — Mid-IR .................................... 229 IR Spectroelectrochemical Absorbance Spectra of Ferrocene/Ferricenium Measured at a Boron-Doped Diamond Optically Transparent Electrode - Far-IR .................................... 230 Ferrocyanide Cyclic Voltammetric Response Measured in the IR Optically Transparent Thin-Layer Electrochemical Cell (Model II) .................................................................................................... 231 IR Spectroelectrochemical Absorbance Spectra of Ferri-lFerrocyanide Measured at a Boron-Doped Diamond Optically Transparent Electrode ................................................................. 233 Comparison of the Mid-IR Spectra of Cytochrome 0 Measured at a Au Grid and at a Boron-Doped Diamond Optically Transparent Electrode ..................................................................................... 235 xvii CHAPTER 1 1 Introduction 1.1 Insight into Metalloprotein Mechanisms Metals in biology are divided into two classes. The first class consists of highly concentrated ions, such as K, Na”, M92+, and Ca2+, which are important in maintaining the structure of proteins and in regulating cell membrane function. The second class consists of the so-called metalloproteins, in which low levels of ionic forms of Mn, Fe, Co, Cu, Zn, Mo, and so on, are incorporated Into the protein. Depending on function, metalloproteins are further divided into two categories: (A) transport and storage proteins and (B) enzymes. Type A includes Fe-containing oxygen-transport proteins such as hemoglobin, myoglobin, hemerythrin, and the Cu-centered hemocyanine; electron-transfer proteins such as cytochromes (Fe), iron-sulfur proteins (Fe), and blue-copper proteins (Cu); and metal storage proteins such as ferritin (Fe) and ceruloplasmin (Cu). The enzyme category, type B, includes hydrolases such as carboxypeptidase (Zn) and aminopeptidase (Zn, Mg); oxidoreductases such as the heme-copper oxidases (Fe, Cu, Mg, Mo), photosynthetic enzymes (Mn, Ca), and nitrogenase (Mo, Fe); and isomerases such as vitamin B12 coenzyme (Co). Understanding the role Of metal ions in metalloproteins first requires knowledge of the coordination chemistry (structure and bonding) at the active sites. Active sites are the discrete regions of a protein where the chemistry occurs. Detailed structural information is difficult to obtain because these sites are usually buried amid a complex protein structure. X-ray crystallography would be the method of choice for this purpose; however, its application is often hampered by the difficulties in growing single crystals of high molecular weight protein molecules and in analyzing diffraction data with high resolution. For a number of proteins, these difficulties have been overcome, and knowledge of precise geometries from x-ray crystallographic data has made a great contribution to our understanding of their biological functions [1, 2]. The crystallization of membrane proteins appears to be the exception, as little x-ray structural information is available or definitive [3-5]. In such cases, complementary information about the structure, bonding and environment of the metal is acquired via a variety of physiochemical measurement techniques. These include electronic, infrared (IR), resonance Raman, magnetic, nuclear, circular dichroism, and Mossbauer spectroscopies, and electrochemical, thermodynamic and kinetic measurements. Information on the interaction of participating molecules is required to understand the molecular mechanisms of enzymes. Infrared spectroscopy has been essential in addressing such problems [6, 7]. Often, only a small part of the large enzyme actually participates in the catalytic reactions. Infrared difference spectra, which result as the difference of two unique states of the enzyme, report only on the modes that undergo a change upon the transition. For small proteins, direct subtraction of the absorbance spectra of the two states, or analysis by band-narrowing techniques [8-10] are successful approaches for obtaining difference spectra. Alternatively, investigation of larger, more complex proteins requires in situ methods, whereby a trigger initiates the transition from one state of the protein to the next, directly in the infrared cell. When combined with sensitive instrumentation, this approach allows for the accurate detection of the subtle (0.1% of the total absorbance) changes in the vibrational spectrum that are associated with enzyme catalysis [6]. Reaction-induced difference spectroscopy is the term given to such a method, in which a trigger is used to advance a molecule from one state to another in an effort to obtain a difference spectrum of the unique states. There are several options for performing reaction-induced difference spectroscopy, including indirect [11, 12] and direct photochemical reactions [13, 14], and electron transfer reactions [15]. Metalloproteins are well suited for the electrochemical difference method, as many contain redox-active metal centers. One of the advantages of using electrochemistry as the trigger is its generality of application, as there is no requirement for highly absorbing chromophores or for a light-initiated reaction cycle—only the need for a redox-active metal. For example, this method has proven useful for studies of the proteins involved in photosynthesis [16] and respiration [17-19]. 1.2 Spectroelectrochemical Methods The combination of electrochemistry and spectroscopy—so-called spectroelectrochemistry—has been in use for almost forty years. Kuwana’s pioneering approach coupled UV/vis spectroscopy with an electrochemical cell to monitor solution concentrations of dissolved redox species [20]. These results were revolutionary, in that, by this method, it was possible to obtain time- and potential-dependent spectra of reaction components from which one could elucidate reaction mechanisms and kinetic information. Without the additional, optical probe, previous experiments of this type were much more difficult. Over the next several years, Kuwana’s group [21-25], and others [26-38], continued to report on UV/vis Spectroelectrochemical studies of solution species, including those of proteins [35, 39-41]. In the mid 1960s, there was a shift in the experimental interest to include investigations of surface processes occurring at the electrode. UV/vis radiation was not useful for probing the details of such surface phenomena; therefore, some focus turned to coupling infrared (IR) spectroscopy with electrochemistry. It was this shift in experimental interest, accompanied by the commercial development and availability of Fourier transform infrared (FTIR) spectrophotometers that spurred the development of IR spectroelectrochemistry. The first reports on the use of internal reflection IR spectroscopy to monitor the spectrum of a species at the electrode surface, during electrolysis, were published in 1966 [42, 43]. At the outset, IR techniques were primarily focused on electrode surface characterization [42-47]. There were few reports of the application of IR to solution studies, due primarily to the large interference of solvent absorption with the absorption by the analyte(s) of interest. However, in that it provided a high level of detail regarding the molecular structure of reaction components, IR offered an advantage over other spectroscopies that had been coupled to electrochemistry (e.g., UV/vis, electron spin resonance, and Mossbauer) for solution studies. Throughout the next two decades, the availability of higher sensitivity instrumentation and novel spectroelectrochemical cell designs resulted in a steady increase in the number of reports of solution IR spectroelectrochemical studies in the fields of organic and inorganic chemistry [48], and resulted in a number of general review articles on the subject [6, 14, 49- 53]. Today, the focus of spectroelectrochemical studies has expanded to include the development of new types of optically transparent electrodes (OTES) [54, 55], the incorporation of OTEs as sensors [56, 57], and the investigations of intricate biological systems [6, 14, 58-63]. In particular, recent advances in solution IR Spectroelectrochemistry include applications that are capable of elucidating single amino acid involvement in complex enzyme mechanisms [15, 64]. Examples of such experiments on the metalloprotein, cytochrome c oxidase (cco), are described in this dissertation. The catalytic mechanism of cco involves an efficient reduction of molecular oxygen to water, concomitant with the uptake of eight protons, of which four are used in the chemistry and four are translocated across the membrane. The way in which cco links the complex oxygen chemistry at different stages to the proton pumping function, in order to drive protons about 50 A across the mitochondrial membrane against an electrochemical gradient, is not understood. Insights into amino acid involvement in this process can be gained via IR spectroelectrochemistry. Spectroelectrochemistry combines two very different techniques (i.e., spectroscopy and electrochemistry); and because of this, the method provides more information about the redox mechanism of electroactive analytes than can be acquired using traditional electrochemical methods alone. Oxidation states are changed electrochemically, by adding or removing electrons at an OTE. The spectroscopic response of the species in solution is measured simultaneously with the electrolysis. Spectroelectrochemical methods of analysis have been used for over three decades to investigate heterogeneous electron transfer processes. Spectroelectrochemistry is useful for obtaining spectra and redox potentials for electrogenerated species in the investigation of electrode reaction mechanisms, as well as in electroanalytical applications. In particular, we are interested in using Spectroelectrochemistry to elucidate structural information about redox-active protein active sites, and, specifically, that of cytochrome c oxidase. Our goals are to be able to routinely measure high-quality electrochemical difference FTIR spectra of cytochrome c oxidase, and then to establish a protocol for performing such measurements in the low-frequency infrared region of the electromagnetic spectrum. Although mid-IR (2500 - 1000 cm°1) measurements have contributed substantially to our understanding of amino acid involvement in catalysis, accessing the low- frequency region (< 1000 cm'1) is also crucial, as it is this region that reports directly on metal-ligand modes of protein active sites. In order to directly monitor substrate binding and intermediate structure via metal center coordination, low- frequency FTIR difference measurements of proteins are necessary. However, many technical difficulties, including low Signal-to-background ratios and limited options of suitable infrared windows, have plagued such measurements, These problems and corresponding solutions are discussed in detail, in a later part of this work. We set out to accomplish our goal in two phases: First, we implemented a conventional transmission method for acquiring mid-IR spectra of redox-active proteins, developed by Mantele and coworkers [65, 66]. This approach uses a Au minigrid OTE in a spectroelectrochemical cell that was specially designed to meet the technical challenges inherent in such measurements. Preliminary results were obtained for a simple electron transfer protein, cytochrome 0, upon which we quickly adapted our setup to handle the more complex enzyme cytochrome c oxidase. The second phase involved efforts to access the low-frequency IR region by modifying this experimental set- up to improve the signal-to-noise ratio of such measurements. These changes included a highly sensitive low-frequency detection scheme (Si bolometer), a stable background signal achieved via precise temperature control, and maximized optical throughput using diamond optical windows. However, despite these efforts, low-frequency measurements were unsuccessful for the modified set-up. Additionally, as part of such efforts, we have begun developing a new type of OTE, boron-doped diamond (BDD), for spectroelectrochemical measurements of proteins, again using cytochrome c as a test system. Part of the development of this novel OTE material included electrochemical characterization in nonaqueous solvents, in preparation for studies of bioinorganic compounds that mimic the structure of protein active sites and, therefore, serve as simple model systems. The model compound studies are not discussed in this dissertation. The electrochemical characterization of cytochrome c was also carried out at this relatively new electrode material. This was important for understanding the nature of the interaction and electron transfer kinetics of the protein with the BDD surface. Besides basic electrochemical characterizations, we aimed to determine the deposition conditions that lead to an optimized balance between electrical conductivity and optical transparency in this material. This dissertation reports on the application of BDD OTES for spectroelectrochemical studies of two test analytes, ferrocyanide and ferrocene, in the UV/vis and IR regions of the electromagnetic spectrum. Spectroelectrochemical investigations were also made of cytochrome c, which enabled us to make a direct comparison of the results obtained for BDD OTES to those obtained with a traditional Au grid OTE in the UV/vis and IR regions of the spectrum. A variety of optical methods have been coupled with electrochemistry, including absorption, as is shown in Figure 1-1, as well as light scattering spectroscopies. Transmission UV/vis spectroscopy is perhaps the most commonly employed method, owing to the ease of cell design and measurement. In the transmission mode (Figure 1-1A, B), the optical beam is directed normal to the surface and passes directly through an OTE and the adjacent solution. Infrared spectroscopy, in both the transmission and reflectance modes, has also enjoyed widespread use in spectroelectrochemistry [50, 51, 53]. The specular reflectance mode (Figure 1-1C) involves passing IR radiation through the solution onto a reflective surface (electrode), at which point the light is reflected back through the solution to the detector. In attenuated total reflectance (ATR) infrared spectroscopy, the probe beam of light is introduced into a highly refractive material, at an angle greater than the critical angle so that the beam is totally internally reflected (Figure 1-1D). Light attenuation occurs through interaction between the evanescent wave of the totally reflected light and the light-absorbing solution species adjacent to the electrode surface. Figure 1-1 Spectroelectrochemical Techniques Using Absorption Spectroscopies Methods employing UV, visible, and IR Spectroscopies can be divided into two classes, those which measure transmission (A, B), and those measuring reflectance (C, D). (A) In the thin-layer orientation, the path/ength is less than or equal to 0.2 mm. (B) Conventional cell orientation, where the solution thickness (~ 1 cm), is greater than the diffusion-layer thickness. (C) Specular reflectance at a non-transparent electrode. (D) lntemal reflectance (commonly called A TR) method illustrating the evanescent wave. Figure adapted from [49]. l——— UV, VISIBLE, IR fié’l'RANSMlSSIONii HIREFLEc'rANca-iw 0.2 mm OTE\ HK electrode A. Thin-layer cell C. External reflectance 1 cm OTE '—"'I \[Y ,, B. Conventional cell D. Internal reflectance An assortment of spectroelectrochemical cell types are reported in the literature [25, 28, 34, 37, 42, 54, 66-84]. Although it is possible, and sometimes necessary, to use a cell geometry analogous to a conventional electrochemical cell (Figure 1-1B), in that the solution thickness is much greater than the diffusion-layer thickness normal to the electrode, an optically transparent thin- layer electrochemical (OTTLE) cell design (Figure 1-1A) is most commonly employed. One of the main advantages of this cell type is that the electroactive species in the thin layer can be completely electrolyzed in a very short time (typically 20 - 120 S). Additionally, diffusion effects are eliminated, simplifying the mathematical descriptions of the electrochemical parameters. Often, in the absorption spectroscopy mode, experimental requirements impose functional limits on the pathlength or thin-layer distance. An example of such a situation will be presented later in this dissertation, in which the difficulties of low-frequency infrared measurements of proteins will be described. 1.3 Optically Transparent Electrodes 1.3.1 Conventional Types Several types of optically transparent electrodes (OTES) are available for use in spectroelectrochemical measurements. The conventional types include thin-films of a conductive material on transparent substrates, and minigrid electrodes. Thin-films of platinum, gold, indium-doped tin oxide (ITO), or carbon have been deposited on glass (visible), quartz (UV-visible), or germanium and silicon (infrared), depending on the region of interest. The transparency (20 - 85%) of the thin-film OTE depends on several properties, including the film thickness. The second type of conventional OTE, the minigrid, consists of a 10 metal (Ni, Pt, Au, Ag or Cu) micromesh of 100 - 2000 wires per inch. In this case, the transparency (20 - 80%) is due to the openings in the mesh. Minigrid OTES have been used primarily in transmission experiments employing thin-layer cell designs [32, 66, 78-80, 84-87]. An example of such a design for UV/vis and IR transmission experiments of redox-active proteins is one of the OTTLE cells that will be presented later, in Chapter 2. 1.3.2 A New Material: Boron-Doped Diamond Much of the work detailed herein involves the development and characterization of a new material for use as an OTE—boron-doped diamond. Although the application of conductive diamond films as electrode materials is a relatively new field, with its beginnings in the early 1990s [88-92], a sound knowledge base has been developed in recent years regarding structure-function relationships between redox analytes and the electrode surface [93-97]. From electrochemical studies in aqueous media [94, 98-100], it is clear that high- quality (i.e., hydrogen-terminated, with minimal nondiamond (spZ-bonded) carbon impurity content, and low secondary nucleation density) BDD thin-film electrodes exhibit a number of advantageous electrochemical properties, including: (i) low and stable background currents and capacitance; (ii) a wide working potential window (3 - 4 V); (iii) rapid electron transfer kinetics for many redox systems without the requirement of conventional pretreatment procedures; (iv) long-term response stability; (v) morphological and microstructural stability during anodic and cathodic polarization (~10 A/cmz) in acidic and alkaline media; and (vi) high 11 resistance to fouling, due to the generally weak molecular adsorption of polar molecules. Experimental focus is currently centered on two types of electrically conducting, chemically vapor deposited (CVD) materials: microcrystalline and nanocrystalline diamond thin-films. The details of BDD film growth are discussed in Chapter 2. Generally, microcrystalline films are grown using a CH4/H2 (~0.5%) source gas mixture. Under these growth conditions, the rate of crystal growth generally exceeds the rate of nucleation, producing large (1 -20 pm), well- faceted diamond grains. The nanocrystalline films are grown using a CH4/Ar (~1°/o) source gas mixture with little or no H2 added [101]. Under these growth conditions, the rate of nucleation generally exceeds the rate of crystal growth, resulting in the formation of small grains (~15 nm). Such films are very smooth, with a 30-50 nm surface roughness, and have a higher fraction of grain boundaries than do the microcrystalline films, due to the smaller grain size [95, 102]. The contact angle measured for water (740°) indicates a hydrophobic surface, and x-ray photoelectron spectrosc0pic data indicates surface oxygen levels of less than 2 atomic percent. lmparting boron-dopant atoms during film deposition renders the normally insulating diamond material electrically conductive. Film conductivity is, therefore, directly dependent upon the level of boron doping (i.e., electron carrier concentration), with the most conductive films doped at levels greater than 1020 B/cm3, yielding resistivities less than 0.05 0 cm. As mentioned above, most of the research with diamond electrodes has involved the use of aqueous redox analytes, with very few studies performed using nonaqueous redox analytes [97, 103-106]. There are even fewer reports of 12 the application of BDD electrodes toward biological electron transfer reactions [107, 108]. Both of these applications are described in this work. Much of the literature on the use of diamond in electrochemistry is in the field of electroanalysis, as, during the past few years, numerous reports have been published on the application of this material in chemical measurements [54, 99, 100, 105-157]. The success of diamond in such studies is attributed to the significant advantages it offers over other carbon and metal electrodes in terms of stability, limit of detection, robustness, and dynamic range [99, 100, 105, 106, 153, 154, 156, 157]. Another unique property of BDD, which has yet to be exploited, is the optical throughput. As an optical material, diamond is mechanically strong, resistant to chemical attack, and, most importantly, it has the widest optical window of any material, extending from the band gap absorption edge in the visible (~ 225 nm) deep into the far infrared (~ 30 cm'1). Although imperfections and impurities in the diamond lattice, such as boron, decrease the Optical throughput, BDD retains a wide optical window transmitting (> 50%) in the visible (300 - 900 nm) and infrared (6000 - 2600 and < 1500 cm'1) regions. Factors that influence the optical throughput of diamond include the defect density, chemical composition, dOping level, film thickness, and grain size [158-176]. Thus, it is possible to manipulate and optimize the Optical properties of this material through adjustments in the CVD growth conditions. Knowledge of these growth parameters is essential for establishing an optimized balance between the electrochemical and optical properties of BDD OTES. 13 A wide optical window, coupled with attractive electrochemical properties makes BDD a viable OTE material. The application of electrically conducting diamond for this purpose is in its infancy [54, 104, 106, 177, 178]. Boron-doped diamond offers several advantages over other traditionally used OTES, such as indium-doped tin oxide (ITO) including: (i) transparency extending from the UV into the far IR, (ii) remarkable stability in aqueous (acidic and alkaline) and nonaqueous (e.g., chlorinated) media, (iii) a wide working potential window in aqueous and nonaqueous media, (iv) a reasonably well-defined and stable surface chemistry, and (v) the ability to withstand cathodic polarization. Additionally, aside from thin-films coated onto a transparent substrate, BDD exists in a freestanding form. The ability to manufacture BDD OTES based on deposited thin-films, as well as in freestanding forms, offers the option for application specificity. Our group's research with the optically transparent BDD electrodes is in the early stages; therefore, we have only begun to test the extent of applications for this material. For instance, our preliminary spectroelectrochemical studies of cytochrome 0 show great promise for future applications of BDD OTES in the field of metalloprotein spectroelectrochemistry. 1.4 Scientific Relevance Spectroelectrochemical studies of metalloproteins have attracted widespread interest and attention. Such studies yield important information about not only the intrinsic thermodynamic and kinetic properties of the redox- active biomolecules, but also the structural properties, which are important for understanding the molecular mechanisms of enzyme function. Beyond the basic 14 research goals to elucidate mechanistic insights into metalloprotein catalysis, applications of a more applied nature do exist. These include the development of spectroelectrochemical biosensors based on redox-active metalloproteins. Currently, such methods involve sol-gel encapsulation of metalloproteins with optical detection schemes [179-185]. An alternate approach, in which both optical and electrochemical sensing are possible, involves the immobilization of protein molecules on an OTE. Preliminary studies of this type have been conducted at nanoporous titanium dioxide (T102) for cytochrome c and hemoglobin [186]. Hemoglobin is of particular interest for optical biosensing methods due to its capability for high-affinity binding of Oz, CO, and NO, which drastically change its absorption spectrum when bound to the heme iron. However, hemoglobin only binds these molecules in the Fe(|l) oxidized state, which, incidentally, is not its resting state (Fe(lll)). Sol-gel methods, therefore, require repeated additions of chemical oxidants and reductants to cycle this enzyme between the Fe(ll) and Fe(lll) states [180, 185]. The ability to perform electrochemistry directly at a transparent electrode would alleviate this problem, which complicates technological development of such sensors. Topoglidis and coworkers have demonstrated the viability of such an approach for sensing CO in aqueous solution by optical detection at a hemoglobin/I102 film [186]. BDD has Shown great promise as an optically transparent electrode and in its application for direct electrochemistry of cytochrome 0. Therefore, it is conceivable to expect that this 15 material may be exploited in the future developments of spectroelectrochemical sensing using biomolecules. 1.5 Dissertation Outline The following chapter, Chapter 2, of this dissertation is referred to throughout as the experimental section, as it details all experimental parameters and procedures used for the measurements discussed in later chapters. Chapters 3 and 4 demonstrate the usefulness of BDD as an electrode material for investigating electron transfer kinetics of two very different systems. Chapter 3 reports on the first system, ferrocene, which is routinely used as a model system in nonaqueous kinetic studies because it demonstrates “ideal” electrochemical behavior in a variety of solvents. Cyclic voltammetry was used to study the ferrocene/ferricenium redox reaction at microcrystalline BDD films as a function of the scan rate and the electrolyte/solvent composition. The response at BDD was compared to a well-characterized carbon electrode, glassy carbon. This study demonstrated that the electrode microstructure does not significantly affect the rate of ferrocene oxidation at these two carbon electrodes, and that BDD is a viable electrode for performing electrochemistry in nonaqueous solvents. In Chapter 4, a more challenging system is investigated by cyclic voltammetry at BDD electrodes. This system, cytochrome c, is a well-studied, electron transfer protein. Until recently, it was believed that, in order to observe facile electron transfer of cytochrome c at a bare electrode, the surface must be terminated with negatively charged, hydrophilic, oxygen groups. The effects of protein concentration, boron-doping level, and microstructure were investigated. 16 This study demonstrated that quasi-reversible, stable electron transfer could be performed at hydrophobic, hydrogen-terminated, BDD electrodes, thus raising important questions of the conventional wisdom regarding the necessary surface structure for cytochrome c electron transfer. In Chapters 5, 6, and 7, the focus turns to spectroelectrochemistry. The technical difficulties associated with performing low-frequency FTIR spectroscopy of biological samples are discussed in Chapter 5, and the concept of reaction- induced infrared difference spectroscopy is described in detail. Chapter 6 discusses applications of electrochemical difference infrared spectroscopy to the metalloprotein cytochrome c oxidase. This technique has been extended to its technical limits, in an effort to elucidate structural and mechanistic aspects of oxygen reduction to water by cco. Isotope-labeling experiments of the protein samples were performed in an effort to simplify mode assignments. Additionally, cco from two different organisms, a bacterial and a mammalian source, were compared. Although the oxidase structure and mechanistic function is highly conserved between these two organisms, there were observable deviations in the IR spectra. These experiments identified a potential histidine amino acid side chain that is linked to the electron transfer reaction of cco and may be important in the proton pumping mechanism. Questions remain regarding whether this histidine is a ligand to the Cue of the active site. These studies are representative of some of the most challenging FTIR experiments currently being conducted, as they are attempting to identify single amino acid residues (AA ~ 10"4 AD.) in complex protein FTIR spectra. 17 The last chapter, Chapter 7, describes BDD OTE spectroelectrochemical studies in the UV/vis and IR regions of the spectrum. The examination of the optical properties of BDD is in its infancy. Of the several types of BDD OTES available, freestanding disks were used for studies in the UV/vis region and thin- films of BDD on undoped silicon substrates for the IR. Novel spectroelectrochemical cells were designed to utilize these two types of OTES. The cell designs were tested with ferrocyanide, ferrocene, and cytochrome c, and key insights into the relationship between the diamond growth conditions and OTE optical properties emerged. For the first time, the spectroelectrochemical response for cytochrome c was measured at BDD OTES. The observed UV/vis and IR spectra were compared to those obtained using a conventional OTE, a Au minigrid. This study indicated the viability of using BDD thin-films as OTES but also demonstrated that further work is required to optimize the growth conditions to yield a BDD OTE that possessed both the conductivity and transparency to compete with 3 Au minigrid OTE. Subsequent studies may determine that a different type of BDD OTE, one with better conductivity and optical transparency, is better suited for this purpose. However, these studies of the protein system are among the most advanced experiments currently being conducted with BDD OTES. The ability to combine direct electrochemistry of biomolecules with UV/vis and IR spectroscopies on one electrode may have further implications in the applied fields of bioanalytical devices and enzyme-based spectroelectrochemical sensors. Such applications could benefit from the resistance to fouling and extreme potential ranges that are possible with BDD OTES. 18 1.6 [1] [2-1 [3-1 [4-1 [51 [6.] [7-1 [8-1 [91 [10.] [11.] [12.] [13.] References L. Esposito, L. Vitagliano and L. Mazzarella Protein Pept. Left. 2002, 9, 95-105. C. M. Wilmot and A. R. Pearson Curr. Opin. Chem. Biol. 2002, 6, 202-207. A. Zouni, H. T. Witt, J. Kern, P. Fromme, N. Krauss, W. Saenger and P. Orth Nature 2001, 409, 739-743. S. Yoshikawa, K. Shinzawa-Itoh and T. Tsukihara J. Bioenerg. Biomembr. 1998, 30, 7-14. T. Tsukihara, H. Aoyama, E. Yamashita, T. Tomizaki, H. Yamaguchi, K. Shinzawaltoh, R. Nakashima, R. Yaono and S. Yoshikawa Science 1996, 272, 1136-1144. C. Zscherp and A. Barth Biochemistry 2001, 40, 1875—1883. J. Heberle and T. Gensch Nat. Struct. Biol. 2001, 8, 195-197. A. Barth Spectrochim. Acta A 2000, 56A, 1223-1232. J. K. Kauppinen, D. J. Moffatt, H. H. Mantsch and D. G. Cameron Appl. Spectrosc. 1981, 35, 271-276. H. H. Mantsch, D. J. Moffatt and H. L. Casal J. Mol. Struct. 1988, 173, 285-298. P. R. Rich and J. Breton Biochemistry 2001, 40, 6441-6449. A. Barth and C. Zscherp FEBS Lett. 2000, 477, 151-156. H. A. Chu, W. Hillier, N. A. Law, H. Sackett, S. Haymond and G. T. Babcock Biochim. Biophys. Acta 2000, 1459, 528-532. 19 [14.] [15.] [16.] [17.] [18.] [19.] [20.] [21 .1 [22.] [23.] [24.] [25.] [26.] W. Mantele In Biophysical Techniques in Photosynthesis; J. Amesz and A. J. Hoff, Eds.; Kluwer Academic Publishers, Dordrecht, 1996; Vol. 3, 137- 160. P. Hellwig, U. Pfitzner, J. Behr, B. Rost, R. P. Pesavento, W. von Donk, R. B. Gennis, H. Michel, B. Ludwig and W. Mantele Biochemistry 2002, 41, 9116-9125. W. Mantele, A. Wollenweber, F. Rashwan, J. Heinze, E. Nabedryk, G. Berger and J. Breton Photochem. Photobiol. 1988, 47, 451-455. J. Zhang, W. Oettmeier, R. B. Gennis and P. Hellwig Biochemistry 2002, 41, 4612-4617. B. Barquera, P. Hellwig, W. Zhou, J. E. Morgan, C. C. Haese, K. K. Gosink, M. Nilges, P. J. Bruesehoff, A. Roth, C. R. D. Lancaster and R. B. Gennis Biochemistry 2002, 41 , 3781-3789. P. Hellwig, B. Rost and W. Mantele Spectrochim. Acta A 2001, 57A, 1123- 1 131. T. Kuwana, R. K. Darlington and D. W. Leedy Anal. Chem. 1964, 36, 2023-2025. T. Kuwana and W. R. Heineman Acct. Chem. Res. 1976, 9, 241-248. J. Zak, M. D. Porter and T. Kuwana Anal. Chem. 1983, 55, 2219-2222. T. Kuwana and N. Winograd In Electroanalytical Chemistry: A Series of Advances; A. J. Bard, Ed.; Marcel Dekker, Inc., New York, 1974; Vol. 7, 1- 78. T. Osa and T. Kuwana J. Electroanal. Chem. 1969, 22, 389-406. J. W. Strojek and T. Kuwana J. Electroanal. Chem. 1968, 16, 471—483. W. R. Heineman, T. P. DeAngeIis and J. F. Goelz Anal. Chem. 1975, 47, 1364-1369. 20 [27.] [28.] [29.] [30.] [31.] [32.] [33.] I341 [35.] [36.] [37.] [38.] [39.] [40.] N. R. Armstrong, R. K. Quinn and N. E. Vanderborgh J. Electrochem. Soc. 1976, 123, 646-649. F. M. Hawkridge, J. E. Pemberton and H. N. Blount Anal. Chem. 1977, 49, 1646-1647. D. E. Albertson, H. N. Blount and F. M. Hawkridge Anal. Chem. 1979, 51, 556-560. J. L. Anderson Anal. Chem. 1979, 51, 2312-2315. E. A. Blubaugh, A. M. Yacynych and W. R. Heineman Anal. Chem. 1979, 51, 561-565. R. K. Rhodes and K. M. Kadish Anal. Chem. 1981, 53, 1539-1541. E. E. Bancroft, H. N. Blount and F. M. Hawkridge Adv. Chem. Ser. 1982, 23-49. E. F. Bowden, D. J. Cohen and F. M. Hawkridge Anal. Chem. 1982, 54, 1005-1008. K. A. Rubinson and H. B. Mark, Jr. Anal. Chem. 1982, 54, 1204-1206. H. O. Finklea, R. K. Boggess, J. W. Trogdon and F. A. Schultz Anal. Chem. 1983, 55, 1177-1179. X. Q. Lin and K. M. Kadish Anal. Chem. 1985, 57, 1498-1501. D. A. Scherson, S. Sarangapani and F. L. Urbach Anal. Chem. 1985, 57, 1501-1503. L. N. Mackey and T. Kuwana Bioe/ectrochem. Bioenerg. 1976, 3, 596-613. F. R. Shu and G. S. Wilson Anal. Chem. 1976, 48, 1676-1679. 21 [41.] [42.] [43.] [44.] [45.] [46.] [47.] [48.] [49.] [50.] [51.] [52.] [53.] [54.] [55.] C. W. Anderson, H. B. Halsall and W. R. Heineman Anal. Biochem. 1979, 93, 366-372. H. B. Mark, Jr. and B. S. Pons Anal. Chem. 1966, 38, 119-121. W. N. Hansen, T. Kuwana and R. A. Osteryoung Anal. Chem. 1966, 38, 1810-1821. W. N. Hansen, R. A. Osteryoung and T. Kuwana J. Am. Chem. Soc. 1966, 88, 1062-1063. J. D. E. McIntyre and D. M. Kolb Symp. Faraday Soc. 1970, NO. 4, 99- 113. W. J. Plieth Symp. Faraday Soc. 1970, No. 4, 137-144. H. B. Mark, Jr. and E. N. Randall Symp. Faraday Soc. 1970, No. 4, 157- 172. H. E. Toma and K. Araki Curr. Org. Chem. 2002, 6, 21-34. W. R. Heineman Anal. Chem. 1978, 50, 390A-402A. J. K. Foley and B. S. Pons Anal. Chem. 1985, 57, 945A-956A. K. Ashley Talanta 1991, 38, 1209-1218. F. Siebert In Biochemical Spectroscopy, 1995; Vol. 246, 501-526. J. J. Niu and S. J. Dong Rev. Anal. Chem. 1996, 15, 1-171. J. K. Zak, J. E. Butler and G. M. Swain Anal. Chem. 2001, 73, 908-914. A. Neudeck and L. Kress J. Electroanal. Chem. 1997, 437, 141-156. 22 [56.] [57.] [58.] [59.] [60.] [61.] [62.] [63.] [64.] [65.] [66.] J. N. Richardson, A. L. Dyer, M. L. Stegemiller, l. Zudans, C. J. Seliskar and W. R. Heineman Abstracts of Papers, 223rd ACS National Meeting, Orlando, FL, United States, April 7-11, 2002. M. Maizels, M. Stegemiller, S. Ross, A. Slaterbeck, Y. Shi, T. H. Ridgway, W. R. Heineman, C. J. Seliskar and S. A. Bryan In ACS Symposium Series 2001, Vol. 778, 364-378. J. D. Burgess and F. M. Hawkridge In Electroanalytical Methods for Biological Materials; A. Brajter—Toth and J. O. Chambers, Eds; Marcel Dekker, New York, 2002, 109-142. C. H. Taboy, C. Bonaventura and A. L. Crumbliss In Redox Cell Biology and Genetics, Pt B, 2002; Vol. 353, 187-209. N. B. Ugulava, B. R. Gibney and J. T. Jarrett Biochemistry 2001, 40, 8343- 8351. P. Yao, Y. H. Wang, Y. Xie and Z. X. Huang J. Electroanal. Chem. 1998, 445, 197-201. D. J. Cohen, B. C. King and F. M. Hawkridge J. Electroanal. Chem. 1998, 447, 53-62. S. J. Dong, J. J. Niu and T. M. Cotton In Biochemical Spectroscopy, 1995; Vol. 246, 701-732. P. Hellwig, J. Behr, C. Ostermeier, O.-M. H. Richter, U. Pfitzner, A. Odenwald, B. Ludwig, H. Michel and W. Mantele Biochemistry 1998, 37, 7390-7399. F. Baymann, D. A. Moss and W. Ma‘nteIe Anal. Biochem. 1991, 199, 269- 274. D. Moss, E. Nabedryk, J. Breton and W. Mantele Eur. J. Biochem. 1990, 187, 565-572. 23 [67.] [68.] [69.] [70.] [71.] [72.] [73.] [74.] [75.] [76.] [77.] [78.] [79.] [80.] [81.] P. A. Flowers, M. A. Maynor and D. E. Owens Anal. Chem. 2002, 74, 720- 723. l. S. Zavarine and C. P. Kubiak J. Electroanal. Chem. 2001, 495, 106-109. D. Daniel and I. G. R. Gutz Electroanal. 2001, 13, 681-685. J. 8. Yu, C. Yang and H. O. Fang Anal. Chim. Acta 2000, 420, 45-55. Y. J. Ma, J. M. Zheng and S. M. Zhu Chin. J. Inorg. Chem. 1999, 15, 61- 67. D. Kardash, J. M. Huang and C. Korzeniewski J. Electroanal. Chem. 1999, 476, 95-100. M. E. Rosa-Montanez, H. De Jesus-Cardona and C. R. Cabrera-Martinez Anal. Chem. 1998, 70, 1007-1011. P. A. Flowers and S.-A. Callender Anal. Chem. 1996, 68, 199-202. J. J. Niu and S. J. Dong Electrochim. Acta 1995, 40, 823-828. J. J. Niu, G. J. Cheng and S. J. Dong Chin. Chem. Lett. 1994, 5, 317-320. F. Hartl, H. Luyten, H. A. Nieuwenhuis and G. C. SchoemakerAppl. Spectrosc. 1994, 48, 1522-1528. F. Yang, X. Lin and E. Wang Fenxi Huaxue 1993, 21, 857-860. M. Krejcik, M. Danek and F. Hartl J. Electroanal. Chem. 1991, 317, 179- 187. X. Q. Lin, Y. Liu and E. K. Wang Acta Chim. Sin. 1990, 48, 1080-1084. P. B. Graham In Univ. Massachusetts,Amherst, MA, USA, 1990, p 299 pp. 24 [82.] [83.] [84.] [85.] [86.] [87.] [88.] [89.] [90.] [91.] [92.] [93.] [94.] [95.] X. Lin, D. Liu and E. Wang WuIiHuaxue Xuebao 1989, 5, 719-722. K. M. Kadish, X. H. Mu and X. Q. Lin Electroanal. 1989, 1, 35-41. R. W. Murray, W. R. Heineman and G. W. O'Dom Anal. Chem. 1967, 39, 1666-1668. K. Jiao, L. Q. Ren and T. Yang Chin. J. Anal. Chem. 2001, 29, 1299-1302. J. P. Bullock, D. C. Boyd and K. R. Mann Inorg. Chem. 1987, 26, 3084- 3086. W. R. Heineman, J. N. Burnett and R. W. Murray Anal. Chem. 1968, 40, 1974-1978. M. lwaki, S. Sato, K. Takahashi and H. Sakairi Nucl. Instrum. Methods Phys. Res. 1983, 209-210, 1129-1133. Y. V. Pleskov, A. Y. Sakharova, M. D. Krotova, L. L. Builov and B. V. Spitsyn J. Electroanal. Chem. 1987, 228, 19-27. K. Patel, K. Hashimoto and A. Fujishima J. Photochem. Photobiol., A 1992, 65, 419-429. K. Patel, K. Hashimoto and A. Fujishima DenkiKagaku1992, 60, 659-661. G. M. Swain and R. Ramesham Anal. Chem. 1993, 65, 345-351. M. C. Granger and G. M. Swain J. Electrochem. Soc. 1999, 146, 4551- 4558. M. C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M. D. Koppang, J. E. Butler, G. Lucazeau, M. Mermoux, J. W. Strojek and G. M. Swain Anal. Chem. 2000, 72, 3793-3804. O. Y. Chen, D. M. Gruen, A. R. Krauss, T. D. Corrigan, M. Witek and G. M. Swain J. Electrochem. Soc. 2001, 148, E44-E51. 25 [96.] Q. Chen and G. M. Swain Langmuir1998, 14, 7017-7026. [97.] S. Alehashem, F. Chambers, J. W. Strojek, G. M. Swain and R. Ramesham Anal. Chem. 1995, 67, 2812-2821. [98.] M. C. Granger, J. Xu, J. W. Strojek and G. M. Swain Anal. Chim. Acta 1999, 397, 145-161. [99.] J. S. Xu, M. C. Granger, Q. Y. Chen, J. W. Strojek, T. E. Lister and G. M. Swain Anal. Chem. 1997, 69, A591-A597. [100.] G. M. Swain, A. B. Anderson and J. C. Angus MRS Bulletin 1998, 23, 56- 60. [101.] D. M. Gruen Annu. Rev. Mater. Sci. 1999, 29, 211-259. [102.] B. Fausett, M. C. Granger, M. L. Hupert, J. Wang, G. M. Swain and D. M. Gruen Electroanal. 2000, 12, 7-15. [103.] S. Haymond, G. T. Babcock and G. M. Swain Electroanal. 2002, in press. [104.] S. Haymond, J. K. Zak, J. E. Butler, G. T. Babcock and G. M. Swain in preparation 2002. [105.] G. Pastor-Moreno and D. J. Riley Electrochim. Acta 2002, 47, 2589-2595. [106.] M. Witek, J. Wang, J. Stotter, M. Hupert, S. Haymond, P. Sonthalia, G. M. Swain, J. K. Zak, Q. Chen, D. M. Gruen, J. E. Butler, K. Kobashi and T. Tachibana J. Wide Bandgap Materials 2002, 8, 171-188. [107.] S. Haymond, G. T. Babcock and G. M. Swain J. Am. Chem. Soc. 2002, 124, 10634-10635. [108.] F. Marken, C. A. Paddon and D. Asogan Electrochem. Commun. 2002, 4, 62-66. 26 [109.] N. Wangfuengkanagul and O. Chailapakul J. Phann. Biomed. Anal. 2002, 28, 841-847. [110.] K. Van Hege, M. Verhaege and W. Verstraete Electrochem. Commun. 2002, 4, 296-300. [111.] K. Tsunozaki, Y. Einaga, T. N. Rao and A. Fujishima Chem. Lett. 2002, 502-503. [112.] l. Troster, M. Fryda, D. Herrmann, L. Schafer, W. Hanni, A. Perret, M. Blaschke, A. Kraft and M. Stadelmann Diam. Relat. Mat. 2002, 11, 640- 645. [113.] C. Terashima, T. N. Rao, B. V. Sarada, D. A. Tryk and A. Fujishima Anal. Chem. 2002, 74, 895-902. [114.] N. Spataru, B. V. Sarada, D. A. Tryk and A. Fujishima Electroanal. 2002, 14, 721-728. [115.] D. Sopchak, B. Miller, R. Kalish, Y. Avyigal and X. Shi Electroanal. 2002, 14, 473-478. [116.] T. N. Rao, B. H. Loo, B. V. Sarada, C. Terashima and A. Fujishima Anal. Chem. 2002, 74, 1578-1583. [117.] C. Prado, S. J. Wilkins, F. Marken and R. G. Compton Electroanal. 2002, 14, 262-272. [118.] C. Prado, G.-U. Flechsig, P. Gruendler, J. S. Foord, F. Marken and R. G. Compton Analyst 2002, 127, 329-332. [119.] K. Ohnishi, Y. Einaga, H. Notsu, C. Terashima, T. N. Rao, S.-G. Park and A. F ujishima Electrochem. Solid-State Lett. 2002, 5, D1-D3. [120.] F. Montilla, P. A. Michaud, E. Morallon, J. L. Vazquez and C. Comninellis Electrochim. Acta 2002, 47, 3509-3513. 27 [121.] K. J. McKenzie and F. Marken Electrochem. Solid-State Left. 2002, 5, E47-E50. [122.] A. Manivannan, M. S. Seehra, D. A. Tryk and A. Fujishima Anal. Lett. 2002, 35, 355-368. [123.] T. A. Ivandini, B. V. Sarada, C. Terashima, T. N. Rao, D. A. Tryk, H. Ishiguro, Y. Kubota and A. Fujishima J. Electroanal. Chem. 2002, 521, 117-126. [124.] W. Haenni, J. Gobet, A. Perret, L. Pupunat, P. Rychen, C. Comninellis and B. Correa New Diamond and Frontier Carbon Technology 2002, 12, 83-88. [125.] A. Fujishima, C. Terashima, K. Honda, B. V. Sarada and T. N. Rao New Diamond and Frontier Carbon Technology 2002, 12, 73-81. [126.] S. Ferro and A. De Battisti J. Phys. Chem. B 2002, 106, 2249-2254. [127.] S. Ferro and A. De Battisti Physical Chemistry Chemical Physics 2002, 4, 1915-1920. [128.] S. Ferro and A. De Battisti Electrochim. Acta 2002, 47, 1641-1649. [129.] N. G. Ferreira, L. L. G. Silva and E. J. Corat Diam. Relat. Mat. 2002, 11, 657-661. [130.] A. Chatterjee, R. Wiltshire, K. B. Holt, R. G. Compton, J. S. Foord and F. Marken Diam. Relat. Mat. 2002, 11 , 646-650. [131.] P. Canizares, F. Martinez, M. Diaz, J. Garcia-Gomez and M. A. Rodrigo J. Electrochem. Soc. 2002, 149, 0118-0124. [132.] B. Boye, P.-A. Michaud, B. Marselli, M. M. Dieng, E. Brillas and C. Comninellis New Diamond and Frontier Carbon Technology 2002, 12, 63- 72. [133.] M. A. Witek and G. M. Swain Anal. Chim. Acta 2001, 440, 119-129. 28 [134.] N. Wangfuengkanagul and O. Chailapakul Anal. Sci. 2001, 17, a349-a352. [135.] R. Uchikado, T. N. Rao, D. A. Tryk and A. Fujishima Chem. Lett. 2001, 144-145. [136.] Y.-C. Tsai, B. A. Coles, K. Holt, J. S. Foord, F. Marken and R. G. Compton Electroanal. 2001, 13, 831-835. [137.] C. Terashima, T. N. Rao, B. V. Sarada and A. Fujishima Chemical Sensors 2001, 17, 124-126. [138.] A. J. Saterlay, S. J. Wilkins, K. B. Holt, J. S. Foord, R. G. Compton and F. Marken J. Electrochem. Soc. 2001, 148, E66-E72. [139.] A. J. Saterlay, J. S. Foord and R. G. Compton Electroanal. 2001, 13, 1065-1070. [140.] M. A. Rodrigo, P. A. Michaud, I. Duo, M. Panizza, G. Cerisola and C. Comninellis J. Electrochem. Soc. 2001, 148, D60-DG4. [141.] T. N. Rao, B. H. Loo, D. A. Tryk and A. Fujishima Chemical Sensors 2001, 17, 160-162. [142.] M. Panizza, P. A. Michaud, G. Cerisola and C. Comninellis Electrochem. Commun. 2001, 3, 336-339. [143.] M. Panizza, P. A. Michaud, G. Cerisola and C. Comninellis J. Electroanal. Chem. 2001, 507, 206—214. [144.] F. Montilla, P. A. Michaud, E. Morallon, J. L. Vazquez and C. Comninellis Port. Electrochim. Acta 2001, 19, 221-226. [145.] F. Marken, Y.-C. Tsai, A. J. Saterlay, B. A. Coles, D. Tibbetts, K. Holt, C. H. Goeting, J. S. Foord and R. G. Compton J. Solid State Electrochem. 2001, 5, 313-318. [146.] J.-J. Lee, B. Miller, X. Shi, R. Kalish and K. A. Wheeler J. Electrochem. Soc. 2001, 148, C183-C190. 29 [147.] T. A. Ivandini, B. V. Sarada, C. Terashima, T. N. Rao, D. A. Tryk and A. Fujishima Chemical Sensors 2001, 17, 163-165. [148.] J. Iniesta, P. A. Michaud, M. Panizza and C. Comninellis Electrochem. Commun. 2001, 3, 346-351. [149.] J. Iniesta, P. A. Michaud, M. Panizza, G. Cerisola, A. Aldaz and C. Comninellis Electrochim. Acta 2001, 46, 3573-3578. [150.] P. L. Hagans, P. M. Natishan, B. R. Stoner and W. E. O'Grady J. Electrochem. Soc. 2001, 148, E298-E301. [151.] L. Gherardini, P. A. Michaud, M. Panizza, C. Comninellis and N. Vatistas J. Electrochem. Soc. 2001, 148, 078-082. [152.] L. Gherardini, C. Comninellis and N. Vatistas Anna/i di Chimica (Rome, Italy) 2001, 91, 161-168. [153.] A. Deneuville and E. Gheeraert Vide-Sci. Techn. Appl. 2001, 56, 427-+. [154.] A. Fujishima NASA Conference Publication 2001, 210948, 150-157. [155.] O. Chailapakul, P. Aksharanandana, T. Frelink, Y. Einaga and A. Fujishima Sensors and Actuators, B: Chemical 2001, B80, 193-201. [156.] J. C. Angus, H. B. Martin, U. Landau, Y. E. Evstefeeva, B. Miller and N. Vinokur New Diamond and Frontier Carbon Technology 1999, 9, 175-187. [157.] A. Fujishima, T. N. Rao and D. A. Tryk Proc. - Electrochem. Soc. 2000, 99-32, 383-388. [158.] A. M. Zaitsev Optical Properties of Diamond: A Data Handbook; Springer- Verlag, Berlin, 2001. [159.] E. Bustarret, F. Pruvost, M. Bernard, C. Cytermann and C. Uzan—Saguy Phys. Status Solidi A 2001, 186, 303-307. 30 [160.] J. Wang, G. M. Swain, M. Mermoux, G. Lucazeau, J. Zak and J. W. Strojek New Diamond and Frontier Carbon Technology 1999, 9, 317-343. [161.] E. Gheeraert, A. Deneuville and J. Mambou Diam. Relat. Mat. 1998, 7, 1509-1512. [162.] Z. Yin, Z. Akkerman, B. X. Yang and F. W. Smith Diam. Relat. Mat. 1997, 6, 153-158. [163.] H. Opyrchal, K. K. Chin, E. Kohn and W. Ebert Diam. Relat. Mat. 1997, 6, 940-943. [164.] A. Ogasawara, T. Inushima, T. Shiraishi, S. Ohya, S. Karasawa and H. Shiomi Diam. Relat. Mat. 1997, 6, 835-838. [165.] E. Gheeraert, A. Deneuville, J. Mambou and P. Ribot Shallow-Level Cent. Semicond, Int. Conf., 7th 1997, 159-164. [166.] F. Zhang, X. Erqing, B. Yang, Y. Cal and G. Chen BandaotiXuebao1995, 16, 468-472. [167.] R. Locher, J. Wagner, F. Fuchs, C. Wild, P. Hiesinger, P. Gonon and P. Koidl Mater. Sci. Eng, B 1995, B29, 211-215. [168.] J. I. Pankove and C.-H. Qui In Synthetic Diamond: Emerging CVD Science and Technology; K. E. D. Spear, J.P. Dismukes, Ed.; John Wiley & Sons, Inc., New York, 1994, 401-407. [169.] K. M. McNamara, B. E. Williams, K. K. Gleason and B. E. Scruggs Electrochem. Solid-State Lett. 1994, 76, 2466-2472. [170.] F. Zhang, B. Yang, E. Xie, Y. Cai, X. Jiang and G. Chen Int. Conf. Appl. Diamond Films Relat. Mater., 2nd 1993, 451-456. [171.] A. Khomich, V. Polyakov, P. Perov, M. Ermakov, l. Teremetskaya, V. Varnin and G. Sokolina Int. Conf. Appl. Diamond Films Relat. Mater., 2nd 1993, 387-392. 31 [172.] G. Zou, S. Yu, Z. Jin, X. Lu and Z. Yang Proc. SP/E-lnt. Soc. Opt. Eng. 1992, 1759, 224-228. [173.] S. R. P. Silva, G. A. J. Amaratunga and C. P. Constantinou J. Appl. Phys. 1992, 72, 1149-1153. [174.] K. Okano, T. Kurosu, M. lida, T. Eickhoff, H. Wilhelm and D. R. T. Zahn Vacuum 1990, 41, 1387-1389. [175.] K. Okano, T. Kurosu, M. lida, T. Eickhoff, H. Wilhelm and D. R. T. Zahn Proc. SPlE-lnt. Soc. Opt. Eng. 1990, 1275, 25-35. [176.] C. Wild, N. Herres, J. Wagner, P. Koidl and T. R. Anthony Proc. - Electrochem. Soc. 1989, 89-12, 283-295. [177.] H. B. Martin and P. W. Morrison Electrochem. Solid-State Lett. 2001, 4, E17-E20. [178.] J. M. Stotter, J. K. Zak, Z. Behler, Y. Show and G. M. Swain submitted for publication Anal. Chem. 2002. [179.] S. Bharathi and O. Lev Appl. Biochem. Biotechnol. 2000, 89, 209-216. [180.] D. J. Blyth, J. W. Aylott, D. J. Richardson and D. A. Russell Analyst 1995, 120, 2725-2730. [181.] D. J. Blyth, S. J. Poynter and D. A. Russell Analyst 1996, 121, 1975-1978. [182.] D. J. Blyth, J. W. Aylott, J. W. B. Moir, D. J. Richardson and D. A. Russell Analyst 1999, 124, 129-134. [183.] K. E. Chung, E. H. Lan, M. S. Davidson, B. S. Dunn, J. S. Valentine and J. I. Zink Anal. Chem. 1995, 67, 1505-1509. l184-] E. H. Lan, B. Dunn, J. S. Valentine and J. I. Zink J. Sol-Gel Sci. Technol. 1996, 7, 109-116. 32 [185.] E. H. Lan, B. C. Dave, J. M. Fukuto, B. Dunn, J. I. Zink and J. S. Valentine J. Mater. Chem. 1999, 9, 45-53. [186.] E. Topoglidis, T. Lutz, R. L. Willis, C. J. Barnett, A. E. G. Cass and J. R. Durrant Faraday Discuss. 2000, 35-46. 33 CHAPTER 2 2 Materials and Methods 2.1 Boron-Doped Diamond Growth Conditions 2.1.1 Microcrystalline Boron-Doped Diamond The polycrystalline boron-doped diamond (BDD) thin films were deposited on Si using microwave-assisted chemical vapor deposition (CVD). The reactor was a 1.5 kW, 2.54 GHz commercial system (ASTex, Inc., Lowell, MA) with a 1 L quartz bell jar. The Si substrates (largely undoped, gift from Texas Instruments, Inc. (spectroscopy measurements) or p-Si(100), < 0.001 0 cm, Virginia Semiconductor, Inc. (electrochemistry measurements» were pretreated in the following manner: (i) organic solvent cleaning for 10 s each in toluene, dichloromethane, acetone, methanol, and isopropanol, (ii) a 60 s soak in concentrated hydrofluoric acid, followed by a rinse with ultrapure water and air drying, (iii) a 20 min ultrasonication in a diamond powder (0.1 pm diameter, GE Superabrasives)-acetone mixture followed by a rinse with clean acetone. The scratched substrates were then placed in the CVD reactor on top of a boron diffusion source (GS-126, BoronPlus, Techniglas, Inc., Perrysburg, OH), which, itself, rests on top of the molybdenum substrate stage. This diffusion source provides the boron dopant atoms during film growth. In some depositions, the substrates were also placed around a small block of boron nitride, which also served as a source for boron dopant atoms. The films were deposited for 4 - 20 h 34 using the following parameters: a 0.35% CH4/H2 volumetric ratio, a total gas flow of approximately 200 sccm, a system pressure of 45 torr, and a microwave power of 1000 W. The substrate temperature was estimated to be 850 °C, as measured with a disappearing filament optical pyrometer. At the end of the growth period, the methane flow was stopped, and the films were hydrogen- plasma treated for an additional 15 min at the deposition conditions listed above. Following this treatment, the plasma power and pressure were slowly reduced to 400 W and 20 torr over a 30 min period to cool the sample to less than 400 °C (estimated) in the presence of atomic hydrogen. The plasma was then extinguished and the films were further cooled under a flow of hydrogen gas for 1 h. Recently, a gas-handling system was incorporated onto one of the CVD reactors, allowing for well-controlled gas-phase doping with B2H5. This set-up enabled a substantial decrease in the film growth times, and therefore, thickness, without sacrificing film quality or B content. Post-deposition, the diamond films were chemically treated by a two-part treatment to remove adventitious metal and Sp2-bonded nondiamond carbon impurities from the surface: (i) exposure to warm 3:1 HNOg/HCI (v/v) for 30 min followed by a rinse with ultrapure water and (ii) exposure to warm 30% H202 (w/v) (Aldrich Chemical) for 30 min followed by a rinse with ultrapure water [1]. The films were dried and then placed into a microwave-assisted CVD system for hydrogen-plasma treatment. The wet chemical treatments introduce significant levels of oxygen, ca. 12- 15 atomic percent, so the final hydrogen-plasma 35 treatment is required to rehydrogenate the surface. The treatment conditions were as follows, 45 min at a total flow of 200 sccm, a system pressure of 35 torr, and a microwave power of 1000 W. The films were slowly cooled over a 15 min period in atomic hydrogen, as described above. The methane and hydrogen gases were of ultrahigh purity (99.999%). Films were characterized by Optical and atomic force microscopy (morphology), Raman spectroscopy (microstructure), cyclic voltammetry, and 4-point probe resistivity measurements (electrical conductivity). The estimated film thickness was between 2 and 5 pm, as measured by scanning electron microscopy, and the apparent in-plane resistivity was, generally, 0.1 0 cm or less. This resistivity is referred to as “apparent" because the measurement was made with the diamond attached to the electrically conducting Si substrate. Electrochemical characterization involved measuring the cyclic voltammetric responses for two redox couples, ruthenium hexaamine ([Ru(NH3)5] +3M) and —3/-4 ferrocyanide ([Fe(CN)5] ). If the peak currents varied linearly with the square root of the scan rate, and the peak splitting was near 0.070 V at 0.1 V/s, the electrodes were considered high quality. The heterogeneous electron transfer rate constant for [Ru(NH3)5] +3” is sensitive to the density of electronic states at the formal potential, while the rate constant for [Fe(CN)5] "3"4 is sensitive to both the density of states and the diamond electrode surface chemistry. Previous studies of BDD electrodes showed that the highest rate constants for [Fe(CN)5] "3"4 are observed at the hydrogen-terminated diamond surface [2]. The doping level was estimated to be in the mid 10’19 to low 10'20 range, based 36 on boron nuclear reaction analysis measurements performed on other films deposited using similar conditions. Optical microscopy revealed that the polycrystalline films completely covered the substrate, showing no cracks or voids. Atomic force microscopy showed a well-faceted, polycrystalline morphology with a nominal grain size of 1 — 2 pm. The crystallites were primarily of octahedral and cubo-octahedral shape with a relatively low fraction of secondary growths [1]. Raman spectroscopy revealed a sharp, one-phonon diamond line centered near 1332 cm'1. The line width was 6 - 8 cm’1, which is slightly larger than the 2 cm'1 width observed for a single-crystal diamond reference. The broader line width is due, in part, to scattering by grain boundaries, which reduces the phonon lifetime. The one—phonon diamond line exhibited some asymmetry, due to the Fano-effect, and was superimposed on a very low photoluminescence background [3]. There was also negligible scattering intensity between 1500 and 1600 cm'1, due to spz-bonded nondiamond carbon impurities. The Fano interaction is attributed to a quantum mechanical interference between the Raman phonon and transitions from the broadened impurity band to continuum states composed of excited acceptors and valence band states [3, 4]. The Fano line shape is asymmetric, with an intensity enhancement of the high wavenumber side and an intensity reduction on the other flank (low wavenumber side). Fano-like Iineshapes are observed as the boron-dopant concentration 3 reaches 1019 cm' , and above. The doping level may be qualitatively estimated, based on the level of asymmetry. For the observed asymmetry, the doping level 37 was estimated to be in the high 1019 cm'3 range [4-6]. It Should be mentioned, however, that the dOping level threshold for the appearance of the diamond line asymmetry varies somewhat from report to report [7]. 2.1.2 Nanocrystalline Boron-Doped Diamond The nanocrystalline BDD thin films were deposited using microwave- assisted CVD, as described above. The films were deposited on scratched and cleaned Si (p-Si (100), < 0.001 0 cm), using the following source gas mixture: 1 sccm CH4, 4 sccm H2, 1 sccm B2H5 (0.1%) diluted in H2, and 94 sccm Ar. The microwave power was 800 W, the system pressure 140 torr and the estimated substrate temperature was 800 °C. The growth time was 2 h, resulting in a film thickness of ca. 4 pm, as determined from weight change measurements. The resulting electrodes were highly conductive (< 0.05 0 cm) being prepared from a gas phase BZHG concentration of approximately 10 ppm. The electrodes were used as prepared, with no additional pretreatment (e.g., polishing) or potential cycling required for activation. 2.1.3 Freestanding Boron-Doped Diamond OTE A freestanding, mechanically polished, BDD disk (380 pm thick and 8 mm diameter) was used as the optically transparent electrode (OTE) for the UV/vis measurements of ferrocene and cytochrome c, which are discussed in Chapter 7. This disk was grown by microwave plasma CVD (2.54 GHz) In a customized reactor at the Naval Research Lab, based on a 5 kW commercial system (model 38 5400, ASTeX, Inc). The ultrahigh-purity reactants (CH4 and BzHe prediluted in H2) were hydrogen purified, using a Pd diffusion cell. A computerized system was used to control and monitor the reactor pressure and reactant gas flow rates. Samples were grown on 2 in diameter refractory metal substrates, which released freestanding plates of CVD diamond upon rapid cooling from the deposition temperature. The diamond plate was then laser machined to the required sample size and cleaned as follows: aqua regia followed by a boiling mixture of concentrated sulfuric and nitric acids. The sample was then mechanically polished using a resin-bonded diamond wheel (Coburn Eng. Co. Ltd.). Further details of this growth process are reported In [8]. The other freestanding films discussed in Chapter 7 have the following thicknesses: (I) freestanding CVD white diamond disk, 230 um; (ii) polished BDD on CVD white diamond disk, 210 pm; and (iii) unpolished BDD on CVD white diamond disk, 220 pm. Specifically, the growth conditions used for the freestanding disk consisted of a tungsten substrate at 810 °C, reactor pressure of 117 torr, 5.0 kW of microwave power, and gas flow rates of 115 sccm for H2, 8 sccm for CH4, and 0.4 sccm of 0.1% BZHG diluted in H2. After polishing, AFM (Nanoscope Ill, Veeco Instruments, Santa Barbara, CA) measurements in the contact mode revealed a smooth surface, with a root-mean-square roughness of 7 nm over a 10 um2 area. 39 2.1.4 Boron-Doped Diamond IR OTE The OTE used for the IR measurements reported in Chapter 7 consisted of a microcrystalline BDD thin film (~ 2 pm thick) deposited onto a highly resistive (i.e., very low doped) Si substrate (650 pm thick, Texas Instruments, Inc.) via microwave-assisted CVD. The deposition procedure was the same as is detailed above, in Section 2.1.1. A nonconducting Si substrate was used to maximize the optical throughput. Typically, films were doped with 1 ppm BZHS diluted in H2 using growth times ranging from 4 - 10 h. The effect of the growth time on the optical throughput will be discussed in the last chapter. 2.2 Sample Preparation All aqueous solutions were prepared using water purified using either a Barnstead E-Pure or Millipore system (resistivity > 17 MO cm). Ferrocyanide (Aldrich) was recrystallized from ethanol. All other analyte and solution preparation procedures are described below. 2.2.1 Ferrocene Ferrocene and sodium perchlorate (NaCIO4) were purchased from Aldrich and used as received, without further purification. Tetrabutyl ammonium perchlorate (TBACIO4) salt was purchased from Fluka and used without further processing. Dichloromethane (Spectrum, spectrophotometric grade) was distilled over calcium hydride. Acetonitrile (Merck, HPLC grade) was dried by 40 distillation over calcium hydride, followed by twice passing over a column of activated alumina. 2.2.2 Cytochrome c Cytochrome c stock solutions were obtained from the Ferguson-Miller lab (MSU Biochemistry), courtesy of Dr. Denise Mills. For the experiments described herein, the stock solution (3.5 mM cytochrome c, 0.3 M NaCI, 10 mM Tris HCI pH 7) was diluted with purified water to the desired protein concentration. Stock solutions of cytochrome c were prepared by chromatographically purifying and concentrating horse heart cytochrome 0 purchased from Sigma Chemical Company, following a published procedure [9]. For ionic strength studies, measured amounts of NaCI were added to the protein solution, described above, to yield the desired ionic strength. Cytochrome 0 concentrations were determined optically from the amplitude of the reduced minus oxidized difference band at 550 nm using 17 mM'1 as the extinction coefficient [10]. 2.2.3 Cytochrome c Oxidase The cytochrome c oxidase (cco) protein samples were also received as part Of a collaboration with the Ferguson-Miller lab. The bovine heart oxidase was isolated from bovine heart muscle, as previously described by Yoshikawa [11]. The enzyme was concentrated to 0.7 - 1.0 mM in a solution of 40 mM HEPES buffer pH 7.4, by centrifugal filtration. The Rhodobacter Sphaeroides oxidase was isolated from the YZ-300 strain [12] engineered to over-express a histidine-tagged form of the enzyme [13] and was cultured aerobically in 41 Sistrom’s growth media at 30 °C with rapid swirling. Global 15N labeling was achieved by substituting 15NH4CI (99% 15N; Isotec, Miamisburg, OH) into the growth media which constituted ~95% of the cell’s nitrogen. The isolated enzyme was concentrated In a 40 mM HEPES pH 7.4, 1 mM EDTA and 0.1% Iauryl maltoside buffer medium with a 50 kDa centrifugal filter (Ultrafree filter, Millipore, MA). Deuterium exchange of both oxidases was performed by repetitive concentration and incubation of the enzyme in 99.5% DZO enriched buffer. To ensure maximal exchange, four concentration steps and four incubation steps of 6 - 8 h were performed (and the enzyme was turned over). Enzyme concentrations were determined from UV/vis absorbance spectra, using 26°6— 3630 = 24 mM'1 cm‘1 [14]. 2.3 Cyclic Voltammetric Experiments Figure 2-1 shows a drawing of the standard 3-electrode glass cell that was employed for the cyclic voltammetric (CV) experiments described in Chapters 3 and 4 [1]. The cell volume was approximately 5 mL. In this cell, a ChemRaz® (Ace Glass) o-ring defined the electrode area (0.39 cm2). This type of o-ring was selected due to its chemical resistance toward organic solvents. Contact to the working electrode was made through a conductive lnGa alloy applied to the backside of the cleaned silicon substrate. This Side of the substrate was pressed onto a copper current-collecting plate, and electrical connection was made to this plate. In the ferrocene measurements, a Ag wire, encased in Teflon®, served as 42 the quasi-reference electrode, and 3 Pt mesh was used as the counter electrode. For the CV measurements of cytochrome c, a similar, smaller-scale cell was employed, requiring a much smaller volume (ca. 0.5 mL), the counter was a Pt wire coil, and the reference was changed to a commercial (Cypress Systems, Lawrence, KS) Ag/AgCl (1.0 M KCI) electrode. The exposed electrode area in this cell was 0.2 cm2, as determined by a Viton o-ring. All solutions were deaerated by gently bubbling with nitrogen, which was then kept flowing over the solution during measurements. Figure 2-1 Conventional 3-Electrode Glass Electrochemical Cell A diagram of the glass electrochemical cell used in cyclic voltammetric experiments. Cell volume ca. 5 mL. A conventional three-electrode configuration is employed, including (a) a copper current collector plate, (b) a Ag wire quasi-reference electrode, (c) 3 Pt mesh counter electrode, (d) a nitrogen gas inlet, (a) an o-ring, and (t) a BDD thin-film on Si substrate as the working electrode. t t I i a: it :3: 15% b\ s it, . a? a '-" I! \ Eagwmyamg f Im| mi] 43 Cyclic voltammetric experiments were conducted using either an EG&G Princeton Applied Research Model 263A (cytochrome c) or a CH Instruments Model 650 (ferrocene) computer-controlled potentiostat. Low analyte concentration (0.1 mM) and iR compensation were used for the CV measurements of ferrocene, in an effort to minimize ohmic distortion. Digital compensation of the uncompensated solution resistance, via a positive-feedback method, was applied at 70% to generate this data, which was used for the rate constant calculations. Glassy carbon electrodes (GO-30), obtained from Tokai Ltd. (Japan), were used for some CV measurements of ferrocene described in Chapter 3. These electrodes were polished smooth on felt pads, using slurries of 1.0, 0.5 and 0.3 pm alumina (Buehler) in ultrapure water. The electrodes were rinsed with deionized and purified water between each polishing step. 2.4 Spectroelectrochemistry 2.4.1 CcO Experiments - Au Grid OTTLE Cell Figure 2-2 shows a schematic of the optically transparent thin-layer electrochemical (OTTLE) cell used for the spectroelectrochemistry of cytochrome c oxidase, described in Chapter 6. This cell was custom built, according to a design proposed by Mantele and coworkers [15, 16]. The protein solution formed a thin layer on a 70% transparent, 6 pm thick gold grid working electrode (200 wires per inch; Buckbee-Mears, St Paul, MN) between two Ban windows. The gold mesh electrode was in electrical contact with a platinum wire counter 44 electrode and a AglAgCl reference electrode (0.197 V vs. NHE; BAS Instruments, West Lafayette, IN). Figure 2-2 Optically Transparent Thin-Layer Electrochemical Cell Design Employing a Au Grid Optically Transparent Electrode Design schematic of the optically transparent thin-layer electrochemical (O TTLE) cell used for U V/vis and F TIR spectaoelectrochemical studies of cytochrome c oxidase. Components of the cell: (a) Delron plastic body (b) window (25 mm) mounting piece (0) o-ring (d) Au mesh working electrode (e) Au ring contact for working electrode (0 solution inlet channel (g) Pt wire counter electrode (h) reference electrode port/solution outlet channel and (i) lR windows (top 20 x 2 mm, bottom 25 x 2 mm). Drawing not to scale. t - e \ ‘ h I "\i‘: . . "if"? x \ .:.'. .‘ \ \ ;\‘ . \\;\\ \\_\";V :36:- I .~\. ‘\.\" \ N“ If: \ ‘ d A :‘LL'EtE‘LI'i \ "'1'“. ”1 a c \ The cell was filled with the 25 mm window, and the working and counter electrodes in place, in the DelronTM plastic body, as is shown in Figure 2-2. A 10 pL drop of protein solution was placed onto the gold grid and the two windows were progressively pressed together by screwing the top window into place against the resistance of rubber o-rings, until the sample drop completely filled the 20 mm diameter. Contact to the counter and reference electrodes was 45 established by filling the space around the periphery of the windows with supporting electrolyte solution via a tangential port. Although the sample solution was in direct contact with the surrounding electrolyte solution, the rate of dilution was minimal (< 5% in 24 h). The sample solution consisted of 0.7 - 1 mM cco in 40 mM HEPES buffer (pH 7.4), 200 mM KCI as the supporting electrolyte, and a cocktail of mediators (Table 2-1) at 40 DM each. This series of mediators was necessary to accelerate the electrochemistry. At such a relatively low concentration, the mediator solution did not interfere with the spectra of the cco enzyme. Prior to use, the electrode grid was cleaned in an acid solution (3:1 (v/v) H2804 and 30% H202) by soaking for 30 min followed by copious rinsing with ultrapure water. The cleaned grid was then chemically modified by immersion in a 2 mM cysteamine (Aldrich) aqueous solution, according to procedures established for obtaining direct electrochemistry of cytochrome c [17, 18]. After 2 h, excess modifier was removed by rinsing with ultrapure water, and the mesh was dried under a stream of nitrogen gas. The cysteamine formed a monomolecular layer on the gold electrode by self-assembly, due to strong Au-S bonds. This monolayer prevents proteins from adhering and denaturing on the surface. Proteins may interact directly with the modifier to undergo electron transfer at the electrode. However, when large proteins contain deeply buried metal centers that result in sluggish electron transfer between the protein and the modified electrode, additional efforts are required to facilitate rapid electron 46 transfer. Therefore, a cocktail of mediators with midpoint reduction potentials, spaced ~0.06 - 0.10 V, was added to the sample solution to exchange electrons between the electrode and the enzyme. Table 2-1 lists the mediators and their redox midpoint potentials. Table 2—1 Mediators Used in Spectroelectrochemical Measurements of Cytochrome c Oxidase Em (V) vs. Mediator AgIAgCl Solvent Ferricyanide 0.216 "'20 Dimethylparaphenylendiamine (DMPPD) 0.163 DMSO 1,1-dimethylferrocene 0.133 DMSO Tetrachlorobenzoquinone 0.072 DMSO Tetramethylparaphenylendiamine (TMPPD) 0.062 DMSO 2,6-dichlorophenolindophenol (DCPIP) 0.009 H2O Hexaammineruthenium (III) chloride -0.008 H2O 1,2-napthoquinone -0.063 DMSO Trimethylhydroquinone -0.108 DMSO Menadione -0.220 DMSO 2-hydroxy-1,4-naphthoquinone -0.333 DMSO Anthraquinone-Z-sulfonate -0.433 DMSO Benzyl viologen -0.568 H2O Methyl viologen -0.654 H2O DMSO, dimethylsulfoxide 47 Spectroscopy was performed in the OTTLE cell, as a function of applied potential. An EG&G Princeton Applied Research 263M potentiostat was used to initiate the electrochemical reactions. This instrument was operated at a scan rate of 0.010 V/s and under no iR compensation. The UV/vis spectroscopy was recorded with a Hewlett Packard 8453 diode array spectrometer; and FTIR spectroscopy was performed, separately, with a Bruker Equinox 55 instrument, using a MCT detector. The IR optics consisted of a KBr beamsplitter, used in conjunction with a Ge filter (OCLI L02547-9, Santa Ross, CA), which reduced spectral bandwidth. Temperature control was maintained at 10 1 0.01 °C as described below (section 2.5). A stable background spectrum could be achieved after temperature equilibration of ~ 60 min. At this point, the potentiostat was used to electrochemically poise cytochrome c oxidase in a reduced or oxidized form. The enzyme was held at 0.50 V (vs. Ag/AgCl) to generate the oxidized form and —0.50 V (vs. Ag/AgCI) for the reduced form. Electrolysis was complete within 2 min and was recognized by a steady-state current reading and identically shaped successive spectra. Typically, the oxidized and reduced spectra were a ratio of 800 scans for a single-beam spectrum. Each 800-scan acquisition lasted -1 , . . ~ 80 5. All spectra were collected at 4 cm resolutIon. No baseline corrections were made to any of the spectra. 2.4.2 BDD UVlvis OTTLE Cell The freestanding diamond OTE (growth described in Section 2.1.3) was mounted into a specially designed thin-layer cell for the transmission UV/vis 48 spectroelectrochemical measurements detailed in Chapter 7. A schematic of this design is shown in Figure 2-3. This is a very Simple and user-friendly design, as the entire cell is conveniently housed in a 1 cm2 quartz cuvette. Figure 2-3 UV/vis Spectroelectrochemical Cell Employing a Freestanding Boron-Doped Diamond Optically Transparent Electrode Side and front views of the UV/vis OTTLE cell designed for freestanding BDD disk OTES. As presented, the cell consists of (a) the BDD freestanding disk working electrode, (b) a mica spacer, (c) the Kel-FTM body, (d) a Pt foil contact to the working electrode, (a) a Ag wire reference electrode, (0 8 Pt coil counter electrode, (9) a standard 1 cm2 quartz cuvette and (h) springs. Drawing not to scale. light—> SIDE FRONT The thin-layer cavity was formed between the diamond OTE and the wall of the cuvette, using an 80 — 150 pm mica spacer (cell volume ~ 3 — 6 [.IL). An o-ring, contained within the KeI-fm body, combined with spring rods, pressed the electrode against the spacer. Electrical contact with the diamond disk was made with a piece of platinum foil, also contained within the KeI-fTM body. A small channel was cut in the spacer to connect the thin-layer cavity with the solution 49 contacting the reference and counter electrodes. A Ag wire, coated with Teflonm, was used as the reference electrode. The potential of this reference electrode was —0.01 V relative to a saturated calomel reference electrode, as measured in 1 M KCI. A coil of Pt wire was used as the counter electrode. The solution was introduced into the cell, using a syringe, taking care to avoid trapping air in the thin-layer cavity or in the solution compartment hosting the reference and counter electrodes. The active area of the electrode was defined by the opening of the mica spacer (0.39 cm2). UV/vis spectrophotometric measurements were made, using a Hewlett Packard Model 8453 spectrometer with a diode array detector. An EG&G Princeton Applied Research 263M potentiostat was used to control the applied potential. A cyclic voltammetric waveform was used with a 0.002 V/s scan rate and no iR compensation when the current response was measured simultaneously with the spectra. Othenrvise, a potential step platform was employed, in which the potential was stepped to a constant value while acquiring an Optical Spectrum. A steady-state current reading and identical successive spectra indicated that complete electrolysis had been achieved. 2.4.3 BDD IR OTTLE Cell I The diamond OTE for FTIR measurements was a diamond thin-film deposited on undoped Si, described above In Section 2.1.4. For the ferrocene measurements reported in Section 7.3.2, this OTE was mounted onto the metal body of a custom designed cell shown in Figure 2-4. A TeflonTM spacer defined 50 the thickness (25 to 200 pm) of the thin-layer cavity, which was formed between the AgCI window and diamond electrode. The cell volume is ca. 3 pL using a 25 um spacer. The cell was filled by filling the solution port containing the reference and counter electrodes. An lnGa alloy was used to make a contact between the metal body and diamond electrode. A Pt coil counter electrode was wrapped around a Ag wire reference electrode that was encased in Teflonm. This cell was the prototype for an IR spectroelectrochemical cell designed for BDD thin-film on Si OTES. Figure 2-4 IR Spectroelectrochemical Cell Employing a Boron-Doped Diamond Optically Transparent Electrode - Model I The spectroelectrochemical cell consists of the following components (a) a TeflonTM spacer, (b) a AgCI window (25 x 2 mm), (c) the metal body, (d) 3 Pt wire counter electrode, (a) 3 Ag wire encased in Teflon (reference electrode), (0 the BDD OTE working electrode, and (g) the lnGa contact to the working electrode. 51 A major drawback of this design was the method in which the OTE was mounted in the cell. One difficulty encountered in working with the BDD thin- films on nonconductive Si was the necessity to make electrical contact to the front of the electrode, the same side, which is exposed to the analyte solution. Therefore, the electrode was placed into the metal body and an epoxy mask was used to prevent the electrical contact point from encountering the solution. The epoxy mask did Shield the lnGa alloy from the solution; however, it did not enable easy exchange of OTES, as OTES were sacrificed upon removal from this cell. In our next design, we aimed to find a different method for shielding the working electrode contact from the solution without permanently fixing the OTE to the cell body. No electrochemical characterization was carried out for this cell. At the time this cell was used, there was no capability to produce films using the gas phase dopant source. Instead, films were doped using a solid B source, requiring longer growth times, and therefore, yielding either highly conductive, thicker films with poor transparency, or lightly doped, thinner films with sufficient optical transparency. The thinner films were conductive enough to sustain electron transfer reactions and were used in the potential step experiments. However, attempts to measure CVs at such films were unsuccessful. Potential step spectroelectrochemical experiments for ferrocene demonstrated that BDD thin-films on Si were useful as OTES in the IR region of the spectrum. Upon this finding, we quickly changed our OTTLE cell to the design described below. 52 2.4.4 BDD IR OTTLE Cell II Figure 2-5 IR Spec; --' ‘ L ' ‘ Cell Employing a Boron-Doped Diamond Optically Transparent Electrode — Model II Schematic of the three-piece IR OTTLE cell designed to incorporate a BDD thin film on Si OTE. The cell was constructed from Kel-F plastic and consists of the following components (a) a solution inlet port, (b) the reference electrode/solution outlet port, (c) the BDD thin-film on Si OTE, (d) a Pt wire counter electrode, (9) the Pt ring contact for the working electrode, (f) o-rings, and (9) an IR window (9 x 5 mm). 9 IR WINDOW BDD OTE MOUNTING PIECE CENTER BODY MOUNTING PIECE The spectroelectrochemical cell used for the FTIR measurements of ferrocyanide and cytochrome 0 (Sections 7.3.4 and 7.3.5) consisted of a three- piece assembly constructed from KeI-fm. For such measurements, a diamond thin film deposited on undoped Si, described above in Section 2.1.4, was used as the OTE. The diamond OTE was mounted into the thin-layer cell, as shown in Figure 2-5. This piece was held onto the center body by four screws. A 1 — 3 ].IL drop was placed on the diamond OTE and pressed to a thickness of ca. 10 pm by screwing the NaCl window piece into place. The volume around the NaCI window, containing the Pt wire counter electrode, was then filled with electrolyte 53 via an external port. Electrical contact to the diamond electrode was achieved by pressing a flat Pt ring onto its surface. An o-ring with a smaller diameter than the Pt ring ensured that the solution was isolated from this contact. The reference electrode, 3 Ag wire encased in Teflonm, fit into a small cavity in the center body. FTIR spectra were measured, using a Bruker Equinox 55 Spectrometer with a MCT detector; the potential was applied by an EG&G Princeton Applied Research 263M potentiostat in potential step mode. Typically, difference spectra were an average of 200 scans acquired at 15 °C, using the temperature control apparatus detailed later in Section 2.5. All spectra were collected at 4 cm'1 resolution, and no baseline corrections were made. As was mentioned for the cco studies, temperature equilibration required ~60 min. At this point, the potentiostat was used to electrochemically poise the analyte in a reduced or oxidized form. Within 2 min, the electrolysis was complete and was indicated by a steady-state current reading and identical successive spectra. 2.5 Temperature-Control Set Up The necessity for precise temperature control during IR spectroelectrochemical measurements will be discussed in Chapter 5. The system designed to meet such experimental needs is shown in Figure 2-6. This system maintained a very precise temperature (i 0.005 °C), yet allowed for facile sample changing and high optical throughput. 54 Figure 2-6 Temperature-Control Set Up Schematic of the experimental setup used for maintaining precise temperature control during the spectroelectrochemical experiments. The components include (a) copper coil, (b) liquid nitrogen dewar, (0) copper transfer line insulation, (d) dewar transfer arm, (e) nichrome heater element, (0 Si diode temperature sensor, and (g) cryostat box. As depicted, the nitrogen gas flow originates in the copper coil and is vented from the cryostat box. Lake Shore 340 T control unit \ —1 gas flow inlet fl / t~ \(D gas flow F outlet C A :: 69 9 The sample temperature was controlled by surrounding the sample with a stream of gas at a precisely controlled temperature. To achieve this, nitrogen gas, at room temperature, was flowed into a copper coil, which was submerged into a dewar of liquid nitrogen. The gas was then passed through a glass-dewar transfer arm that contained a nichrome heating element, used to heat the gas to 55 the desired temperature in a controlled manner. The heated gas flowed into the cryostat box that was located In the sample compartment of the FTIR spectrometer. Figure 2-7 Cryostat Box Diagrams Top and side views of the cryostat box sample holder employed in the IR spectroelectrochemical measurements. The components include: (a) Spectroelectro- chemical cell mount, (b) sample compartment adapter, (0) Germanium filter, (d) temperature sensor, (e) temperature-controller connection, (f) electrochemical connections, and (g) KBr infrared window. gas flow inlet gas flow a \ F outlet TOP VIEW SIDE VIEW The cryostat box, shown in Figure 2-7, was constructed from DelronTM plastic and contained a removable top. This style was well suited for the IR spectroelectrochemical measurements described in the following chapters (Chapters 6 and 7), as it allowed for quick purging (i.e., within 20 S) of the sample compartment, precise temperature control, and easy sample changing. The box was designed to fit entirely inside of the sample compartment of the Bruker 56 Equinox 55 spectrophotometer. A mounting screw was used to ensure that the position Of the box was reproducibly and properly aligned in the IR beam. The front side of the box was equipped with an outlet for the temperature controller and for electrochemical connections between the potentiostat and the electrodes in the sample cell. A Si diode temperature sensor (Lake Shore Cryotronics, Inc., DT-471-BR), mounted onto an aluminum-alloy heat-Sink material, was attached to the inside wall of the cryostat box. The Si diode was important in this design because it had a very fast response time, relative to other types of temperature sensors (e.g., thermocouples or platinum wires). Two sample compartment adaptors were used to seal the cryostat box to the instrument, maintaining an air-free purge. Purging was important to suppress spectral contributions from the air, such as those due to carbon dioxide and water vapor. The box was sealed from the instrument by attaching a KBr window and a Ge window. The Ge window acted as a 3.3 pm filter, cutting out high- frequency radiation greater than 4500 cm'1. Decreasing the bandwidth, in this way, served to improve the Signal-to-noise ratio and decreased required scanning time. The spectroelectrochemical cell was mounted onto a support that ensured maximum optical throughput. The temperature was controlled to i 0.005 °C by a control unit (Model 340, Lake Shore Cryotronics, Inc.). The thermal contributions to the spectra were largely eliminated with this level of control. 57 2.6 I1] I21 I31 [41 [5.] [51 [7-1 [8-1 [9-1 [10.] [11.] [12.] References M. C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M. D. Koppang, J. E. Butler, G. Lucazeau, M. Mermoux, J. W. Strojek and G. M. Swain Anal. Chem. 2000, 72, 3793-3804. M. C. Granger and G. M. Swain J. Electrochem. Soc. 1999, 146, 4551- 4558. U. Fano Phys. Rev. 1961, 124, 1866-1878. J. W. Ager, W. Walukiewicz, M. McCluskey, M. A. Plano and M. l. Landstrass Appl. Phys. Lett. 1995, 66, 616-618. K. Ushizawa, K. Watanabe, T. Ando, l. Sakaguchi, M. Nishitani-Gamo, Y. Sato and H. Kanda Diam. Relat. Mat. 1998, 7, 1719-1722. P. Gonon, E. Gheeraert, A. Deneuville, F. Fontaine, L. Abello and G. Lucazeau J. Appl. Phys. 1995, 78, 7059-7062. M. Mermoux, B. Marcus, G. M. Swain and J. E. Butler J. Phys. Chem. B 2002, 106, 10816-10827. J. K. Zak, J. E. Butler and G. M. Swain Anal. Chem. 2001, 73, 908-914. D. L. Brautigan, S. Ferguson-Miller and E. Margoliash J. Biol. Chem. 1978, 253, 130-139. M. Brunori, P. Sarti, A. Closimo, G. Antonini, F. Malatesta, M. G. Jones and M. T. Wilson EMBO J. 1985, 4, 2365-2368. S. Yoshikawa, T. Tera, Y. Takahashi, T. Tsukihara and W. S. Caughey Proc. Natl. Acad. Sci. U. S. A. 1988, 85, 1354-1358. Y. Zhen, J. Qian, K. Follmann, T. Hayward, T. Nilsson, M. Dahn, Y. Hilmi, A. G. Hamer, J. P. Hosler and S. Ferguson-Miller Protein Expression and Purification 1998, 13, 326-336. 58 [13.] D. M. Mitchell and R. B. Gennis FEBS Lett. 1995, 368, 148-150. [14.] W. Vanneste Biochemistry 1966, 5, 838-848. [15.] D. Moss, E. Nabedryk, J. Breton and W. Mantele Eur. J. Biochem. 1990, 187, 565-572. [16.] F. Baymann, D. A. Moss and W. Mantele Anal. Biochem. 1991, 199, 269- 274. [17.] P. M. Allen, H. A. 0. Hill and N. J. Walton J. Electroanal. Chem. 1984, 178, 69-86. [18.] l. Taniguchi, K. Toyosawa, H. Yamaguchi and K. Yasukouchi J. Electroanal. Chem. 1982, 140, 187-193. 59 CHAPTER 3 3 Ferrocene Electron Transfer Kinetics Measured at Boron-Doped Diamond Electrodes 3.1 Introduction Good quality, hydrogen-terminated diamond thin-film electrodes exhibit properties superior to other carbon electrodes: (i) a low and stable background current over a wide potential range, (ii) a resistance to fouling, due to weak adsorption of polar molecules on the nonpolar, H-terminated surface, (iii) a wide working potential window, (iv) excellent morphological and microstructural stability at extreme anodic and cathodic potentials, and current densities, and (v) good electrochemical responsiveness for several redox systems without any conventional pretreatment [1]. There have been a number of papers reviewing the electrochemical response of boron-doped diamond (BDD) thin films toward a variety of aqueous redox systems [1-5]. However, there have been very few reports of the electrode’s responsiveness toward nonaqueous redox systems [6- 9]. A question that needs to be answered is, do diamond electrodes possess the same attractive features in nonaqueous media that they do in aqueous media? Using the ferrocene/ferricenium redox couple as a test system, one can begin to answer this question. Ferrocene is routinely used as a model system in electrochemical kinetic studies, because it demonstrates “ideal” behavior in a variety of solvents. Many researchers rely on the well-behaved kinetics of ferrocene as an internal standard in nonaqueous electrochemistry [10-12]. The 60 one-electron oxidation of ferrocene to the stable ferricenium cation is a simple electron transfer reaction, in that there are no mechanistically complicated steps arising from adsorption or associated chemical reactions. Generally, this redox couple exhibits highly reversible, well-behaved voltammetric responses at a variety of electrode materials (Pt, Au, glassy carbon (GC)) [11, 13-21]. A schematic for the electron transfer reaction of the ferrocene/ferricenium couple is Shown below, in which the formal oxidation state of the metal changes from +2 to +3. 4? SR“ +2 +3 .. Fe ' Fe + e E0 = 0.307 V VS SCE FERROCENE FERRICENIUM A cyclic voltammetric study of the ferrocene/ferricenium redox reaction at boron-doped, microcrystalline diamond thin-film electrodes as a function of the scan rate, the composition of the electrolyte solution is reported on herein. 0 app , were calculated Apparent heterogeneous electron transfer rate constants, k from the dependence of the peak potential separation, AEp, on the scan rate, according to the theory developed by Nicholson [22, 23]. Cyclic voltammetry (CV) is one of the most versatile electroanalytical methods for the study of redox-active species. The technique has been used extensively in electroanalytical studies and for the characterization of organic and inorganic compounds, as well as for probing the properties of electrode surfaces, and, more recently, for studies of biological molecules (see [24] for a recent, 61 comprehensive review of dynamic electrochemical methods). The effectiveness of this technique lies in its ease of measurement combined with the wealth of information contained in the resulting voltammogram. From cyclic voltammetric studies, one can obtain information about the thermodynamic and kinetic properties of an electroactive analyte, as well as information on the nature of its interaction with the electrode surface. Often, CV is the first experiment performed In the study of a new compound or electrode material. A CV experiment is conducted by cycling the potential of an electrode, which is immersed, Often unstirred, in a solution containing the electroactive analyte, and measuring the resulting current. A potentiostat is used to apply the potential and measure the resulting current. For optimal performance, three electrodes are used in a standard CV experiment. A conventional three- eIectrode glass cell employed for CV measurements was presented as Figure 2-1 in the previous chapter. The electrolysis takes place at the working electrode, to which a potential is applied relative to the reference electrode. The current required for electrolysis flows through the counter electrode. In an unstirred solution, diffusion is responsible for transporting the redox analyte to and from the electrode/solution interface. The voltage applied to the working electrode is linearly scanned from an initial value to a “switching” potential, at which the direction of the scan is reversed. This results in a triangular waveform of voltage as a function of time. The slope of the waveform gives the scan rate in units of potential/time. The resulting voltammogram is a plot of the current response as a function of the applied potential. 62 A reversible cyclic voltammogram, with the current limited by semi-infinite linear diffusion, has a very characteristic shape, as is shown in Figure 3-1. Important features of the voltammogram include the initial (and final) potential, E], the switching potential, E), the reduction and oxidation peak potentials, Egedand ~red on respectively, the reduction peak current, Ip p , , and the oxidation peak current, ig" . To a first approximation, the formal reduction potential, E", can be determined as the average of the peak potentials. Figure 3-1 Simulated Cyclic Voltammogram - Diffusion-Limited Case Simulated cyclic voltammogram for a reversible, one electron transfer reaction at a planar electrode (A = 0.39 cmz) for the diffusion-limited case at 0.1 V/s. The electrolyte solution initial/y contained the reduced form but not the oxidized form of the redox analyte. Other parameters used in the simulation include: E0 = 0.025 V, a = 0. 5, ke) = 0.1 cm/s, c0 = 1x10‘3 M, and D0 = 5x105 cmZ/S. -20 40 -60 - red . A I L I A lEp A 1 + I A I -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 Potential (V) 63 3.2 Results The oxidation of ferrocene is a one-electron process leading to a chemically stable ferricenium cation. Figure 3-2A shows a cyclic voltammetric i-E curve for 0.1 mM ferrocene in acetonitrile containing 0.1 M TBACIO4 at a diamond film electrode. The scan rate was 0.2 V/s. The background response at the same scan rate is also presented, for comparison. The scan was initiated at 0.70 V and was swept in a negative direction, at a constant rate of 0.2 V/S, to the switching potential, 0.10 V. Scanning in the negative direction results in the reduction of electrogenerated ferricenium to ferrocene, as indicated by the peak at approximately 0.45 V. At 0.10 V, the direction of the scan is reversed, and the peak due to the oxidation from ferrocene to ferricenium is observed at approximately 0.50 V. According to theory, for a redox reaction to be kinetically reversible, as dictated by the Nernst equation, EZE”'_flln{m , NF [0] PO a reaction must occur as fast as the potential sweep rate, so that the concentrations of the reduced, [R], and oxidized, [0], forms are in equilibrium with each other at the surface. In this expression, R is the molar gas constant (8.314 J/mol K), T is temperature (K), n is the number of electrons transferred per mole, F is Faraday’s constant (96,485 C/mol) and E” is the formal reduction potential of the couple. Reversible systems are characterized by three criteria: 64 (1) a peak potential separation, AEp, of 59m mV A5,, = E12" — 13;?" = 59/n m V (2) an oxidation and reduction peak current ratio of 1 -(),\l Il’ _ -r't~'t/ /) (3) a peak potential separation, AEp, and therefore, an E1,2 value that is independent of scan rate Ejsx _ E;)Ud — AE [) 2 ‘2' 151/2 = A redox couple whose AEp increases with increasing scan rate is considered to be quasi-reversible (10'2 >k° >10'6 cm/s) or irreversible (k0 < 10'6 cm/S). For app app a reversible system under control by semi-infinite linear diffusion, the peak current, ip, is given by the Randles-Sevcik equation, ip = (2.69x105)n3"2 A DJ,” v“ (7;. The electrode area, A (cmz), the diffusion coefficient for the oxidized species, Do (cm/s), the scan rate, v (V/s), and the bulk concentration, C; (moI/cms), yield current in amperes [25]. From this relationship, it is seen that the current varies linearly with the square root of the scan rate and the concentration. 65 Figure 3-2 Ferrocene Cyclic Voltammetric Response at a Boron-Doped Diamond Electrode (A) Cyclic voltammetric i-E curve for 0.1 mM ferrocene in acetonitrile containing 0.1 M TBAC/O4 at 0.2 V/s. The featureless response for the blank electrolyte solution is also Shown. (B) Oxidation peak current versus the square root of the scan rate plot for 0.1 mM ferrocene in acetonitrile containing 0.1 M TBAC/O4. Data are Shown for scan rates from 0.05 to 1.0 V/s. 15 40 j I A I —LtnearfIt 35» B 10» 3.: , .. 30. <3 5? E * P x L E 0 § 20. t3 . 5 15+» '5' E 10. '10)- E 5” 0T 0230.3 A 0.4‘0.5T 0.64 0.7 93.0 ‘ d2 ‘ 0T4 ‘ 0.6 ‘ ole L 11.0 Potential (V vs. Ag wire QRE) V10 ((V/s)”2) A well-defined voltammetric response was observed for ferrocene at BDD (Figure 3-2A) with a AEp of 0.074 V, an E1/2 of ca. 0.440 V, and a peak current ratio of 1. The oxidation peak current varied linearly with the square root of the scan rate between 0.05 and 1.0 V/s, as shown in Figure 3-28. The correlation coefficient was 0.999 and the y—axis intercept was near zero. This indicates the ferrocene oxidation reaction is limited by semi-infinite linear diffusion of the reactant to the interfacial reaction zone. The calculated rate constant is consistent with reversible electrochemical reaction kinetics. It can be seen that the background current is low and featureless over the potential range—a characteristic feature of diamond electrodes [1]. 66 Cyclic voltammetric AEp values were recorded at different scan rates, to determine the apparent heterogeneous electron transfer rate constant, kgpp, in the three different electrolyte/solvent systems. The AEp can be used to determine kgpp, using the classical treatment developed by Nicholson [22, 23]. This method relates AEp to the dimensionless kinetic parameter, lIt, and is useful for determining rate constants for redox systems with peak separations of 0.065 — 0.20 V. Nicholson generated a working curve demonstrating the variation of peak potential separation with the dimensionless kinetic parameter [22]. From the experimental AEp, the corresponding value of ti! is determined from the working curve. If the scan rate and diffusion coefficients are known, the rate constant can [.00] k, DR — ([0,, III/(itF/RT) ' be calculated from ’2” 0,22 In most cases, the assumption that [DO] is near unity is valid, with the R exception being cases of large differences between Do and DR, regardless of the value of the transfer coefficient, a. Table 3-1 lists the kgpp values calculated for both untreated diamond and freshly polished GC. One of the properties of high-quality diamond electrodes is the inherently active response for many redox systems, and long-term 67 maintenance of this response, without any conventional pretreatment. Triplicate results were obtained and averaged for five scan rates between 0.05 and 1.0 V/s. The k° app values were calculated using Nicholson’s method, assuming that O = 0.5. The diffusion coefficient for ferrocene in acetonitrile was taken as DR: 2.4 x 10'5 cm2 s'1 [26]. The value used for ferrocene in dichloromethane, DR = 2.0 x 10'5 cm2 S1, was calculated from the Stokes-Einstein relation, 6nr;r}=K—T, I) applying the Walden rule, I), r] = constant . It is assumed that the radius of the solvated ion, ri, does not change drastically in acetonitrile relative to dichloromethane [27]. The viscosities, n, used for acetonitrile and dichloromethane were 0.345 and 0.390 cP, respectively. 0 app values for diamond There are no striking differences between the k and those for GO in any of the three electrolyte/solvent systems. This suggests that there is no significant affect Of the electrode microstructure (spz- vs. sp3-bonding) on the electrode kinetics for this particular redox system. Therefore, the most important factor influencing the kinetics appears to be the density of electronic states, which is related to the number of charge carriers, at the formal potential. Clearly, both electrodes possess a sufficient density of states to support relatively rapid electron transfer kinetics. 68 Table 3—1 Apparent Heterogeneous Electron Transfer Rate Constants for Ferrocene at Untreated Boron-Doped Diamond and Polished Glassy Carbon Electrodes 0.1 mM Ferrocene BDD k3,, (cm/s) GC-30 kgpp (cm/s) 0.1M NaClO4/CH3CN 0.042 2‘. 0.015 0.059 : 0.014 0.1M TBAC|O4ICH3CN 0.048 i 0.015 0.051 i 0.003 0.1M TBACIO4/CH2CI2 0.008 t 0.002 0.006 i 0.001 BDD, Boron-Doped Diamond Electrode; GC-30, Tokai Glassy Carbon Electrode The rate constants reported herein can be compared to values reported in O app values reported herein for the literature for a variety of solid electrodes. The k diamond and glassy carbon are 5-10 times larger than some values reported in the literature [14, 17]. However, Panzer and Elving reported similar values for AEp at 0.1 V/s in 0.1 M LiCIO4/CH3CN on glassy carbon [18]. The k° app values reported for platinum are generally 2 -4 times larger than those observed for carbon electrodes. For example, the values reported in acetonitrile range from 0.02 to 0.22 cm/s [11, 13, 15, 16, 20, 21]. Much larger rate constants, up to 1 cm/s, have been observed for gold and platinum microelectrodes, using fast- scan cyclic voltammetry [28]. 0 app at either electrode There appears to be no electrolyte cation effect on k 0 app observed for both in CH3CN. There is, however, an effect of the solvent on k electrodes; with the kgppin CH2CI2 being a factor of 4-6 lower than in CH3CN. 69 The difference is attributed to a dielectric constant for CH2C|2 that is a factor of 4.3 less than that for CH3CN (£(20 CC) = 9.08 vs. 38.8). The difference in dielectric constant is slightly less than the difference in rate constants, which is a factor of 5 lower for diamond and a factor of 8.5 lower for glassy carbon in CH2CI2. The lower solvent dielectric constant can affect the CV response, from which k° app is determined, in at least two ways. First, the solution will have a lower ionic conductance because of less effective charge separation. This results in reduced ionic strength and, therefore, less ion conductivity. The lower ionic conductivity Increases the magnitude of iRu. In the experiment, the cell current, i, is measured as a function of Eappl, which is the applied potential difference between the working and reference electrodes. However, the true potential across the working electrode/solution interface, EW, is related to Eapp. by Ew = Euppl _ l.Ru _ (pref I where RU is the uncompensated portion of the solution resistance and (Ilref is the 0 reference electrode potential. The evaluation of AED, and its relation to kapp, becomes problematic if iRu becomes significant (2 mV or greater). In the present data analysis, 70% of the iRu was electronically compensated for, so solution resistance is expected to have little affect on the measured AEp values, or the related kgpp values. 70 Second, changing the dielectric constant of the solvent can strongly influence the electric double-layer structure—in particular, the magnitude of the potential drop at the plane of closest approach, the outer Helmholtz plane, OHP (Gouy-Chapman-Stern model) [25]. It is supposed that this is the main effect in the present data. A lower dielectric constant means there will be less accumulated ions at any given potential in the compact layer to compensate for the excess surface charge on the electrode. Thus, for the excess charge on the surface to be balanced by the charge in the solution, qm = -q5, there will have to be a significant diffuse-layer thickness. A larger diffuse-layer thickness means that the magnitude of the electric field at OHP, t’iCID/t5xx:x2 , will be less. This is the so-called Frumkin effect [25]. The potential at the OHP, (1)2, will not be equal to the potential in bulk solution, tits, because of the potential drop through the diffuse layer. Thus, the effective potential is Eapp. — (1)2. The kgpp values can be corrected for such a double-layer effect according to k” k” (an—z)F¢2 . = . . ex ti/rp true p RT where kf’me is the true heterogeneous electron transfer rate constant, a is the transfer coefficient, n is the number of electrons transferred per molecule, 2 is the charge on the molecule, and F, R, and T have their usual meanings. This correction is possible, if one can calculate (>2. However, one must know the potential at the point of zero charge, in order to calculate this value; and, for diamond, this potential is unknown. The fact that the relative rate constant 71 difference is larger for glassy carbon than for diamond may indicate that the overall structure of the electric double-layer of these two carbon electrodes is different. 3.3 Conclusions The electron transfer kinetics for ferrocene were investigated at microcrystalline, thin-film BDD electrodes as a function of the solvent and electrolyte composition. The background current was low and stable with time. Well-defined CV responses were observed in all three electrolyte/solvent systems, consistent with quasi-reversible electron transfer reaction kinetics. Measured values of k° were in the low to mid 10'2 cm/s range for the untreated app BDD thin-film electrodes, very similar to the values for freshly polished glassy carbon. Hence, there is no apparent effect of the electrode microstructure on the electrode kinetics for this redox system. The kinetics are primarily affected by the density of electronic states at the formal potential. Clearly, these two electrodes have a sufficient density of states to support relatively rapid kinetics. A comparison of the rates measured in dichloromethane and acetonitrile reveals the expected trend of decreased rates 0 app in the lower dielectric solvent, dichloromethane. The relatively large k observed for ferrocene at untreated BDD demonstrates the viability of using these electrodes for electrochemical analysis of nonaqueous redox systems and that the attractive features of this electrode material exhibited for aqueous redox systems are retained for nonaqueous redox systems. 72 3.4 [I] I21 I31 [41 [5-1 [5-1 [7-1 [81 [9-1 [10.] [11.] References M. C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M. D. Koppang, J. E. Butler, G. Lucazeau, M. Mermoux, J. W. Strojek and G. M. Swain Anal. Chem. 2000, 72, 3793-3804. T. A. Ivandini, B. V. Sarada, C. Terashima, T. N. Rao, D. A. Tryk, H. Ishiguro, Y. Kubota and A. Fujishima J. Electroanal. Chem. 2002, 521 , 117-126. M. Witek, J. Wang, J. Stotter, M. Hupert, S. Haymond, P. Sonthalia, G. M. Swain, J. K. Zak, Q. Chen, D. M. Gruen, J. E. Butler, K. Kobashi and T. Tachibana J. Wide Bandgap Materials 2002, 8, 171-188. G. M. Swain In Thin Film Diamond; C. Nebel and J. Ristein, Eds; Academic Press, 2002, in press. J. Wang, G. M. Swain, M. Mermoux, G. Lucazeau, J. Zak and J. W. Strojek New Diamond and Frontier Carbon Technology 1999, 9, 317-343. S. Haymond, G. T. Babcock and G. M. Swain Electroanal. 2002, in press. M. Yoshimura, K. Honda, T. Kondo, R. Uchikado, Y. Einaga, T. N. Rao, D. A. Tryk and A. Fujishima Diam. Relat. Mat. 2002, 11 , 67-74. Z. Y. Wu, T. Yano, D. A. Tryk, K. Hashimoto and A. Fujishima Chemistry Letters 1998, 503-504. S. Alehashem, F. Chambers, J. W. Strojek, G. M. Swain and R. Ramesham Anal. Chem. 1995, 67, 2812-2821. A. M. Bond, E. A. McLennan, R. S. Stojanovic and F. G. Thomas Anal. Chem. 1987, 59, 2853-2860. J. W. Diggle and A. J. Parker Electrochim. Acta 1973, 18, 975-979. 73 [12.] [13.] [14.] [15.] [16.] [17.] [18.] [19.] [20.] [21.] [22.] [23.] [24.] [25.] R. R. Gagne, C. A. Koval and G. C. Lisensky Inorg. Chem. 1980, 19, 2854-2855. N. R. Armstrong, R. K. Quinn and N. E. Vanderborgh J. Electrochem. Soc. 1976, 123, 646-649. L. Bjelica, R. Parsons and R. M. Reeves Proc. - Electrochem. Soc. 1980, 80-3, 190-212. T. Gennett, W. Geiger, B. Willett and F. C. Anson J. Electroanal. Chem. 1987, 222, 151-160. K. M. Kadish, J. Q. Ding and T. Malinski Anal. Chem. 1984, 56, 1741- 1744. M. Noel, V. Suryanarayanan and R. Santhanam Electroanal. 2000, 12, 1039-1045. R. E. Panzer and P. J. Elving J. Electrochem. Soc. 1972, 119, 864-874. I. Ruff, V. J. Friedrich, K. Demeter and K. Csillag J. Phys. Chem. 1971, 75, 3303-3309. M. Sharp, M. Petersson and K. Edstroem J. Electroanal. Chem. 1980, 109, 271-288. M. Sharp Electrochim. Acta 1983, 28, 301-308. R. S. Nicholson Anal. Chem. 1965, 37, 1351. R. S. Nicholson and I. Shain Anal. Chem. 1964, 36, 706. J. L. Anderson, L. A. Coury and J. Leddy Anal. Chem. 2000, 72, 4497- 4520. A. J. Bard and L. R. Faulkner Electrochemical Methods: Fundamentals and Applications; 2nd ed.; John Wiley 8 Sons, Inc., New York, 2001. 74 [26.] T. Kuwana, D. E. Bublitz and G. Hoh J. Am. Chem. Soc. 1960, 82, 5811- 5817. [27.] J. Koryta, J. Dvorak and L. Kavan Principles of Electrochemistry; 2nd ed.; John Wiley & Sons, New York, 1993. [28.] D. O. Wipf, E. W. Kristensen, M. R. Deakin and R. M. Wightman Anal. Chem. 1988, 60, 306-310. 75 CHAPTER 4 4 Direct Voltammetry of Cytochrome c at Boron-Doped Diamond Electrodes 4.1 Introduction Electron transfer reactions play an integral role in many biological processes, including photosynthesis and respiration. Additionally, many biosensor and biocatalytic devices are based on electron transfer processes involving biomolecules. Thus, the physiological relevance and exciting application possibilities have made direct bioelectrochemistry the subject of a number of recent reviews [1-3]. The practice of directly measuring biological electron transfer reactions spans over three decades, beginning with early work aimed at determining the reduction potentials of proteins via potentiometric methods [4]. From protein electrochemistry, one can Obtain valuable thermodynamic and kinetic information, which is useful when accompanied by detailed structural information available from crystallographic and spectroscopic methods. Biochemists and bioinorganic chemists routinely use cyclic voltammetry to determine the solution redox potentials of small proteins and to investigate the influence of site-directed mutations and experimental conditions on protein redox centers [5-7]. Electrochemists, however, are primarily concerned with using voltammetry to understand the interfacial interactions between proteins and electrode surfaces via electron transfer kinetics, adsorption kinetics, and thermodynamic measurements [1]. From the outset, the difficulties 76 of making direct electrochemical measurements of proteins were realized, as an early report of the direct reduction Of the electron transfer protein cytochrome c at a platinum electrode indicated that application of very negative potentials were necessary (La, -1.2 V) to observe reduction [8]. The field has progressed beyond the first measurements of cytochrome c to include direct methods that routinely observe quasi-reversible electron transfer for small electron transfer proteins, such as cytochrome c, the ferredoxins, and the blue-copper proteins, and has expanded to include the investigation of increasingly complex enzyme systems with multiple redox-active metal centers, such as cytochrome c oxidase. However, cytochrome c continues to be frequently used in the electrochemical investigation of new electrode materials and chemically modified surfaces and remains the subject of numerous bioelectrochemistry papers per year. Cytochrome c is a relatively small (12,400 Da), water-soluble protein that functions as an electron carrier in many biological reactions. The redox-active metal center consists of a Fe-containing heme unit, in which the Fe is ligated to a Met and His residue, as shown in Figure 4-1A. The structure of the heme 0 center is also shown in Figure 4-1, for clarity. The protein contains three conserved core helices which form a "basket" around the heme group with one heme edge exposed to the solvent [9]. Located near the heme edge is a phenylalanine residue, Phe82, which also serves as the exit point for a positive dipole moment. This may be the region involved in the closest approach of the heme edge to the electrode surface [10]. The redox reaction of cytochrome 0 occurs through a single, outer-sphere electron transfer step for which the formal 77 reduction potential of the heme Fe(l|l) to Fe(ll) transition is accepted to be 0.260 V (vs. NHE) [3]. Physiologically, electrostatic interactions involving several lysine residues on the surface of cytochrome c enable it to dock at its reaction partner protein, promoting the necessary redox center orientations required for electron transfer [11-15]. Electrochemists have used this knowledge to mimic the cytochrome c docking site at metal and chemically-modified electrode surfaces, in an effort to facilitate direct, rapid electron transfer between cytochrome c and the electrode surface [3, 16-21]. Figure 4-1 Cytochrome c Heme Redox Center (A) Heme coordination scheme for class l type cytochromes, in which the redox-active heme Fe is ligated to Methionine and Histidine amino acid residues. (B) Detailed schematic of the cytochrome c heme, heme c, structure. A B Cys\ S Met SvCYS \s | H I N 6 / HIS HN O OH OH O The cyclic voltammetric response for cytochrome c has been studied at a variety Of electrodes. It has been found to undergo rapid electron transfer with clean metal (e.g., Au, Pt, Ag), carbon, and metal oxide electrodes, as well as with chemically-modified metal electrodes [10, 22-25]. As a well-characterized, electron transfer protein, cytochrome c has been used extensively as a test 78 system for redox reactions of proteins. The electrode kinetics of this system are known to be strongly dependent upon a combination of the interfacial electrostatic and chemical interactions, which are derived from the enzyme structure and the nature of the electrode surface. Fouling of the electrode surface caused by protein denaturation and aggregation plagued early work focused on direct voltammetry Of cytochrome c at metal electrodes. However, it has been Since been shown that two factors are necessary to Observe relatively rapid electrode reaction kinetics for cytochrome c at bare electrodes: (i) a chromatographically purified cytochrome c solution (is, free from oligomeric contaminants) and (ii) electrode pretreatment to produce a clean, hydrophilic surface [10]. Armstrong and coworkers observed a marked difference in the electron transfer rate of cytochrome c at the edge vs. basal plane of pyrolytic graphite. The authors report a drastic increase in the kinetics at the polished “edge” plane, which has higher surface O/C ratio, than at the freshly cleaved, hydrophobic “basal” plane [26]. Thus, these results suggest that cytochrome c electron transfer at an uncharged, hydrophobic surface should be sluggish, if a response is observed at all. Herein, we report a somewhat different finding, in that we observe reversible, stable, diffusion-controlled kinetics at an uncharged, hydrogen- terminated electrode surface. We observed this behavior as part of our research work to develop boron-doped diamond (BDD) optically transparent electrodes (OTES) for use in infrared (IR) spectroelectrochemical studies to probe protein structure-function relationships. We began our protein study with cytochrome c, 79 as it is a structurally and electrochemically well-characterized heme protein that serves as a simple model for the more complicated heme-containing systems we intend to study. One of the first steps in our multi-tasked project was to characterize the electrochemical kinetics Of cytochrome c at an untreated BDD electrode. The spectroelectrochemistry of cytochrome c in the UV/vis and IR at BDD OTES will be described in a later chapter (Sections 7.2.4 and 7.3.5). BDD is a new electrode material, as discussed in Chapter 3 [27-30]. Good quality, hydrogen-terminated diamond thin-film electrodes exhibit properties well suited for protein electrochemical studies such as: (i) a low and stable background current over a wide potential range and (ii) a resistance to fouling, due to weak adsorption of polar molecules on the nonpolar surface [31]. This chapter will discuss investigations of the effects of the different surface morphology (micro- and nanocrystalline films), the effect of boron dopant level (1 and 10 ppm films), and the effect of surface pretreatment (acid washing and rehydrogenation) on cytochrome c electron transfer. BDD thin-film electrodes can be produced synthetically by microwave- assisted chemical vapor deposition. Alterations in the growth conditions result in different BDD thin-film morphologies, namely, micro- and nanocrystalline. Microcrystalline films are grown using a CH4/H2 (~ 0.5%) source gas mixture, with growth conditions such that the rate of crystal growth exceeds the rate of nucleation. This produces large (1 -20 um), well-faceted diamond grains. In contrast, the nanocrystalline films are grown using a CH4/Ar (~ 1%) source gas mixture with little or no H2 added [32]. Under these growth conditions, the rate of 80 nucleation generally exceeds the rate of crystal growth, resulting in the formation of small grains (~ 15 nm). Nanocrystalline films are very smooth, with a 30 - 50 nm surface roughness, and have a higher fraction of grain boundaries than do the microcrystalline films due to the smaller grain size [33, 34]. The instantaneous contact angle measured for water is routinely larger than 70.00, indicating a mildly hydrophobic surface. As deposited, the films are H-terminated with XPS data indicating surface oxygen levels Of less than 2 atomic percent. Electrical conductivity is imparted to both types of films by boron doping (~ 1019 - 1021 B cm'3) and both types of films exhibit excellent electrochemical responsiveness. We have exploited these properties and demonstrated that BDD is a viable material for studying bioelectrochemical redox reactions—in particular, the reduction and oxidation of cytochrome c [35]. Our preliminary results indicated that quasi-reversible electron transfer could be easily observed for cytochrome c at BDD thin-film electrodes. We then turned our focus towards determining the surface properties that would yield optimal and reproducible heterogeneous electron transfer rates for cytochrome c at these electrodes. 4.2 Results All electrochemical measurements were performed using a conventional three-electrode cell, as described in the experimental chapter, Section 2.3. Details of the BDD electrode fabrication (Sections 2.1.1-2.1.2) and the cytochrome 0 sample preparation procedure (Section 2.2.2) were also provided in Chapter 2. The 1 ppm and 10 ppm nomenclature used throughout refers to 81 the growth conditions of BDD thin films, specifically, to the gas phase concentration (B/C ratio). These values correspond to carrier concentrations of ca. 1018 B cm'3 (1 ppm) and ca. 1019 B cm'3 (10 ppm). 4.2.1 Microcrystalline BDD Thin-Film Electrodes The cyclic voltammetric i-E curves measured for cytochrome c at 1 ppm and 10 ppm microcrystalline BDD thin-film electrodes are shown in Figure 4-2. Despite the increased carrier concentration present in the 10 ppm film, there is no observable difference in the cytochrome c electron transfer rate constant at these electrodes. This was the exception, rather than the general observation for a variety of redox couples. Comparison measurements using redox analytes like Fe(CN)53'/4' and Ru(NH3)63+/2+ generally revealed larger cyclic voltammetric AEp values for the 1 ppm films. We believe this is due to increased film resistance. This is reflected by the fact that the apparent film resistivities were a factor of 10 larger for the 1 ppm films. These films are called “as deposited” because no surface pretreatment was applied after film deposition. Later studies revealed that rate constants measured for the 10 ppm films were indeed larger than those for 1 ppm films, so the duration of this section describes studies at 10 ppm microcrystalline BDD electrodes. 82 Figure 4-2 Cyclic Voltammetry of Cytochrome c at “As Deposited” 1 ppm and 10 ppm Microcrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i-E curves for 200 pM cytochrome c with 20 mM NaCl and 1 mM Tris buffer (pH 7) measured at “as deposited” 1 ppm (dotted line) and 10 ppm (solid line) microcrystalline BDD electrodes. The potential was initially scanned in the negative direction from 0.30 to -0. 20 V at 0.010 V/S. Electrode area = 0.2 cm2 ..... 1ppm —10ppm -02 A .01 l 0.0 A 0.1 ‘ 0.2 L 0.3 Potertial (V vs. Ag/AgCI) Cyclic voltammograms (CVs) for cytochrome c measured at four different 10 ppm microcrystalline BDD electrodes, from the same growth batch, are presented in Figure 4-3. In general, the electrode responsiveness of the “as deposited” 10 ppm films is very good for the aqueous-based redox test analytes, 3+/2+ Fe(CN)63'/4‘ and Ru(NH3)6 , as the CV peak splitting at 100 mV/s is routinely 75 mV or less. Figure 4-3 demonstrates that the response is sluggish for the protein. In contrast, Figure 4-4 shows cytochrome c CVs at the same four electrodes after a surface regeneration procedure. This procedure involves exposing the electrodes to a 2-step acid washing, followed by rehydrogenation in 83 H2 plasma. The details of the acid washing and rehydrogenation were given in Section 2.1.1. It is clear that the electron transfer kinetics and reproducibility greatly improve after the pretreatment. The pretreatment produces a clean, hydrogen-terminated surface. Figure 4-3 Cyclic Voltammetry of Cytochrome c at “As Deposited” Microcrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i-E curves for 200 pM cytochrome c with 50 mM NaCI and 1 mM Tris buffer (pH 7) measured at “as deposited” microcrystalline (10 ppm) BDD electrodes. Data are presented for four different electrodes from the same batch, as indicated by the numbers in the figure legend. The potential was initially scanned in the negative direction from 0. 30 to -0. 20 V at 0.020 V/S. Electrode area = 0.2 cm2. Current (uA) 4L L L 0.2 0.3 0.4 LL L 0. 1 Potential (V vs. Ag/AgCI) -02 A 0.1 A 00 The deposition Of the microcrystalline films involves placing four of the Si SLJbStrates in the microwave reactor, all positioned near the center of the $leStrate stage. The substrate temperature is controlled indirectly by immersion I“ the plasma. However, with four substrates present, it often results that the 84 temperature of one or more Of the substrates can be inhomogeneous across the surface. For example, a substrate placed near the edge of the plasma discharge zone will have the side exposed to the discharge at a higher temperature than the side further removed from the discharge zone. Such temperature gradients result in lower boron incorporation and more poorly conducting amorphous carbon impurity incorporation. This effect is manifested in irreproducible behavior, as shown in Figure 4-3. Figure 4-4 Cyclic Voltammetry of Cytochrome c at “Acid-Washed, Rehydrogenated” Microcrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i-E curves for a 200 )JM cytochrome c with 50 mM NaCl and 1 mM Tris buffer (pH 7) measured at acid-washed, rehydrogenated microcrystalline (10 ppm) BDD electrodes. Data are presented for four different electrodes, as indicated by the numbers in the figure legend. The potential was scanned in a negative direction from 0. 30 to —0. 20 V at 0.020 V/s Electrode area = 0.2 cm2. A l A 0.2 0.3 0.4 L -02 A -01 A 0.0 I 0.1 Potential (V vs. Ag/AgCI) 85 Similar improvements were observed for CVs measured using acid washed, rehydrogenated 1 ppm microcrystalline BDD thin-film electrodes over those measured at as deposited surfaces (data not shown). However, in general, the rate constant values for the 10 ppm films were larger than the 1 ppm films even after pretreatment. The reason for the response improvement is unclear. It is supposed that the pretreatment removes poorly conducting, amorphous carbon impurities from the surface, leading to larger electron transfer rate constants for this surface-sensitive analyte. For all of the remaining data presented in this section, the electrodes were acid washed and rehydrogenated. Figure 4-5 Demonstration of the Cyclic Voltammetric Response Stability for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode Cyclic voltammetric i-E curves for 100 pM solution of cytochrome c containing 50 mM NaCl and 1 mM Tris at pH 7. The potential was initially scanned negative/y from 0. 30 to -0. 20 V at 0. 020 V/s. The numbers in the legend represent the scan number. Data are presented for scans between 1 and 15 at an acid-washed, rehydrogenated microcrystalline (10 ppm) BDD electrode (referred to as #4 in Figures 4-3 and 4-4). Electrode area = 0.2 cm . CUTent (tIA) .01 A 0.0 A 0.1 A 0.2 A 0.3 A 0.4 Potential (v vs. Ag/AgCI) 86 The reproducibility and responsiveness of the electrodes was improved after pretreatment. For the voltammograms presented in Figure 4-4, the applied potential was swept from 0.30 V to —0.20 V at a scan rate of 0.020 V/s. The reduction peak at ca. 0.026 V is indicative of the heme Fe(lll) to Fe(ll) transition. Beyond —0.10 V the reduction current is attributed to the reduction of residual dissolved oxygen. The catalysis of oxygen reduction at the electrode surface by metal macrocycles, such as Fe porphyrins, is well known [36]. Furthermore, experiments involving added oxygen to the solution resulted in similar currents. On the reverse scan, the peak near 0.133 V is due to the oxidation of heme Fe(ll) to Fe(lll). The short-term response stability for cytochrome c at bare electrodes has been shown to be a sensitive test of protein sample purity [10]. The CV signals of cytochrome c samples that are not chromatographically purified and, therefore, contain oligomers, typically decrease over time with increasing scan number. Niki and coworkers have concluded that a decreased signal amplitude over time indicates fouling by denatured protein blocking the electroactive area of the electrode [37]. Therefore, Figure 4-5 demonstrates an important point for cytochrome c electrochemistry at BDD electrodes, and that is there is no observed decrease in signal amplitude or increase in AEp over several minutes with repeated scans. The protein used in all CV experiments was chromatographically purified. This suggests that the interaction of cytochrome c with the H-terminated BDD electrode surface does not result in irreversible adsorption and denaturation of 87 the protein molecules. The fact that the redox potential is in good agreement with that observed for cytochrome c in solution further supports the interpretation that the protein is In its native form. Generally, electrode fouling at metal and metal oxide electrodes is evident within 10 scans at 0.02 V/S (~ 10 min) [2]. Figure 4-6 Cyclic Voltammetric Response for Cytochrome c Measured at a Microcrystalline Boron-Doped Diamond Electrode (A) Background response for 50 mM NaCl, 1 mM Tris HCI at 0. 020 V/s. (B) Cyclic voltammograms for 50 pM horse heart cytochrome c in 50 mM NaCl, 1 mM Tris HCI pH 7 buffer at an acid-washed, rehydrogenated microcrystalline (10 ppm) BDD electrode (referred to as #4 in Figures 4-3 and 4-4). Data are presented for scan rates 0. 002 - 0.100 V/s. Electrode area = 0.2 cm2. . A *___.J l B A 1.0 - 530.5. 2 ’ A 5 0.0 I .N —0.5‘ -1.0 - -1.5 . L on A 0.0 0.1 A 0.2 A 0.3 Potertial (V vs. Ag/AgCl) Figure 468 shows a series of cyclic voltammetric i-E curves for cytochrome c at a 10 ppm microcrystalline BDD electrode at different scan rates from 0.002 to 0.100 V/s. The solution concentration was 200 [M in 1 mM Tris HCI buffer (pH 7) containing 50 mM NaCI. A well-defined voltammetric 88 response, characteristic of a diffusion-controlled reaction, is observed in the potential range of 0.30 to —0.20 V. This response was observed immediately upon introducing cytochrome c solution into the electrochemical cell with no conditioning cycles required. The AEp values ranged from 0.058 to 0.091 V over this scan rate range, indicating quasi-reversible behavior. The background scan (50 mM NaCI solution) collected after the CV measurements of cytochrome c shown in Figure 4-6A was featureless, suggesting no protein adsorption. The apparent heterogeneous rate constant, k was calculated from the scan rate dependence of the peak splitting via the 0 399' Nicholson method, using the data presented in Figure 4-6B [38]. The value obtained from this method, 7.2 x 10'3(: 2) cm/s, compares well with the rate constant reported for other bare electrodes [24]. The standard reduction potential, measured, as the average of the anodic and cathodic peak potentials, remained independent of scan rate over this range. The value of 0.073 V vs. Ag/AgCl (0.270 V vs. NHE) is in very good agreement (1 10 mV) with that reported in the literature for cytochrome c in solution at this pH [39]. The anodic and cathodic peak currents are shown to vary linearly with the square root of the scan rate (Figure 4-7) and the concentration (Figure 4-8), as expected for a one-electron diffusion-controlled process. The diffusion coefficient for the oxidized form of cytochrome c was calculated from the slopes of the plots in Figure 4-8 and found to be .7 x 10’7 cm2/s. The value measured here is in agreement with that reported by Armstrong and coworkers (5.0x 10‘7 cm2/s) from CV studies of cytochrome c at the edge plane of pyrolytic graphite (PG) 89 electrodes [26]. These rates are low compared to the solution value of 11 x 10'7 cmzls, owing to the quasi-reversible nature of the cytochrome c electron transfer reaction at solid electrodes. Figure 4-7 Peak Current Dependence on the Square Root of the Scan Rate Measured for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode Plot of the reduction (squares) and oxidation (circles) peak currents versus the square root of the scan rate. Data points were obtained from the CVS presented in Figure 4-6. Data shown for scan rates from 0. 002 to 0.100 V/s. Linear correlation constants = 0. 999 (oxidation) and 0.998 (reduction) and y-intercepts = 0.010 pA (oxidation) and - 0.010 trA (reduction). Electrode area = 0.2 cm . 0.9 0.6 - 0.3 - 0.0 - -0.3 - Peak Current (HA) -06 _ I 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 San Rate"2 ((V/s)‘0) I l - J 90 Figure 4-8 Peak Current Dependence on the Concentration Measured for Cytochrome c at a Microcrystalline Boron-Doped Diamond Electrode Plot of the peak current versus the cytochrome 0 solution (50 mM NaCl and 1 mM Tris at pH 7) concentration. Data points obtained from cyclic voltammograms measured at 0.020 V/s at an acid-washed, rehydrogenated microcrystalline (10 ppm) BDD electrode (referred to as #4 in Figures 4-3 and 4-4). Data is shown for concentrations from 25 to 200 pM. Linear correlation constants = 0.999 (oxidation) and 0.994 (reduction) and y-intercepts = 0.085 pA (oxidation) and - 0.069 pA (reduction). Electrode area = 0.2 cm2. 1.2 I red 09- . 0x 0.6 - 0.3 - 0.0 - ak Current (IA) -0.3 - Pe 95 a: -0.9 - -12 - 0 25 50 75 100 125 150 175 200 225 Concentration (ttIVI) Although no evidence of strong adsorption was Observed for cytochrome c electron transfer at BDD thin-film electrodes, changes in the electrode surface termination were evident after exposure to cytochrome c. In particular, the background CV after exposure to cytochrome 0 showed a marked reduction current beyond -0.10 V, as is shown in Figure 4-9. As mentioned earlier, surface Oxygen reduction catalysis by porphyrins is well documented. Additionally, Similar current profiles were observed during experiments involving added Oxygen to the electrolyte solution. 91 Figure 4-9 Cyclic Voltammetric Background Curves Measured for a Boron- Doped Diamond Electrode after Exposure to Cytochrome c Background cyclic voltammograms of the electrolyte solution (50 mM NaCl and 1 mM Tris at pH 7) before (solid line) and after (dashed line) cytochrome 0 measurements. Cyclic voltammograms measured at 0.020 V/s at an acid-washed, rehydrogenated microcrystalline (10 ppm) BDD electrode. -0.3 A -02 A .01 A 0.0 A 0.1 A 0.2 A 0.3 Potential (v vs. AglAgCI) The indication of catalytic oxygen reduction coupled with increased surface N levels, as measured by XPS, after the cytochrome 0 measurements suggests the presence of heme molecules on the surface. As shown in Table 4-1, the peak splitting observed for the surface-sensitive test analyte, Fe(CN)53’/4‘ increased after exposure to the protein solution. Granger et al demonstrated that the electron transfer kinetics for this analyte at BDD electrodes is highly sensitive to the surface termination [31]. The lowest peak splitting, (i.e., fastest rates) were Observed for clean, hydrogen-terminated 92 surfaces. Thus, we would expect to see a decrease in the electron transfer rate of Fe(CN)53’/4’ if residual heme molecules, or oxygen groups are present at the 2+ electrode surface. In contrast, Ru(NH3)53+/ is not sensitive to the surface conditions, thus the peak splitting values measured before and after exposure to the protein solution are relatively constant. Table 4—1 Effect of Cytochrome c Exposure on Test Analyte Cyclic Voltammetric Peak Splitting for Acid-Washed Rehydrogenated 10 ppm Microcrystalline Boron-Doped Diamond Electrodes RU(NI‘I3)53+/2+ F8(CN)63A/4A electrode before after before after 1 69 71 80 129 2 69 70 73 135 3 69 73 74 145 4 67 71 73 151 Values measured for 0.1 mM analyte In 1 M KCI at 0.05 V/s In general, the electrochemical and spectroelectrochemical results obtained at BDD thin-film electrodes do not indicate denaturation of the cytochrome c protein upon exposure to this surface. Surface Fe was not detected by XPS, although if present, the levels may be well below that of the detection limit. However, it is well known that cytochrome c adsorbs strongly to glass and can only be removed by acid. The measurements following the cytochrome c electrochemistry were not performed after washing the glassware in acid, thus, the possibility for residual protein contaminating these experiments 93 can not be ruled out, at this time. Further experiments to investigate this issue and the role Of adsorption are required. 4.2.2 Nanocrystalline BDD Thin-Film Electrodes Figure 4-10 Cyclic Voltammetry of Cytochrome c at “As Deposited” 1 ppm and 10 ppm Nanocrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i—E curves for 100 pM cytochrome c with 50 mM NaCl and 1 mM Tris buffer (pH 7) measured at “as deposited” 1 ppm (dotted line) and 10 ppm (solid line) nanocrystalline BDD electrodes. The potential was scanned in a negative direction from 0. 30 to —-0. 20 V at 0.010 V/s. Electrode area = 0.2 cm2. ..... 1ppm —-10F>Pm oz -0.1 A 0.0 0.1 A 0.2 A 0.3 Potential(V vs. Ag/AgCI) 41% A comparison of the cyclic voltammetric i-E curves measured for cytochrome c at 1 ppm and 10 ppm as deposited nanocrystalline BDD thin-film electrodes is shown in Figure 4-10. In contrast to the cytochrome c voltammetry observed at the microcrystalline BDD electrodes represented in Figure 4-2, the rate constant measured for the 10 ppm nanocrystalline film is clearly larger than 94 that for the 1 ppm film. Again, the 10 ppm films have a higher doping level, a greater electron carrier concentration, and, thus, a higher conductivity. A comparison of Figures 4-2 and 4-10 suggests that the as deposited nanocrystalline BDD surface is better suited for promoting relatively rapid cytochrome c electron transfer kinetics. Owing to the increased rate Observed for the 10 ppm films, all subsequent studies were carried out using these electrodes. Figure 4-11 Cyclic Voltammetry of Cytochrome c at “As Deposited” Nanocrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i-E curves for 100 pM cytochrome c with 50 mM NaCl and 1 mM Tris buffer (pH 7) measured at “as deposited” nanocrystalline (10 ppm) BDD electrodes. Data presented for three different electrodes, as indicated by the numbers in the figure legend. The potential was initially scanned in the negative direction from 0. 30 to —0. 20 V at 0.020 V/S. Electrode area = 0.2 cm2. Current (uA) -0.1 0.0 A 0.1 A 0.2 A 0.3 Potential (V vs. Ag/AgCI) In the cyclic voltammograms presented in Figures 4-11 through 4-13, the applied potential was swept from 0.30 V to —0.20 V. Figure 4-11 shows a set of CVS that were collected using three different, “as deposited,” electrodes. The 95 el ('1‘! ‘4 h I: electrodes were grown under identical conditions in three separate runs. Therefore, the surface temperature gradient problem observed for the microcrystalline electrodes was nonexistent. The electron transfer kinetics observed for electrodes 2 and 3 are faster than those observed for electrode 1. In comparison, the cytochrome c CVs presented in Figure 4-12 were measured at the same three electrodes after acid-washing and rehydrogenating each electrode. The pretreatment again improved the reproducibility and decreased the CV peak splitting. The improvement was most dramatic for electrode 1. Figure 4-12 Cyclic Voltammetry of Cytochrome c at “Acid-Washed, Rehydrogenated” Nanocrystalline Boron-Doped Diamond Electrodes Cyclic voltammetric i-E curves for a 100 pM cytochrome c with 50 mM NaCl and 1 mM Tris buffer (pH 7) measured at acid-washed, rehydrogenated nanocrystalline (10 ppm) BDD electrodes. Data presented for three different electrodes, as indicated by the numbers in the figure legend. The potential was initially scanned in the negative direction from 0. 30 to —0. 20 V at 0. 020 V/s. Electrode area = 0.2 cm2. A A 01 A 00 0.1 A 0.2 A 0.3 Potential (V vs. Ag/AgCl) 96 I.A "it. Successive cytochrome c CV scans are presented in Figure 4-13 for one of the acid-washed, rehydrogenated nanocrystalline BDD thin-film electrodes, referred to as electrode 3 in Figures 4-11 and 4-12. As noted above for microcrystalline BDD thin-film electrodes, Figure 4-13 demonstrates short-term signal stability over several minutes with repeated scans. Therefore, it is concluded that like the microcrystalline surface, the nanocrystalline surface morphology does not promote cytochrome c denaturation, such that response attenuation or increased peak splitting were observed. Figure 4-13 Demonstration of the Cyclic Voltammetric Response Stability for Cytochrome c at a Nanocrystalline Boron-Doped Diamond Electrode Cyclic voltammetric i-E curves for 100 pM solution of cytochrome 0 containing 50 mM NaCl and 1 mM Tris at pH 7. The potential was initially scanned negatively from 0. 30 to -0. 20 V at 0. 020 V/s. The numbers in the legend represent the scan number; data are presented for scan numbers between 1and 10 at an acid-washed, rehydrogenated nanocrystalline (10 ppm) BDD electrode (referred to as #3 in Figures 4-10 and 4-11). Electrode area = 0.2 cm . L 0.2 0.3 Potential (V vs. Ag/AgCI) 0.1 A 0.0 A 0.1 97 Figure 4-14 Cyclic Voltammetric Response for Cytochrome c Measured at a Nanocrystalline Boron-Doped Diamond Electrode (A) Background CV at 0.020 V/s for a 50 mM NaCl, 1 mM Tris HCI aqueous solution and (B) CVS for 50 pM horse heart cytochrome c in 50 mM NaCl, 1 mM Tris HCI pH 7 buffer at a 10 ppm nanocrystalline BDD electrode. Data are presented for scan rates from 0. 002 - 0.100 V/s. Electrode area = 0.2 cm2. -0.1 0.0 0.1 0.2 0.3 Potential (V vs. AglAgCl) Figure 4-14B shows a series of cyclic voltammetric i-E curves for cytochrome c at a 10 ppm nanocrystalline BDD electrode at different scan rates from 0.002 to 0.050 V/s. The solution concentration was 50 [M in 1 mM Tris HCI buffer (pH 7) containing 50 mM NaCl. The CV peak splitting ranged from 0.066 to 0.120 V over this scan rate range, indicating reversible electron transfer kinetics. A well-defined voltammetric response, characteristic of a diffusion- controlled reaction, is observed in the potential range from 0.30 to —0.20 V. This response was observed immediately upon introducing cytochrome 0 solution into the electrochemical cell, and no time-dependent changes in the response were 98 observed. Although the data presented herein were Obtained at acid-washed, rehydrogenated electrodes, we have previously demonstrated that it is possible to measure quasi-reversible kinetics for cytochrome c at as deposited 10 ppm nanocrystalline electrodes, with no extensive cycling required to activate or precondition the electrode surface [35]. A background scan for a 50 mM NaCI solution at 0.020 V/s is also shown in Figure 4-14A. This scan was obtained after the cyclic voltammetric measurements of cytochrome c were performed. It is important to note that, even after exposure to the protein solution, the background current profile is featureless and the current magnitude remains the same as that prior to exposure to the protein. This suggests that irreversible adsorption of the protein, either in the active or denatured state, is not occurring to any quantifiable extent. The apparent heterogeneous rate constant, kgpp, was calculated from the scan rate dependence of the peak splitting via the Nicholson method, using the data presented in Figure 4-14B [38]. The value obtained from this method, 4.2 x 10’3 (i 1) cm/s, is in accordance with the rate constant reported for other bare electrodes [24]. The standard reduction potential, measured, as the average of the anodic and cathodic peak potentials, remained independent of scan rate over this range. The value of 0.080 V vs. Ag/AgCI (0.277 V vs. NHE) is in very good agreement (i 10 mV) with that reported in the literature for cytochrome c in solution [39]. 99 Figure 4-15 Reduction Peak Current Dependence on the Square Root of the Scan Rate Measured for Cytochrome c at a Nanocrystalline Boron-Doped Diamond Electrode Plot of the reduction (squares) and oxidation (circles) peak currents versus the square root of the scan rate. Data points were obtained from the CVS presented in Figure 4-14. Data are Shown for scan rates from 0. 002 to 0.100 V/s. Linear correlation constants = 0.997 (oxidation) and 0.999 (reduction) and y-intercepts = 0.011 pA (oxidation) and - 0.013 pA (reduction). Electrode area = 0.2 cm2. .0 (0 ,Ired Cox .0 05 r .0 O.) T Peak Current 04A) $5 0 OJ 0 9'3 O) r '09 4 I L L l . l . 1 . L L 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 Scan Rate” ((V/s)"2) 4L The anodic and cathodic peak currents are shown to vary linearly with the square root of the scan rate (Figure 4-15) and the concentration (Figure 4-16), as expected for a diffusion-controlled process. The ratio of the reduction and oxidation peak currents was 1. The diffusion coefficient for oxidized form of cytochrome c was calculated from the slopes of the plots in Figure 4-16 and found to be 4.5 x 10'7 cm2/s. 100 Figure 4-16 Reduction Peak Current Dependence on the Concentration Measured for Cytochrome c at a Nanocrystalline Boron-Doped Diamond Electrode Plot of the peak current versus the cytochrome 0 solution (50 mM NaCl and 1 mM Tris at pH 7) concentration. Data points obtained from cyclic voltammograms measured at 0. 020 V/s for an acid-washed, rehydrogenated nanocrystalline (10 ppm) BDD electrode (referred to as #3 in Figures 4-11 and 4-12). Data are shown for concentrations from 25 to 200 pM. Linear correlation constants = 0. 999 (oxidation) and 0. 998 (reduction) and y-inztercepts = 0.045 pA (oxidation) and - 0.057 pA (reduction). Electrode area = 0.2 cm. 1.2 0-9 - C ox 0.6 b 0.3 - I 0.0 - -0.3 - Peak Cunent -0.6 b -0.9 E l 0 A25 A so A 75 A100A125A150A175A200A225 [wtochromec](uM) -1.2 As was presented for the microcrystalline case, similar increases in background current and test analyte peak splitting were observed after the cytochrome c electrochemistry measured for BDD nanocrystalline thin-film electrodes. Table 4-2 lists the effects of the protein solution on the CV peak splitting of Fe(CN)53'/4' and Ru(NH3)63+/2+. As discussed above, protein contamination can not be ruled out at this time. 101 Table 4—2 Effect of Cytochrome c Exposure on Test Analyte Cyclic Voltammetric Peak Splitting for Acid-Washed Rehydrogenated 10 ppm Nanocrystalline Boron-Doped Diamond Electrodes Ru(NH3)63*’2” Fe(CN)63"4‘ electrode before after before after 1 72 81 81 168 2 92 141 93 156 3 81 1 19 88 185 Values measured for 0.1 mM analyte in 1 M KCI at 0.05 V/s 4.3 Discussion The increased electrode kinetics for the 10 ppm BDD thin-film electrodes compared to that of the 1 ppm electrodes is not unexpected. The higher gas phase diborane concentration leads to a higher film doping level, or more specifically, a higher carrier concentration. The carrier concentrations at room 3, respectively, for the temperature have been found to be ~ 1019 and ~1020 B cm' 1 and 10 ppm films. Thus, a higher B doping level yields more conductive films, less deformation of the cyclic voltammetric AEp values, and higher calculated electron transfer rate constants. The improvement in the response observed after acid-washing and rehydrogenation of both the microcrystalline and nanocrystalline BDD electrodes suggests that there are low levels of contaminants on the surface that are removed by the treatment. The identity of the contaminants is unknown, but they are likely poorly conducting, amorphous nondiamond (spZ-bonded) carbon. The surface contaminants can increase the ohmic resistance at the electrode surface 102 and/or alter the protein-electrode interaction by increasing the electron tunneling distance, which decreases the electrode kinetics. It is important to point out that for some as deposited 10 ppm nanocrystalline BDD electrodes relatively rapid kinetics were observed, with no extensive cycling required to activate or precondition the surface, and no need for extended equilibrium with the protein solution. It has been shown that reversible kinetics can be achieved for chromatographically purified cytochrome 0 solutions at bare metal electrodes only after pretreatment including polishing followed by sonication or annealing in a hydrogen flame [24]. Without special cleaning procedures, quasi-reversible cytochrome c voltammetry has been measured at polished gold and glassy carbon electrodes only after long equilibration times [40]. The cleaning procedures seem to serve two purposes: (i) to remove surface contaminants that may promote protein adsorption and denaturation, and (ii) to produce a hydrophilic surface [10, 23]. In this work, we focused on two dopant levels, 1 and 10 ppm, which 3 correspond to ~ 1019 and 1020 B cm’ , respectively. At these levels, the BDD thin films act like a semi-metal rather than a semiconductor, enabling rapid electron transfer kinetics over a wide (~ 3 V) potential range [27, 41, 42]. Therefore, it is possible to compare the rates observed for BDD thin-film electrodes to those measured for other solid metal electrodes, such as glassy carbon (GC), Au, Ag, and Pt. Table 4-3 presents a comparison of the heterogeneous electron transfer rate constants measured by direct voltammetry of cytochrome c. All reports of 103 cytochrome c kinetics demonstrate a strong dependence on cytochromec concentration and scan rate range, thus these are included in Table 4-3. The rate constant for the 10 ppm microcrystalline BDD thin-film electrodes increases by an order of magnitude after acid washing and rehydrogenation. As deposited, 10 ppm nanocrystalline films demonstrate the same rate constants as observed for the acid-washed and rehydrogenated 10 ppm microcrystalline electrodes. Thus, there appears to be little influence of the surface morphology and microstructure on the responsiveness of diamond electrodes as long as the electrode is conductive and the surface is clean. Upon acid washing and rehydrogenation, the rate constant for the films shows a marked increase over that for the 1 ppm films. Finally, it is clear that the rates measured for the best BDD electrodes (10 ppm acid washed and rehydrogenated micro- and nanocrystalline BDD) compare well with the observed rates for Ag, Au, and GC. The values measured for the PG edge plane are among the fastest of all solid electrodes. Although the rate constants for BDD are somewhat less than those observed at the hydrophilic edge plane of PG, it is crucial to note that they are faster than the rate constants observed for the hydrophobic basal plane of PG. This is surprising, considering that the conventional wisdom suggests a negatively charged, oxygen-rich, hydrophilic surface is necessary to observe facile electron transfer kinetics of cytochrome 0. Thus, the results presented herein may have implications for understanding the nature of the required surface composition for direct cytochrome c electron transfer. 104 Table 4—3 Apparent Heterogeneous Electron Transfer Rate Constants Measured from the Direct Voltammetry of Cytochrome c at Solid Electrodes kgpp x 103 [cm] 3:er Electrode (cm/s) (pM) (mVIs) Source Au 1.6 (:l: 0.9) 134 4 - 100 [24] Pt 2.8 (i 0.5) 85 8 - 200 [24] Ag 1.5 (.t 0.4) 200 1 - 10 [24] ln02 6.0 52 10 - 500 [22] GO 6.5 40 1 - 400 [43] PG (edge) 4.5 150 2 - 100 [26] PG (basal) 0.25 150 2 — 100 [26] 1 ppm micro BDD (as dep) 0.24 (:t 0.03) 200 2 - 20 this work 1 ppm nano BDD (as dep) 0.23 (.t 0.08) 100 2 - 20 this work 10 ppm micro BDD (as dep) 0.27 (i 0.06) 200 2 - 20 this work 10 ppm nano BDD (as dep) 0.86 (i 0.1) 100 2 - 100 this work 10 ppm nano BDD (as dep) 1.1 (.t 0.2) 200 2 - 50 [35] 10 ppm micro BDD (awrh) 7.2 (i 2) 50 2 - 100 this work 10 ppm micro BDD (awrh) 1.4 (:t 0.3) 200 2 - 100 this work 10 ppm nano BDD (awrh) 4.2 (1: 1) 50 2 - 100 this work 10 ppm nano BDD (awrh) 1.3 (i 0.4) 200 2 - 100 this work cyt c, cytochrome c; anz, indium oxide; GC, glassy carbon; PG, pyrolytic graphite; micro, microcrystalline; nano, nanocrystalline; as dep, as deposited; awrh, acid-washed and rehydrogenated. 105 In all cases, both the microcrystalline and nanocrystalline films exhibited a CV peak current response that increased linearly with (i) the square root of the scan rate, suggesting that the electrode reaction kinetics are limited by semi- infinite diffusion of the protein to the interfacial reaction zone, and (ii) concentration. In the case of cytochrome c heterogeneous electron transfer at bare electrodes, researchers have found that two parameters are necessary for measuring facile electron transfer kinetics at carbon, metal, and metal oxide electrodes: (i) use of chromatographically purified, non-Iyophilized cytochrome c (i.e., free from oligomeric contaminants) and (ii) electrode treatment to produce a hydrophilic surface [23, 26]. However, the factors that determine electron transfer rates at such protein/electrode interfaces are not fully understood. Comparisons of the rates observed for differently-sized hemeproteins (cytochrome c [23, 26], cytochrome c3 [44], cytochrome c553 [45], and myoglobin [46]) suggest that the size of the biological molecule, thus, the distance of closest approach between the heme edge and the electrode surface, is responsible for controlling the rate of heterogeneous electron transfer at bare electrodes. This idea is further supported by direct CV studies of cytochrome c on SAM/electrode surfaces, which show that rates depend exponentially on the chain length (longer chain length, slower rates), consistent with a through-bond tunneling mechanism of electron transfer [47]. It has been proposed that this state of closest approach is achieved by a combination of electrostatic and chemical interactions at the electrode/solution interface [10, 23]. Whether the 106 prevailing effects are chemical or electrostatic in nature depends upon the particular electrode/solution system under consideration. The fact that facile cytochrome c electron transfer kinetics have been observed for a number of different electrode surfaces raises questions regarding the existence of or requirement for a particular electrostatic orientation of cytochrome c at an electrode. Instead, it has been suggested that the point of zero charge (pzc) of an electrode in a given solution may be a key parameter for describing the varying voltammetric responses observed at different metal electrodes [10, 23]. Ignoring purely chemical interactions, at electrode potentials negative of the pzc, excess negative surface charge should interact attractively with the positively-charge exposed heme region of cytochrome 0, thus enhancing electron transfer. Too strong of an interaction would inhibit electron transfer due to slow removal of the protein molecules following electron transfer. Excess positive charge will be present at electrode potentials positive of the pzc, which should interact repulsively with the heme site, decreasing the electron transfer rate. Arguments for dominant electrostatic effects are based on the well- established idea that electrostatic interactions, via lysine residues on the cytochrome 0 surface, play an integral role for the binding of cytochrome c to its natural reaction partners, cytochrome c oxidase, cytochrome c reductase, and cytochrome c perioxidase [9]. These strategically-placed lysine residues form a ring-like pattern around the exposed heme edge and match the arrangements of the negative charges of the partner proteins. Chemical modification or site- 107 directed mutagenesis efforts to neutralize the positions of these lysines showed substantial reduction in the enzymatic activity of the partner proteins toward cytochrome c [48-50]. Thus, it is generally accepted that a specific arrangement of complementary charges in the protein-protein interface is essential for the proteins in the reactive complex to be properly aligned. This alignment ensures that the redox centers are oriented in positions most favorable for intermolecular electron transfer. In addition, studies have shown that the involvement of these lysine residues in the interactions with cytochrome c oxidase are also important for inducing the functionally relevant structural changes in cytochrome c [51]. Cytochrome c undergoes the same structural changes when bound to anionic surfaces exhibiting a largely uniform distribution of negative charges, such as phospholipid vesicles, micelles, polyanions, and electrodes (bare and chemically modified) [51-54]. These studies, combined with the findings from studies of cytochrome c/lipid interactions, clearly suggest that Coulombic forces represent the dominant mode of interaction between cytochrome c and its reaction partners as well as with biological membranes. However, there also exists a body of literature that indicates the necessity of hydrophobic interactions in the formation of the cytochrome c/cytochrome c oxidase electron transfer complex. Site-directed mutagenesis studies of Paracoccus denitrificans cytochrome c oxidase show that the Trp residue at position 121 is crucial for cytochrome c docking, implying that hydrophobic residues of cytochrome 0 must also be involved in the complex formation [55]. Hildebrandt’s group has carried out surface enhanced resonance Raman (SERR) spectroscopy of cytochrome c 108 immobilized on Ag/SAM electrodes in an effort to monitor the heme conformational changes accompanying the electron transfer. They studied the electrode response for both carboxyl— [56] and alkane-terminated SAM electrodes [57]. The SERR results indicate that cytochrome c can be immobilized onto hydrophobic surfaces and that the observed heme structure, in this case, is the same as for the electrostatically-bound case [57]. The authors conclude that the driving force for cytochrome 0 binding is likely the reduction of the solvent-exposed hydrophobic area. Additionally, hydrophobic interactions have been deemed important in computational studies of the cytochrome c/cytochrome c oxidase docking interface. Combining computational technology with information from static protein crystallographic structures, Roberts and Pique were able to model the orientation of this electron transfer complex [13]. Their work resulted in an optimized configuration that placed a cytochrome c oxidase Trp residue (Trp 143, Rhodobacter Sphaeroides; Trp 121, Paracoccus denitn'ficans) within 4 A of the cytochrome c heme edge, suggesting a possible electron transfer pathway, further supporting the idea that this Trp is the point of electron entry [11, 12, 55]. This interface consisted of a relatively hydrophobic contact region, which, in turn, was surrounded by hydrophilic, electrostatic interactions between complementary Lys and Asp residues. Bowden’s group has shown that cytochrome c electron transfer kinetics measured at mixed-monolayer SAM/electrodes, consisting of equimolar amounts of SH(CH2)1OCOOH and SH(CH2)30H, resulted in dramatically increased (e.g., 109 5-fold for horse heart cytochrome 0) rates over those measured at pure HS(CH2)1OCOOH SAM/electrodes [20, 47]. The authors concluded that the more irregular, protein-like surfaces, provided by mixed component SAMs, appear to be better suited for binding electron transfer proteins with optimal electronic coupling. The pzc for BDD is not currently known. BDD is known to possess the following surface characteristics: (i) hydrogen surface termination, (ii) hydrophobicity, (iii) absence of negatively charged (acidic) surface oxygen groups, and (iv) no extended rr-bonded electrons. Therefore, at this point we can only conclude that interactions other than electrostatic coupling between positively charged residues on cytochrome c and negatively charged electrode surface oxygen species are involved in promoting facile electron transfer at BDD electrode surfaces. Interactions between positively charged cytochrome c and a negative excess surface charge at BDD cannot be ruled out. Hydrophobic interactions likely play a key role in the observed voltammetric response for cytochrome c electron transfer at BDD electrodes. It seems that cytochrome 0 adsorption cannot be ruled out, because there is an observed effect of surface structure on the voltammetric response at BDD thin-film electrodes. However, if adsorption is occurring, it likely involves weakly bound cytochrome 0 molecules, as rinsing and successive measurement results in a featureless background signal. This is in contrast to the findings of 82003 and Novak, who present evidence for strongly adsorbed cytochrome c monolayers being involved in the electron transfer mechanism at glassy carbon 110 and Au electrodes [40]. Accordingly, further studies, namely, potential step chronocoulometry and long optical pathlength thin-layer spectroelectrochemical experiments, are required to determine if a similar process occurs at BDD thin-film electrodes. 4.4 Conclusions The reported results demonstrate that reversible, diffusion-controlled electron transfer kinetics for cytochrome c are observed at microcrystalline and nanocrystalline BDD thin-film electrodes without the requirement of long equilibration times or potential cycling. The response for both microcrystalline and nanocrystalline BDD electrodes was shown to improve after acid washing and rehydrogenation, presumably due to a surface cleaning effect. In both cases, the H-terminated electrode response is stable over time with no indication of fouling. This is an interesting result, considering that previous measurements of cytochrome c electrochemical kinetics at bare and modified electrodes have indicated the necessity of a hydrophilic, negatively charged, and oxygen-rich electrode surface. The fastest heterogeneous electron transfer rates were measured for 10 ppm BDD thin-film electrodes and are in accordance with that published for other solid electrodes. Further studies are required to determine if surface adsorption is involved in the heterogeneous electron transfer of cytochrome c at BDD, and also to investigate the effect of exposed surface oxygen groups via anodic polarization. 111 4.5 [1] [2-1 [3] [4-1 [5-1 [5-1 [7-1 [8-1 [9-1 [10.] [11.] References F. A. Armstrong and G. S. Wilson Electrochim. Acta 2000, 45, 2623-2645. F. M. Hawkridge and I. Taniguchi In The Porphyrin Handbook; K. M. Kadish, K. M. Smith and R. Guilard, Eds; Academic Press, Inc., New York, 2000; Vol. 8, 191-202. J. D. Burgess and F. M. Hawkridge ln Electroanalytical Methods for Biological Materials; A. Brajter-Toth and J. O. Chambers, Eds; Marcel Dekker, New York, 2002, 109-142. W. M. Clark Oxidation-Reduction Potentials of Organic Systems; Williams and Watkins Publishing Company, Baltimore, 1960. G. J. Tilley, R. Camba, B. K. Burgess and F. A. Armstrong Biochem. J. 2001, 360, 717-726. K. S. Chen, C. A. Bonagura, G. J. Tilley, J. P. McEvoy, Y. S. Jung, F. A. Armstrong, C. D. Stout and B. K. Burgess Nat. Struct. Biol. 2002, 9, 188- 192. K. S. Chen, Y. S. Jung, C. A. Bonagura, G. J. Tilley, G. S. Prasad, V. Sridhar, F. A. Armstrong, C. D. Stout and B. K. Burgess J. Biol. Chem. 2002, 277, 5603-5610. T. Kono and S. Nakamura Bull. Agr. Chem. Soc. Japan 1958, 22, 399- 403. G. W. Pettigrew and G. R. Moore Cytochrome c - Biological Aspects; Springer-Verlag, Berlin, 1987. F. M. Hawkridge and l. Taniguchi Comments Inorg. Chem. 1995, 17, 163- 187. Y. J. Zhen, C. W. Hoganson, G. T. Babcock and S. Ferguson-Miller J. Biol. Chem. 1999, 274, 38032-38041. 112 [12.] [13.] [14.] [15.] [16.] [17.] [18.] [19.] [20.] [21.] [22.] [23.] [24.] K. F. Wang, Y. J. Zhen, R. Sadoski, S. Grinnell, L. Geren, S. Ferguson- Miller, B. Durham and F. Millett J. Biol. Chem. 1999, 274, 38042-38050. V. A. Roberts and M. E. Pique Journal of Biological Chemistry 1999, 274, 38051-38060. D. Flock and V. Helms Proteins 2002, 47, 75-85. V. Drosou, B. Reincke, M. Schneider and B. Ludwig Biochemistry 2002, 41, 10629-10634. K. Niki, K. R. Pressler, J. R. Sprinkle, H. Li and E. Margoliash Russ. J. Electrochem. 2002, 38, 63-67. A. El Kasmi, M. C. Leopold, R. Galligan, R. T. Robertson, S. S. Saavedra, K. El Kacemi and E. F. Bowden Electrochem. Commun. 2002, 4, 177-181. I. Taniguchi, S. Yoshimoto, M. Yoshida, S. Kobayashi, T. Miyawaki, Y. Aono, Y. Sunatsuki and H. Taira Electrochim. Acta 2000, 45, 2843-2853. E. Lojou and P. Bianco J. Electroanal. Chem. 2000, 485, 71-80. A. El Kasmi, J. M. Wallace, E. F. Bowden, S. M. Binet and R. J. Linderman J. Am. Chem. Soc. 1998, 120, 225-226. S. Terrettaz, J. Cheng and C. J. Miller J. Am. Chem. Soc. 1996, 118, 7857-7858. P. Yeh and T. Kuwana Chem. Lett. 1977, 1145-1148. E. F. Bowden, F. M. Hawkridge and H. N. Blount J. Electroanal. Chem. 1984, 161, 355-376. S. C. Sun, D. E. Reed, J. K. Cullison, L. H. Rickard and F. M. Hawkridge Mikrochim. Acta 1988, 3, 97-104. 113 [25.] [26.] [27.] [28.] [29.] [30.] [31.] [32.] [33.] [34.] [35.] [36.] [37.] F. A. Armstrong, H. A. 0. Hill, B. N. Oliver and N. J. Walton J. Am. Chem. Soc. 1984, 106, 921-923. F. A. Armstrong, H. A. 0. Hill and B. N. Oliver J. Chem. Soc, Chem. Commun. 1984, 976-977. M. C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M. D. Koppang, J. E. Butler, G. Lucazeau, M. Mermoux, J. W. Strojek and G. M. Swain Anal. Chem. 2000, 72, 3793-3804. J. S. Xu, M. C. Granger, Q. Y. Chen, J. W. Strojek, T. E. Lister and G. M. Swain Anal. Chem. 1997, 69, A591-A597. G. M. Swain In Thin Film Diamond; C. Nebel and J. Ristein, Eds; Academic Press, 2002, submitted. G. M. Swain, A. B. Anderson and J. C. Angus MRS Bulletin 1998, 23, 56- 60. J. S. Xu, O. Y. Chen and G. M. Swain Anal. Chem. 1998, 70, 3146-3154. D. M. Gruen Annu. Rev. Mater. Sci. 1999, 29, 211-259. B. Fausett, M. C. Granger, M. L. Hupert, J. Wang, G. M. Swain and D. M. Gruen Electroanal. 2000, 12, 7-15. Q. Y. Chen, D. M. Gruen, A. R. Krauss, T. D. Corrigan, M. Witek and G. M. Swain J. Electrochem. Soc. 2001, 148, E44-E51. S. Haymond, G. T. Babcock and G. M. Swain J. Am. Chem. Soc. 2002, 124, 10634-10635. Y. 0. Su, T. Kuwana and S. M. Chen J. Electroanal. Chem. 1990, 288, 177-195. T. Sagara, K. Niwa, A. Sone, C. Hinnen and K. Niki Langmuir1990, 6, 254-262. 114 [38.] [39.] [40.] [41.] [42.] [43.] [44.] [45.] [46.] [47.] [48.] [49.] [50.] [51.] R. S. Nicholson Anal. Chem. 1965, 37, 1351. F. M. Hawkridge and T. Kuwana Anal. Chem. 1973, 45, 1021-1027. A. Sziics and M. Novak J. Electroanal. Chem. 1995, 383, 75-84. M. Hupert, A. Muck, J. Wang, J. Stotter, Z. Cvackova, S. Haymond, Y. Show and G. M. Swain Diam. Relat. Mat. 2002, in press. M. Witek, J. Wang, J. Stotter, M. Hupert, S. Haymond, P. Sonthalia, G. M. Swain, J. K. Zak, Q. Chen, D. M. Gruen, J. E. Butler, K. Kobashi and T. Tachibana J. Wide Bandgap Materials 2002, 8, 171-188. W. R. Hagen Eur. J. Biochem. 1989, 182, 523-530. W. F. Sokol, D. H. Evans, K. Niki and T. Yagi J. Electroanal. Chem. 1980, 108, 107-115. K. B. Koller, F. M. Hawkridge, G. Faque and J. LeGall Biochem. Biophys. Res. Commun. 1987, 145, 619-625. B. C. King and F. M. Hawkridge J. Electroanal. Chem. 1987, 237, 81 -85. 8. Song, R. A. Clark, E. F. Bowden and M. J. Tarlov J. Phys. Chem. 1993, 97, 6564-6572. S. H. Speck, S. Ferguson-Miller, N. Osheroff and E. Margoliash Proc. Natl. Acad. Sci. U. S. A. 1979, 76, 55-59. H. T. Smith, N. Staudenmayer and F. Millett Biochemistry 1977, 16, 4971- 4974. R. Rieder and H. R. Bosshard J. Biol. Chem. 1980, 255, 4732-4739. H. Wackerbarth, D. H. Murgida, S. Oellerich, S. Dopner, L. Rivas and P. Hildebrandt J. Mol. Struct. 2001, 563, 51-59. 115 [52.] S. Dopner, P. Hildebrandt, F. I. Rosell, A. G. Mauk, M. von Walter, G. Buse and T. Soulimane Eur. J. Biochem. 1999, 261, 379-391. [53.] P. Hildebrandt Biochim. Biophys. Acta1990, 1040, 175-186. [54.] J. Peschke and H. Mohwald Colloids and Surfaces 1987, 27, 305-323. [55.] H. Witt, F. Malatesta, F. Nicoletti, M. Brunori and B. Ludwig J. Biol. Chem. 1998, 273, 5132-5136. [56.] D. H. Murgida and P. Hildebrandt J. Phys. Chem. B 2001, 105, 1578- 1586. [57.] L. Rivas, D. H. Murgida and P. Hildebrandt J. Phys. Chem. B 2002, 106, 4823-4830. 116 CHAPTER 5 5 Low-Frequency Electrochemical Difference FTIR Spectroscopy of Metalloproteins: Challenges and Solutions 5.1 Introduction Vibrational spectroscopy is a vital tool in the process of elucidating structural or mechanistic information about metalloproteins. Of particular interest is information about bonding interactions between the metal site and its ligands, which is found in the low-frequency (<1000 cm'1) region of the infrared (IR) spectrum. Information on metal-ligand vibrations provides insight into the molecular structure and reaction intermediates of the enzymatic mechanism. Examples of structures in metalloenzymes that would give rise to modes in the low-frequency region of the spectrum are shown in Table 5-1. Identification of low-frequency modes would allow specific questions regarding the detailed chemistry of the mechanism to be addressed. Until recently, experimental access to low-frequency vibrations of metalloproteins has been limited to resonance Raman spectroscopy. This technique is useful for studying many biological systems, especially those containing highly absorbing chromophores. However, the technique is not useful for studying all metal sites in metalloproteins, particularly those without a Raman- excitable pigment, or those with multiple chromophores. The alternate technique for obtaining vibrational information is IR spectroscopy. This technique provides 117 information on all lR-active vibrational modes of a protein. According to theory, a nonlinear molecule, with N atoms, will exhibit 3N - 6 modes. Considering that even a small (~ 10,000 Da) protein has 1000’s of atoms, the IR spectrum of such a molecule is information rich. For example, the bacterial form of cytochrome c oxidase (Rhodobacter Sphaeroides), a 125,000 Da protein, is comprised of 90,000 atoms. The IR spectrum of this protein then contains over 250,000 vibrational modes. Therefore, difficulty arises not in the measurement of protein spectra, but in the detection of the small amplitude signals arising from the active site structural changes. The detection of such single modes is difficult because they are superimposed on a large background due to water and global protein mode absorbance. To achieve adequate levels of signal-to-noise ratio, accurate background subtraction, stable background absorbance, minimization of the background absorbance, and maximization of the signal absorbance are required. In this chapter, methods are described, in which reaction-induced difference infrared spectroscopy (i.e., electrochemical difference FTIR spectroscopy) is used. Accurate solvent subtraction is achieved, resulting in peak-to-peak noise levels of 105-10'6 absorbance units, thus, rendering the acquisition of low-frequency vibrational information possible. In addition to the problem of the large water absorbance, other technical difficulties associated with accessing the low-frequency region of the spectrum will be discussed. 118 5.2 Structural Information Possible (Far-IR Region) The acquisition of low-frequency infrared spectra of proteins is a relatively new field of IR spectrosc0pic research. While there has been much work performed to assign protein vibrational modes in the mid-IR region (> 1000 cm'1), there is little information to use as a basis for the assignment of metal-ligand modes in the low-frequency region. Predictions about several possible vibrational modes of polypeptides have been made based upon studies of model peptides, such as N-methyl acetamide [1]. Table 5-1 summarizes these modes. This table predicts that the protein secondary structure should display several low-frequency vibrations, arising from primarily out-of-plane peptide backbone vibrations. More interesting information, necessary for understanding mechanistic details, lies in those modes that originate not from the backbone, but from the metal-ligand interactions of the protein active site. Table 5-1 lists a number of metal-ligand vibrations from synthetic model compounds and proteins, such as heme-ligand complexes and the photosystem M protein (PSII). Several of these have been detected by FTIR, whereas others are reported from resonance Raman studies [2-4]. Assignment of vibrational modes is necessary to extract structural data from an information-rich protein spectrum. Isotope exchange plays a critical role in the assignment of vibrational modes. This is especially important when there are many peaks in a spectrum, and it is desired to assign one to a specific molecular species. Substitution of a heavier atom will affect the frequency of the mode according to the following relation, 119 where Vis the frequency of the mode, [1 is the reduced mass, k is the force constant and c is the speed of light. From this equation, it is clear that introducing a heavier atom will result in a decrease in the frequency of the affected mode. Isotope exchange can be accomplished by solvent exchange (e.g., D20 buffers vs. H20 buffers), metal exchange (e.g., 63Cu vs. SSCu), atom exchange (e.g., 15M vs. 14N or 180 vs. 160) and specifically labeled amino acids (e.g., 15N His, 13C Ala). Metal-oxygen species, such as those proposed to exist in the catalytic cycles of cytochrome c oxidase (cco) and PS“, will exhibit infrared absorptions sensitive to isotope exchange. Site-directed mutagenesis is another important tool in the mode assignment of protein spectra. Identification of the mid-IR mode due to a mechanistically important amino acid residue in cco, Glu 286, has been achieved through experiments involving site-directed mutants of this residue [5]. It is clear from Table 5-1 that a method, in which low-frequency modes of metalloenzymes can be measured, will be integral in elucidating mechanistic details of these proteins. 120 Table 5—1 Low-Frequency Vibrational Modes of Metalloenzymes Vibrational Frequency Structure (cm'1) Reference Protein Backbone Modes Amide V (NHob, CN() 725 [6, 7] Amide IV (Cob, CCS, CNCd) 625 [6, 7] Amide VI (COob, CN[) 600 [6, 7] Amide Vll (NHob, CNt, com) 200 [6, 7] Metal-Ligand Modes Cu—OEC 400 [8-10] Cu—N(His) 260 [11, 12] Cu—OH 550-500 [13] CU-OHz 440 [14] Fe—O(Tyr) 600 [1 5] Fe—N(His) 220-250 [16, 17] Fe—OEC 475-575 [8-10] Fe—OH 500 [13] Fe—OH2 390 [14] Fe=O 775-850 [18-23] Mn—OH 400-500 [13] Mn—OH2 395 [14] Mn=O 750 [24] s, stretch; d, deformation; t, torsion; ib, in-plane-bend; ob, out-of-plane-bend; His, Histidine; Tyr, Tyrosine 121 5.3 FTIR of Aqueous Samples Biology uses water as the solvent for a myriad of biochemical reactions. However, water does pose certain difficulties for obtaining protein IR spectra. The vibrational modes of water give rise to very intense absorption throughout regions of the mid-IR and they overlap with many protein modes of interest. Rotational and vibrational modes of water also effectively obscure the low- frequency IR region (below 1000 cm'1). Figure 5-1 demonstrates the four regions of spectral influence by water. A series of O-H stretching modes at 3920, 3490 and 3280 cm’1 give rise to an intense and broad absorbance band centered at 3400 cm'1. The peak at 1645 cm"1 arises from the H-O-H bending mode and the absorbance at 2125 cm'1 is attributed to the association of hydrogen bonds The broad feature below 1000 cm'1 is due to the strongly coupled lattice vibrations of aggregates of water molecules [1, 25]. In order to observe vibrational modes that are obscured by the water absorbance, ways must be found to either minimize this background absorbance or to stabilize it so that it can be accurately subtracted out of the spectrum. In particular, methodologies are discussed, through which one can observe high-quality spectra for metalloenzymes in the mid-IR and possibly, the far-IR regions of the spectrum. 122 Figure 5-1 Infrared Absorption Spectra of H20 and D20 A water spectrum (solid line) and a deuterium oxide (99.96% D) spectrum (dashed line) shown for comparison. The spectra were acquired at 15 °C using AgBr windows separated by a 6 pm spacer. Each spectrum is an average of 32 scans collected versus an air background using a narrow band MCT detector. Absorbance (Abs. Units) .N O r .3 01 r _L O r .0 (11 I *Mww-w: "1 a" . I ‘ -. ‘T' R...‘ . 4000 3500 3000‘2500‘2000 Frequency (cm'1) Table 5—2 Infrared Properties of H20 and DZO Mode H20 (cm'T £(M'1cm'1) ozo (cm'1) e(M“cm") O-X stretch 3920 0.83 2900 0.60 3490 62.7 2540 59.8 3280 54.5 2450 55.2 Association 2125 3.23 1555 1.74 X-O-X Bending 1645 20.8 1215 16.1 Libration Broad absorbance below 1000 cm'1 123 Figure 5-2 Comparison of the H20 Absorption Spectrum and the Noise Spectrum for an Aqueous Sample Comparison of (A) a water absorbance spectrum and (B) a resulting noise spectrum. The noise spectrum was obtained by subtracting two successive spectra. Spectra collected using a MCT detector, AgBr windows, a 6 pm pathlength, and averaging 32 scans at 15 OC. Note the different absorbance scales. 4000'35'00'3000'2500 2000'15'00'1000' Frequency (cm.1) The vibrational modes associated with the prominent water absorbances are well characterized [1, 25]. A list of these vibrational modes and corresponding extinction coefficients are presented in Table 5-2. In general, in order for a solvent spectrum to be accurately subtracted from a protein sample, 124 the overall absorbance of a given peak must be less than one absorbance unit [26]. Therefore, the pathlength of an aqueous solution should not exceed 10 um, based on the extinction coefficient of 20.8 (M'1 cm'1) for the water bending mode at 1645 cm'1. Protein mid-IR spectra, showing high-quality mathematical subtraction of the overlapping water signal, are routinely acquired by using highly concentrated samples and optical pathlengths of 10 pm or less [27]. Highly concentrated protein samples serve to increase the signal amplitude and decrease the effective water concentration. While this method is successful for the mid-IR region, it is not applicable to measurements in the low-frequency region. The broad water lattice vibrations below 1000 cm"1 are extremely sensitive to minute changes in pathlength or sampling conditions (i.e., temperature, ionic strength) [25]. Large spectral artifacts remain even after careful subtraction when measured at pathlengths as short as 6 pm, as shown in Figure 5-2. This Figure illustrates that regions of prominent water absorbance correspond to regions of increased noise. Reaction-induced difference infrared spectroscopy is an essential technique for the precise subtraction of solvent peaks, such as those due to water, especially in protein samples. The principle of this method involves the subtraction of a protein absorbance spectrum in state A from that of the protein in state B. The resulting difference spectrum (B minus A) reports on the small and subtle vibrational changes occurring in the protein as it undergoes the chemical reaction. A trigger initiates a reaction to advance the protein from state A to state B. Several triggers are applicable, including intrinsic photochemical 125 reactions, electron transfer reactions, and indirect photochemical reactions [28, 29]. Reaction-induced difference spectroscopy is very powerful in that it introduces selectivity by subtracting out the background absorbance due to functional groups that are not involved in or influenced by the chemical reaction. It also allows the issue of solvent subtraction to be finessed, as no physical perturbation of the sample is involved, resulting in a constant optical pathlength and no gross perturbation of the solvent concentration or ionic strength. Thus, the relatively large baseline distortions that are common in solvent subtracted data are eliminated. Accurate background subtraction results in spectra without baseline distortions, which greatly improves the signal-to—noise ratio and enables the detection of weakly intense signals obscured by the background absorbance. Global protein structural modes, particularly the so—called Amide l mode, commonly exhibit infrared absorption approaching one absorbance unit in magnitude. However, the reaction of interest will result in only slight perturbations, affecting a few amino acids or just the active site. These single mode changes, arising from the reaction-induced perturbation, will constitute a small fraction of a percent of the overall signal. These single modes have intensities of approximately 10'4 absorbance units, and are commonly detected using this methodology, with sufficient signal-to-noise ratios. It is these single mode signals, which may reflect changes in the enzyme structure that are directly related to the catalytic cycle we are seeking to acquire. 126 Reaction-induced difference FTIR spectroscopy has been successfully applied using a direct photochemical approach for light-activated photosynthetic proteins in the mid- and far-IR [2, 3, 30, 31]. Generalizing the method by using electrochemical reactions instead of actinic light reactions will enable access to a wider class of metalloenzymes. Therefore, any metalloprotein with a redox- active metal would be responsive to this technique. This electrochemically- initiated method has proven successful for mid-IR measurements of several proteins, from simplistic, single redox center proteins (e.g., cytochrome c, cytochrome f) [32] to more complex enzymes, which undergo multiple electron transfer steps at several redox metal centers (e.g., cytochrome c oxidase) [33]. It has not, as yet, been utilized in far-IR measurements, owing to a number of complications, as discussed in this chapter. The major drawback of using the electrochemicalIy-initiated method in the transmission mode is the requirement for a supporting electrolyte solution. Therefore, the measurements suffer from water absorption interference in the low-frequency IR region. This reaction modulated difference technique provides a very detailed look at changes that occur as an enzyme system undergoes its catalytic cycle. Figure 5-3 shows a theoretical construction of such a difference spectrum. A possible step in the catalytic cycle of cytochrome c oxidase, the mitochondrial respiratory enzyme that is responsible for the four-electron reduction of oxygen to water, is shown in Figure 5-3. The Figure depicts the affect of the copper oxidation state on the Cu-N(His) stretching vibration. In the initial state (Figure 5-3A), the copper ion is in the oxidized (+2) state and a positive peak 127 attributed to the metal-nitrogen vibrational mode is shown in spectrum 5-3A. The electron transfer reaction is initiated and the copper is reduced (+1), resulting in a downshift in the frequency of the Cu-N(His) mode (Figure 5-38). Two other peaks are present at the same position in both states of the enzyme. Figure 5-3C shows the reduced minus oxidized electrochemical difference spectrum, which is the result of subtracting the initial (oxidized) spectrum from that of the final (reduced). This spectrum reveals a derivative feature, consisting of both positive and negative bands, which are characteristic of a difference spectrum. The positive modes are indicative of vibrational modes present in the final (reduced) state and the negative bands represent modes for the initial (oxidized) state. The peaks that were unaffected by the oxidation state change of the metal are subtracted out, leaving a flat baseline feature in the difference spectrum. In theory, modes due to ligands (e.g., amino acids and substrate molecules) in the vicinity of the active site and those coordinated to the active site will undergo a frequency shift upon a change in the oxidation state of the metal. Any vibrational modes that are unaffected by the reaction will display exactly the same vibrational frequency in the initial and final spectra. This includes modes that are not redox-sensitive, or are distant from the active site where the chemistry is occurring. These modes will be subtracted out and, therefore, do not contribute to the difference spectrum. Any mode that is slightly perturbed (e.g., an amino acid residue undergoing a H-bonding change, a solvent molecule that is in the active site, or a subtle change in the composition of the vibrational mode) by the reaction will appear as difference peaks. 128 Figure 5-3 Theoretical Construction of an Electrochemical Difference FTIR Spectrum Schematic showing the components of an electrochemical FTIR difference spectrum. (A) The absolute (solvent subtracted) spectrum for the initial (oxidized) form. (8) The absolute (solvent subtracted) spectrum for the final (reduced) form after an electron transfer step. (C) The resultant difference spectrum generated by subtracting the initial from the final spectrum (reduced minus oxidized). Peaks have different amplitudes due to differences in extinction coefficient. -1 < Frequency (cm ) The importance of this type of difference spectrum lies in its ability to provide evidence that could validate (or invalidate) a proposed model of the enzyme’s catalytic cycle. Using the example in Figure 5-3, it is has been suggested that one of the histidine ligands on the Cu is involved in the proton pumping process—an important step in the enzymatic reaction of cytochrome c oxidase [34, 35]. This particular copper ion, CUB, is “invisible” to other types of 129 spectroscopy, such as EPR and Resonance Raman, due to interference by the surrounding hemes. Therefore, low-frequency data (for this metal center) obtained by IR would provide valuable information regarding mechanistic questions involving the role of Cue in the catalytic cycle. Reaction-induced difference spectroscopy is a useful method for simultaneously introducing selectivity and minimizing solvent contribution to complex protein infrared spectra. However, in order to obtain an accurate solvent subtraction, it is necessary to use sample absorbances that do not exceed the limitation for a linear detector response. Systematic errors will be present in difference spectra acquired under conditions where the detector response is unreliable. As already mentioned, the use of short pathlengths can ensure that fully hydrated samples do not exceed the absorbance limitations of the instrumentation. This method has worked well for studies of highly concentrated protein samples in the mid-IR region [32, 36-40]. However, additional considerations are required for handling water absorbance below 1000 cm“. A commonly applied approach for circumventing problems associated with the presence of strongly absorbing water bands in the mid-IR region of the spectrum involves substituting deuterium oxide for water. This approach is particularly suited for cases in which the observation of sample absorption is compromised by overlapping solvent absorption. Studies of hydrated lipid dispersions [41], as well as those aimed at determining protein secondary structure [42], have been successfully performed using this method. As shown in 130 Figure 5-1, substituting DZO for H20 will shift the solvent absorption to lower frequencies, thereby enabling observation of the underlying sample bands. This approach is not entirely free of potential problems, as shifting the solvent absorption to lower frequencies may compromise the detection of other sample absorption bands of interest. However, access to the low-frequency region can be considerably extended through the use of deuterated buffers. Examination and comparison of spectra acquired in H20 and D20 would enable one to access the entire spectrum of an aqueous protein sample. Acquiring spectra in deuterated solvents can pose a more serious problem—the H/D exchange of protons within the protein. This will result in loss of the absorption band of a protonated species and its replacement with an absorption band of the deuterated species. The latter band will likely be at a lower frequency, and in some cases, not observable (i.e., obscured by other modes). If the H/D exchange rates are slow, the disappearance of the Amide l band and subsequent replacement by its deuterated counterpart (Amide I‘) can be exploited to monitor the rates and extent of exchange. As will be discussed in Chapter 6, such information provides insight into the exposure of exchangeable proton sites to the bulk solvent. Although H/D exchange may introduce further complexities to the protein difference spectrum, major problems in mode assignment can be overcome by appropriately considering such occurrences when interpreting the data. 131 5.4 Detectors Detector choice is a critical factor in determining the quality of FTIR spectra of biological systems, as IR detectors have different sensitivities, linear dynamic ranges, and response times. This is particularly true for measurements made in the low-frequency region because of the inherently low signal-to-noise levels. The drop in signal-to-noise ratio observed with decreasing frequency is due to decreased light intensity from the source in this region, and smaller vibrational mode extinction coefficients. Absorption of low-frequency IR radiation by optical materials also contributes to the decreased throughput observed in the far-IR region. The useful range and sensitivity of the IR detector greatly affects the signal-to-noise ratio. There are many types of lR-sensitive detectors available for use, however, the two most common are mercury-cadmium-telluride (MCT) and deuterated triglycine sulfate (DTGS). These detectors differ in their sensitivity, but are similar in that they enable signal collection across a broad region of the infrared spectrum (i.e., mid-IR with some access into the low-frequency region). In general, the MCT is 5 to 50 times more sensitive than the DTGS detector, depending on the bandwidth (accessible range) of the MCT detector [43, 44]. Another type of infrared-sensitive detector, the liquid helium-cooled bolometer, is superior to the MCT for low-frequency IR detection [44]. Infrared detectors are evaluated in terms of their noise-equivalent power (NEP) and compared using their specific detectivity, or D* value [26, 43, 44]. The 132 sensitivity of an infrared detector is usually expressed in terms of the NEP (w Hz‘”), NEP = 117— R" In this expression, Vn is the root mean square (rms) of the detector voltage in V Hz‘y’, and RV is the voltage responsivity of the detector to signals in V W'1 [26]. The specific detectivity, D* (cm szz W4), of a detector is related to the NEP by In *: (1‘1/)) - NEP where AD is the detector element area. The D* value is convenient for comparing the sensitivity of different types of infrared detectors, as to a first approximation; it is independent of the area of the element. When comparing infrared detectors, a higher D* value is indicative of a more sensitive detector. The D* expression is related to the overall signal to noise ratio (SNR) of the FTIR instrument by the following U,—,(T)-®-M-t“2 -.§ NEP SEN’R : U,—,(T).o-.\v.i"'2-§-D* A52 U ,7 (T) represents the spectral energy density, O is the optical throughput of the system, 017 is the spectral resolution, t is the acquisition time, and § is the efficiency factor of the instrument [26]. From this expression, it is clear that the signal-to-noise ratio of a given spectrum is linearly dependent upon the D* value 133 of the detector, the resolution, and the throughput. Therefore, it is integral to optimize these values in order to acquire the highest quality spectra in the shortest amount of time. The choice of detector will affect the type of FTIR measurements undertaken. Thermal type detectors are the most common, as they have the most linear response function. These detectors operate by sensing the change in temperature of an absorbing material as it experiences a change in the infrared radiation striking it. The pyroelectric bolometer is the most common class of thermal-type detectors. The sensing element in a pyroelectric bolometer is a material that changes its electrical resistance when its temperature changes. The most commonly used material is DTGS. DTGS detectors have the advantage that they can be operated at room temperature while displaying highly linear responses for varying intensities of incident IR light. The main drawback of this detector, which renders it unsuitable for biological spectroscopy, is its relatively poor sensitivity (0* = ~ 107 - 108 ). This level of sensitivity makes the detection of minute signals, obscured by large background signals, very difficult. The advent of semiconductor-type quantum detectors over the last decade has revolutionized the field of biological infrared spectroscopy. This class of highly sensitive detectors has made it possible for investigators to detect small amplitude signals in highly absorbing samples. The MCT detectors are the most common of this class. The D* values for MCT detectors range from 109 - 1011, and higher. The principle of operation is based on a ternary semiconductor detector element whose bandgap corresponds to the energies present in IR 134 radiation. These detectors must be cooled, usually to liquid nitrogen temperature (77 K), to suppress the contributions of thermally excited electrons to the detected signal. The specific stoichiometric ratio of the constituent elements determines the useful range of these detectors. Although, an MCT detector can be optimized for response at low-frequency wavelengths (peak response at ~ 500 cm'1) the D* value is approximately an order of magnitude less than one optimized for the mid-IR. MCT detectors are often classified as high-sensitivity (narrow-band) or low sensitivity (broadband, or wide-band) detectors. Therefore, it follows that an MCT with an intermediate bandwidth will possess an intermediate relative detectivity [26]. An alternate choice over MCT detectors for low-frequency IR detection is the liquid helium-cooled bolometer. This quantum detector utilizes a material, such as silicon or germanium, as the detector element, and operates under the same technology as the MCT detector. However, in this type of detector, the detector element temperature must be maintained below that attainable by liquid nitrogen cooling. Therefore, liquid helium is used, necessitating the need for liquid helium cryostat hardware. The additional complexity associated with the cryostat is the key difference that results in a distinct cost differential between this type of detector and the simpler nitrogen-cooled MCT. Helium-cooled bolometers have D* values larger than MCT detectors (D* ~1012 -1014), and their response to radiation in the far-IR exceeds that of any MCT [26, 44]. Chu and coworkers have utilized a helium-cooled Si bolometer for low-frequency FTIR measurements of the Photosystem Il protein to detect single modes due to 135 the active site metal-ligand interactions [3, 45]. Their results demonstrate the potential utility of helium-cooled bolometers in measurement of high quality, low- frequency FTIR spectra of biological samples. 5.5 Window Materials One difficulty encountered with transmission spectroscopy involves the bandpass characteristics of the window materials. The material must be transparent to the incident IR radiation, and must also possess a high enough optical throughput to achieve sufficient signal-to-noise in the resulting spectra. Numerous materials are available for use as IR windows; however, the number of these substantially decreases when one is interested in utilizing aqueous samples and observing the low-frequency region (< 1000 cm'1) of the spectrum. Additionally, due to the large amount of useful information in the mid-IR region of a protein difference spectrum, it would be beneficial to use a material that has sufficient optical throughput from 2000 crn’1 down to less than 400 cm'1, ensuring that all of the vibrational modes of interest are observable. Table 5-3 lists common IR window materials and corresponding frequencies of transmission, indices of refraction, and water solubility. There are several materials that transmit into the low-frequency region and are insoluble in water. The most commonly used are the silver halide salts. Silver bromide and silver chloride transmit from the visible region to 285 and 435 cm'1, respectively. Both have a tendency to physically distort (cold flow), and prolonged exposure to light leads to photoreduction of the window (e.g., Ag+ to Ag) and gradual 136 darkening, resulting in decreased transmission properties over time. The silver halide salts have been successfully used in light-induced difference FTIR experiments of PSII [2, 3, 30, 31]. However, in tests of the electrochemical difference method, they were found to undergo electrochemical attack by the gold contact wire and gold mesh used as electrodes in this setup. Zinc selenide and cadmium telluride have not seen widespread use for biological FTIR experiments due to their higher cost, and the availability of cheaper materials that exhibit similar optical properties. Germanium is not well suited for transmission experiments, due to its high refractive index. An added disadvantage is its low frequency cutoff of 600 cm‘1, as many metal-ligand vibrations fall below this cutoff. Silicon is most often used as a reflection element, owing to its high refractive index. Transmission measurements are possible using thin windows of Si; however, transmission levels are relatively low due to high reflection losses. Polyethylene suffers from a limited transmission range, but is a good option for studies only concerned with low-frequency vibrations. Of all the materials listed, diamond has the widest optical window, transmitting from the UV down to the far- IR region. This material possesses added advantages of being mechanically strong, chemically inert, and when doped with boron, useful as an optically transparent electrode. One of the most significant improvements over the light- induced method is the incorporation of diamond windows into our design. With these windows in place, the limitations on the signal-to—noise ratio due to light throughput are eliminated. 137 Table 5—3 Properties of Infrared Transmissive Materials Material Transmission Refractive Water Solubility Range (cm'1) Index (at 10W“) (gl100 9 H20) NaCI 40,000 - 625 1.49 Soluble KBr 40,000 - 400 1.52 Soluble Csl 10,000 - 200 1.74 Soluble AgCI 25,000 - 435 1.98 0.00015 AgBr 20,000 - 285 2.2 0.000012 ZnSe 10,000 - 500 2.4 Insoluble CdTe 5,000 - 320 2.67 Insoluble Ge 5,000 - 600 4.0 Insoluble Si 10,000 — 1540 3.4 Insoluble 500 - 30 Polyethylene 625 - 30 1.54 Insoluble Diamond 40,000 - 33 2.42 Insoluble 5.6 Temperature Control In the case of reaction-induced difference FTIR experiments, where one is requiring very stable background subtraction, temperature control is a crucial element in the experimental setup. Fluctuations in sample temperature result in large baseline distortions, which map to regions of strong absorption in the infrared spectra of aqueous systems [25]. Figure 5-4 displays this effect on the baseline, as a series of noise spectra are shown for temperature drift from 5 to 0.005 °C. The spectra were collected for an aqueous sample with a 10 pm 138 pathlength at different temperatures. To acquire such a noise spectrum, a spectrum was measured at one temperature, and after decreasing the temperature a second spectrum was collected (i.e., Ti = 15 °C, Tf = 10 °C, so AT: 5°C). The difference of these two successive spectra yields a noise spectrum. It is clear that only with precisely controlled temperature does the baseline become flat and featureless, as the spectra for the more extreme temperature drifts contain peak-like features. The appearance of these “thermal bands” can easily be mistaken for real peaks and have undoubtedly contributed to many imagined peaks in the published literature. Figure 5-4 Effect of Temperature Drift on Baseline Linearity A series of noise spectra acquired by subtracting successive spectra at different temperatures. The presented spectra are representative of the following temperature drifts (0 C): (A) 5, (B) 0. 5, (C) 0.05, and (D) 0. 005. Spectra were collected for a 10 pm thick aqueous solution pressed between two BaFg windows. Each spectrum is an average of 200 scans and the final temperature was set to 10 DC in all cases. 0.0003 g a O 8 (a) AAbsorbance 00006 I — 2000.1800‘1600‘1400‘12100 1000 Frequency (cm-1) 139 The spectrometer set-up designed to maintain the required temperature control precision was detailed in Section 2.5 of the experimental section. Employment of this set-up enables temperature control to :t 0.005 °C. Figure 5-40 illustrates that at this level of precision, undesired thermal contributions to the baseline are largely eliminated, resulting in a baseline that is linear with an extremely low peak-to-peak noise level (AA ~ 1 x 10'5). 5.7 Conclusions There are a number of factors that contribute to the technical difficulty of obtaining high quality, IR difference spectra of proteins. A protocol has been developed using advanced materials and instrumentation that alleviates many of the experimental difficulties with electrochemical difference FTIR. The detection of weak absorption signals, in an othenivise highly absorbing sample is possible at high signal-to-noise ratio using this experimental set-up. Implementing reaction-induced difference spectroscopy with a precise background subtraction and limited path lengths of less than 10 um eliminates the large background interference from water in the mid-IR region. A highly sensitive quantum detector (MCT) is employed to improve the signal-to-noise ratio. Lastly, precise temperature control of the system is utilized to diminish baseline distortions and to eliminate thermal contributions to the difference spectra, further improving the signal-to-noise ratio. To date, attempts to measure low-frequency electrochemical difference transmission FTIR spectra have been plagued by large water absorbance, even 140 at pathlengths less than 10 pm. In the light-initiated method used to measure low-frequency IR spectra of the photosystem II protein, samples were dried directly onto the IR windows, successfully decreasing the amount of water present [2, 3]. However, as previously mentioned, the electrochemical approach requires some amount of aqueous solution for maintaining electrical contact between the electrodes. An alternate approach for performing electrochemical difference transmission IR measurements, relies on the method of protein film voltammetry [4649] Protein film voltammetry is a general concept that is based on the fact that direct electron transfer has been observed for protein films immobilized at electrode surfaces. There are a variety of ways to achieve such electroactive protein films, including self-assembled monolayer technology and cast biomembrane-like films [47]. In the latter approach, the protein is embedded into a membrane-like film comprised of surfactant molecules that allows for direct electron transfer between the protein and electrode surface. This method looks promising for decreasing the water content required for electrochemistry, as these protein films are essentially dried, yet they retain conductivity, as well as native protein structure and catalytic activity [48, 50, 51]. Protein film voltammetry has proven successful for a number of heme proteins at a variety of electrode surfaces. In particular, Hawkridge and coworkers have successfully immobilized cco in a lipid bilayer membrane on a Au electrode. They demonstrated that the cco undenivent direct electron transfer with the electrode and mediated the electron transfer from cytochrome c in solution to the electrode [52]. 141 Alternatively, a reflection method, for example, attenuated total reflection (ATR), could be employed, instead of the transmission mode. The advantages of ATR include less interference from water absorption, the possibility for increased signal amplitude due to multiple reflections, and ease of spectroelectrochemical cell design. Additionally, as shown in Table 5-3, many of the materials that are transmissive below 1000 cm’1 and are not soluble in water possess high indices of refraction, and are, therefore, commonly employed as ATR crystals Chapter 7 details the viability of using a boron-doped diamond thin-film on Si as an optically transparent electrode for IR transmission spectroelectrochemical measurements of a protein (cytochrome 0). Owing to the high refractive index properties of both BDD and Si, this OTE could also be implemented in reflection mode. Aside from attempts to minimize the water absorption below 1000 cm'1, a highly sensitive detector is required for measuring the weakly absorbing modes in the low-frequency IR region. Attempts to interface the He-cooled bolometer with the Bruker FTIR spectrophotometer using a homemade external detector compartment were unsuccessful. As discussed earlier, the MCT detector does not possess adequate sensitivity in the low-frequency IR region, which is required for protein measurements in aqueous solution. It does not seem possible to measure high-quality low-frequency data using this FTIR spectrophotometer until this detector problem is solved. 142 5.8 [1-1 [2-1 [3-1 [4-1 [5.] [6.] [7.] [3-1 [9-1 [10.] [11.] [12.] References J. L. Arrondo, A. Muga, J. Castresana and F. M. Goni Prog. Biophys. Mol. Biol. 1993, 59, 23-56. H.-A. Chu, M. T. Gardner, J. P. O'Brien and G. T. Babcock Biochemistry 1999, 38, 4533-4541. H.-A. Chu, M. T. Gardner, W. Hillier and G. T. Babcock Photosynth. Res. 2001, 66, 57-63. M. T. Gardner, G. Deinum, Y. Kim, G. T. Babcock, M. J. Scott and R. H. Holm Inorg. Chem. 1996, 35, 6878-6884. P. Hellwig, B. Rost, U. Kaiser, C. Ostermeier, H. Michel and W. Mantele FEBS Lett. 1996, 385, 53-57. K. ltoh, T. Shimanouchi and M. Oya Biopo/ymers 1969, 7, 649-658. T. Miyazawa, T. Shimanouchi and S. Mizushima J. Chem. Phys. 1958, 29, 611-616. J. Ramsden and T. G. Spiro Biochemistry 1989, 28, 3125-3128. S. 2. Hu, K. M. Vogel and T. G. Spiro J. Am. Chem. Soc. 1994, 116, 11187-11188. M. Tsubaki, R. B. Srivastava and N. T. Yu Biochemistry 1982, 21 , 1132- 1140. L. Nestor, J. A. Larrabee, G. Woolery, B. Reinhammar and T. G. Spiro Biochemistry 1984, 23, 1084-1093. T. J. Thamann, P. Frank, L. J. Willis and T. M. Loehr Proc. Natl. Acad. Sci. U. S. A. 1982, 79, 6396-6400. 143 [13.] [14.] [15.] [16] [17.] [18.] [19.] [20.] [21.] [22.] [23.] [24.] [25.] [26.] S. A. Asher, L. E. Vickery, T. M. Schuster and K. Sauer Biochemistry 1977, 16, 5849-5856. I. Nakagawa and T. Shimanouchi Spectrochim. Acta 1964, 20, 429-439. K. Nagai, T. Kagimoto, A. Hayashi, F. Taketa and T. Kitagawa Biochemistry 1983, 22, 1305-1311. K. Nagai, T. Kitagawa and H. Morimoto J. Mol. Biol. 1980, 136, 271-289. J. Teraoka and T. Kitagawa J. Biol. Chem. 1981, 256, 3969-3977. S. Hashimoto, Y. Tatsuno and T. Kitagawa Proc. Jpn. Acad. Ser. B-Phys. Biol. Sci. 1984, 60, 345-348. S. Hashimoto, Y. Tatsuno and T. Kitagawa Proc. Natl. Acad. Sci. U. S. A. 1986, 83, 2417-2421. R. T. Kean, W. A. Oertling and G. T. Babcock J. Am. Chem. Soc. 1987, 109, 2185-2187. W. A. Oertling, R. T. Kean, R. Wever and G. T. Babcock Inorg. Chem. 1990, 29, 2633-2645. A. J. Sitter, C. M. Reczek and J. Terner J. Biol. Chem. 1985, 260, 7515- 7522. A. J. Sitter, C. M. Reczek and J. Terner Biochim. Biophys. Acta 1985, 828, 229-235. J. T. Groves and M. K. Stern J. Am. Chem. Soc. 1988, 110, 8628-8638. K. Rahmelow and W. Huebner Appl. Spectrosc. 1997, 51, 160-170. P. R. Griffiths and J. A. de Haseth Fourier Transform Infrared Spectrometry; Wiley-Interscience, New York, 1986. 144 [27.] [28.] [29.] [30.] [31 .1 [32.] [33.] [34.] [35.] [36.] [37.] [38.] [39.] J. L. Koenig and D. L. Tabb Infrared Spectra of Globular Proteins in Aqueous Solution; D. Reidel Publishing Co., New York, 1980. C. Zscherp and A. Barth Biochemistry 2001, 40, 1875-1883. W. Mantele In Biophysical Techniques in Photosynthesis; J. Amesz and A. J. Hoff, Eds; Kluwer Academic Publishers, Dordrecht, 1996; Vol. 3, 137- 160. H. A. Chu, W. Hillier, N. A. Law and G. T. Babcock Biochim. Biophys. Acta 2001, 1503, 69-82. H. A. Chu, W. Hillier, N. A. Law, H. Sackett, S. Haymond and G. T. Babcock Biochim. Biophys. Acta 2000, 1459, 528-532. D. Moss, E. Nabedryk, J. Breton and W. Mantele Eur. J. Biochem. 1990, 187, 565-572. P. Hellwig, U. Pfitzner, J. Behr, B. Rost, R. P. Pesavento, W. von Donk, R. B. Gennis, H. Michel, B. Ludwig and W. Mantele Biochemistry 2002, 41 , 9116-9125. H. Michel Biochemistry 1999, 38, 15129-15140. M. Wikstrém Biochim. Biophys. Acta 2000, 1458, 188-198. F. Baymann, D. A. Moss and W. Ma'ntele Anal. Biochem. 1991, 199, 269- 274. J. Behr, P. Hellwig, W. Mantele and H. Michel Biochemistry 1998, 37, 7400-7406. P. Hellwig, S. Grzybek, J. Behr, B. Ludwig, H. Michel and W. Mantele Biochemistry 1999, 38, 1685-1694. P. Hellwig, C. Ostermeier, H. Michel, B. Ludwig and W. Mantele Biochim. Biophys. Acta 1998, 1409, 107-112. 145 [40.] [41.] [42.] [43.] [44.] [45.] [46.] [47.] [48.] [49.] [50.] [51.] [52.] P. R. Rich and J. Breton Biochemistry 2001, 40, 6441-6449. R. N. A. H. Lewis and R. N. McElhaney In Infrared Spectroscopy of Biomo/ecules; H. H. Mantsch and D. Chapman, Eds; Wiley-Liss, Inc., New York, NY, 1996, 159-202. B. Stuart Biological Applications of Infrared Spectroscopy; John Wiley & Sons, New York, 1997. J. D. lngle and S. R. Crouch Spectrochemical Analysis; Prentice Hall, Englewood Cliffs, NJ, 1988. J. H. Moore, C. C. Davis and M. A. Coplan Building Scientific Apparatus: A Practical Guide to Design and Construction; 2nd ed.; Perseus Books, Reading, MA, 1991. H. A. Chu, H. Sackett and G. T. Babcock Biochemistry 2000, 39, 14371- 14376. J. F. Rusling Acct. Chem. Res. 1998, 31, 363-369. F. A. Armstrong and G. S. Wilson Electrochim. Acta 2000, 45, 2623-2645. A.-E. F. Nassar, W. S. Willis and J. F. Rusling Anal. Chem. 1995, 67, 2386-2392. J. F. Rusling and A. E. F. Nassar J. Am. Chem. Soc. 1993, 115, 11891- 11897. J. F. Rusling and Z. Zhang Electroanalytical Methods for Biological Materials 2002, 195-231. J. F. Rusling and Z. Zhang Handbook of Surfaces and Interfaces of Materials 2001, 5, 33-71. J. K. Cullison, F. M. Hawkridge, N. Nakashima and S. Yoshikawa Langmuir 1994, 10, 877-882. 146 CHAPTER 6 6 Electrochemical Difference FTIR Spectroscopy of Cytochrome c Oxidase 6.1 Introduction Cytochrome c oxidase (cco) is the terminal electron acceptor in the respiratory chain and is found in the mitochondria of eukaryotes, in many bacteria, and in the archea. It is the last of the transmembrane proteins involved in the respiratory electron transfer chain and is one member of the heme-Cu terminal oxidase protein family [1]. The chemistry catalyzed by cco involves an efficient reduction of molecular oxygen to water with rates that exceed 2000 s'1. Concomitant with this reduction chemistry is the uptake of four chemical protons and four pump protons, which are translocated across the membrane. It is the net translocation of protons across the membrane that generates a membrane potential and is subsequently used for ATP synthesis [1-3]. Simply written, the chemistry catalyzed by cco can be presented by: 4e‘+8H,;,+02 —9 2r120+4H+ ()Ul ' The chemistry of 02 reduction and proton pumping centers on the involvement of a number of cofactors and specific amino acid residues. The structural framework for these cofactors is comprised of a number of protein subunits. The bacterial cco contains 3 - 4 subunits and the mammalian oxidase from bovine heart contains 13 subunits. Three different forms of cco proteins 147 have been structurally characterized. These are the bovine heart [4, 5], Paracoccus denitn'ficans (P. denitn'ficans) [6, 7], and Rhodobacter Sphaeroides (lwata unpublished results) proteins. A schematic of subunits l-lll from Rhodobacter Sphaeroides (R. Sphaeroides) is shown in Figure 6-1. Thus far the function of the additional 9- 10 subunits found in the mammalian oxidase is unclear. They do not appear to affect the chemistry of 02 reduction, but may eventually be found to have regulatory roles or aid in protein stability. Figure 6-1 Cytochrome c Oxidase from Rhodobacter Sphaeroides Simplified reaction scheme of membrane-bound cco including subunits I, II and III. The electron transfer steps are indicated by arrows, starting at cytochrome c, to the binuclear CuA site, to heme a, and finally to the binuclear center (heme a3 and CUB). Two proton uptake pathways are shown starting at residues K362 and D132. Other key residues are listed for reference. Oxygen binding, proton exit and water exit pathways are also shown. 148 The rapid rate of 02 reduction by cco directly involves four cofactors that participate in a sequence of long-range electron transfer events. Electron transfer is initiated by electron donation from cytochrome c to the dinuclear copper A (CuA) site. This is followed by subsequent electron transfer to the heme a, and then into the heme a3 - CuB binuclear site where 02 reduction and bond cleavage is catalyzed. The redox potentials (vs. NHE) for the metal centers are accepted to be 0190-0245 V for CuA, 0.190-0.210 V for heme a, 0350-0385 V for heme 33 and 0.350 V for CuB, as determined by spectroelectrochemical [8], redox titration [9] and direct electrochemical [10-12] methods. The initial binding of 02 to the reduced form of the enzyme (states: CuA', Fea”, Fea3", CuB') results in formation of the heme Fea3H—Oz state (intermediate A), which is then converted via a Fe33IV=O intermediate (Pr) to the oxidized state of the enzyme (CuA', Fea', Feagm-OH, CuB'l-OH). The electron inventory between the fully oxidized form and the fully reduced form of the enzyme is 4 and the kinetic steps have been resolved with optical and vibrational spectroscopies [1, 2, 13, 14]. Other metal ions present that do not participate directly in electron transfer include Mg, Ca and Zn ions. The requirement for, and, therefore, the role of these metals in 02 chemistry is unclear. Coupled with the 02 electron transfer reaction is the catalytic pumping of four protons across the membrane and the requirement of four protons as the substrate 02 is reduced to H20. Two proton transfer pathways, termed the K 149 and D channels, have been identified in subunit l for facilitating the movement of protons from the interior side of the membrane towards the binuclear site. The K channel pathway is termed for an essential lysine residue K(l-362) (R. Sphaeroides numbering) and the D channel for an essential aspartate residue D(I-132) that leads to glutamic acid E(l-286). The elucidation of these pathways is derived from structural information [4-7] and from mutational and kinetic measurements [15-17]. Further pathways for the pumped protons have also been proposed [5, 18], as have pathways for the substrate 02 [19-21] and product H20 [22-24]. Recent mechanistic emphasis with the heme-Cu oxidase enzymes has been directed at understanding the important bioenergetic question concerning the coupling of electron transfer with proton pumping. In particular, questions remain regarding the proton pathways and which reaction steps of 02 reduction are coupled to H+ [25-28]. In order to pump protons against the electrochemical gradient it is commonly held that amino acid groups undergo a pK change in response to metal ion reduction-oxidation steps. In cco, a small number of amino acid residues have been specifically implicated in this process. One experimental approach offering insight into the involvement of amino acid residues in the catalysis is Fourier Transform Infrared (FTIR) difference spectroscopy, as it offers the sensitivity to reveal single amino acid changes in response to protonation, ligation, and redox state changes of an enzyme. The FTIR approach can be addressed in a number of ways, from a kinetic to a static perspective. One approach is to follow events initiated by the 150 photolysis of CO from heme a3, and at low temperatures (< 140K), the transfer of CO to CuB. This approach has been used by a number of groups to study the binuclear site [29-32] and has also been used successfully to study changes in a time-resolved manner [33-35]. Another approach examines changes between the fully oxidized and reduced states of cco and involves the electron transfer steps of Cu and heme a, in addition to the chemistry of the binuclear site. This approach lies solely in the physiological temperature regime, and may be initiated via several means, including the photochemical reduction of cco [36, 37], the indirect redox titration of the enzyme [38], or additionally, the pioneering approach of electrochemical titration (accelerated with redox mediators) of cco [39-42]. In the future, it may be possible to perform these spectroelectrochemical experiments with cco directly anchored to an electrode [12]. This chapter discusses the application of electrochemical difference FTIR to examine cco from R. Sphaeroides and bovine heart. Isotope labeling experiments were performed to simplify the spectral assignments. Global 15N labeling of cco of R. Sphaeroides, together with deuterium exchange, has identified a histidine signal in oxidized minus reduced FTIR spectra. A positive 1 upon 15N exchange and to band at 1105 cm", which shifted to 1095 cm' 1104 cm'1 upon deuteration, was assigned to the C-N stretching mode of a histidine side chain. The relatively high frequency at 1105 cm'1 and the observed insensitivity to deuteration indicated that the imidazole ring of this histidine has an N6 protonated form. Definitive assignment of this mode to a specific histidine 151 side chain can only be achieved by further experiments to measure the FTIR spectra of R. Sphaeroides samples containing specifically labeled 15N histidine residues and site-directed mutants replacing histidine. 6.2 Results 6.2.1 Electrochemical Difference Spectra The oxidized and reduced forms of R. Sphaeroides cco were obtained by electrochemical poise of the enzyme in the OTTLE cell facilitated by redox mediators. The UV/vis optical spectra of the fully oxidized and reduced forms of the enzyme were then recorded at potentials of +0.50 V and —0.50 V (vs. Ag/AgCl), respectively. The optical properties of cco report on the oxidation state of the heme a, heme a3, and CuA metal centers, as well as the coordination state of the heme iron atoms. The UV/vis spectrum of cco contains two major bands, the Soret and d-bands The most intense of the two peaks is the Soret band, which occurs between 400 - 450 nm, depending on the nature of the heme iron and the state of the enzyme [43]. This band is attributed to the rr—>n* transitions of the hemes, and is therefore a very sensitive indicator of the heme iron oxidation state. The or-band (~ 606 nm) is of much lower intensity, and is due to n—m" transitions of heme a. 152 Figure 6-2 UV/vis Electrochemical Difference Spectra of Rhodobacter Sphaeroides Cytochrome c Oxidase Optical UV/vis electrochemical difference spectra of the oxidized minus reduced (solid line) and reduced minus oxidized (dashed line) forms of cco from Rhodobacter Sphaeroides. The oxidized and reduced spectra were acquired at +0. 50 and —0. 50 V (vs. Ag/AgCI), respective/y. Potentials were applied externally via a potentiostat and electrochemistry was accelerated with a series of redox mediators—see Table 2-1 for details. l0 3 AA 0.05 Wavelength (nm) The oxidized minus reduced (solid line) and reduced minus oxidized (dashed line) difference spectra are shown in Figure 6-2. When fully oxidized, the Soret maximum occurs at 423 nm, and when fully reduced, the Soret maximum shifts to 445 nm and a prominent d-band at 606 nm appears. In the oxidized minus reduced difference spectrum, these features are manifested as a positive peak at 423 nm representing the oxidized form and negative peaks at 445 and 606 nm corresponding to the reduced state. These features are the same as reported earlier [44] and are indicative that the direct electrochemical titration of this enzyme is achievable. It is important to note that the response is stable in time and reproducible with repeated cycling. The two difference spectra 153 are mirror images of the other, as expected for a reversible and stable electron transfer process. After obtaining a UV/vis difference spectrum, the enzyme sample was placed into the FTIR cryostat, the temperature was equilibrated, and electrochemical difference spectra were recorded. Details of the cryostat and temperature control setup were provided in Chapter 2 (section 2.5). Figure 6-3A shows the noise spectrum acquired for this experimental setup. This spectrum is obtained by measuring two consecutive spectra at a constant potential. The average peak-to-peak noise level is below 1 x 10'5 absorbance units. The oxidized minus reduced (solid line) FTIR spectrum for bovine cco for a potential step from -0.50 to 0.50 V is shown in Figure 638. The positive bands in this spectrum correspond to the oxidized state and the negative, to the reduced state. The difference spectrum for the reverse process, (i.e., for a 0.50 to -0.50 V step), the reduced minus oxidized spectrum (dashed line), is also shown in Figure 638, for comparison. As was observed for the optical difference spectra, these spectra are the mirror image of each other, indicating a quantitative, fully reversible electrochemical reaction at the electrode surface. Examination of the background sample containing only the mediators and HEPES buffer, shown in Figure 6-3C, reveals minimal spectral contribution that would interfere with the cco FTIR difference spectrum. However, a small region in the cco difference spectra, shown in Figure 638, between 1700 - 1600 cm'1 and 1040 - 1000 cm'1, appears to exhibit reduced symmetry. These regions coincide with the strong underlying protein/buffer absorption and are indicative of a small baseline artifact. 154 This origin of this baseline phenomena is likely derived from a non-linearity in the response of the MCT detector, and although apparent, is not of any consequence to the findings of this work. In the oxidase difference spectra, a number of vibrational bands can be seen to change in the 1800 - 1000 cm'1 mid-IR regions. Such peaks arise from protein vibrational changes in response to the oxidation and reduction of the Cu and heme Fe centers. The changes in the FTIR difference spectra are detailed, and in many instances correspond to AA < 104. The most intense band in the absolute absorbance spectrum of a protein is the Amide l peak at ca. 1658 cm'1, which represents the nature of the peptide backbone structure via C=O vibrations. Comparing the total absorbance of the protein at this peak, which is ~ 0.7 absorbance units, to that of the strongest band (1662 cm'1) in the difference spectrum (AA ~ 1 x 10'3), it is apparent that the majority of the protein acts as an inert scaffolding during the redox process, undergoing no gross conformational change in the peptide backbone structure. Instead, the FTIR spectrum of the fully oxidized minus reduced state of cco corresponds to prominent changes in a small number of amino acid residues that undergo a change between the oxidized and reduced forms [45]. A significant contribution to the electrochemical difference spectra is also derived from the two heme groups [39, 42]. 155 Figure 6-3 FTIR Electrochemical Difference Spectra of Rhodobacter Sphaeroides Cytochrome c Oxidase Electrochemical FTIR difference spectra (1800 - 1000 cm‘1) recorded at 10 0C. (A) Oxidized minus oxidized noise spectrum. (8) Spectrum of the cco enzyme from bovine heart as the oxidized minus reduced (solid line) and reduced minus oxidized (dashed line) difference spectrum (average of 3 samples). (C) Spectrum of the mediator cocktail and the sample buffer (no enzyme) as an oxidized minus reduced difference spectrum. All spectra were acquired using BaFg windows and a 10 pm pathlength. Potentials of +0. 50 V and -0. 50 V (vs. Ag/AgCI) were applied to obtain the oxidized and reduced states, respectively. A L—u—W ‘W-‘A‘ —==~W E Amide I g ‘ 6(N H) v(C=0) 5 1548 v(CO),5(COH) 1537 Tit v(CN) His v(CC) ,5(CH) Tyr -3 M11 x 10 Amide I v(C=0) C WWW-MM 1800 1600 1400 1200 1000 Frequency (cm'1) 156 As illustrated in Figures 6-3 and 6-4, the mid-IR spectrum can be broken into distinct regions of interest, according to structural moieties. Exclusive contributions from the carbonyl stretching mode of protonated Asp and Glu amino acid side chains are expected in the spectral region above 1700 cm'1 [46]. These modes are sensitive to H/D exchange, experiencing downshifts of 8 - 10 cm'1, as well as to hydrogen bonding, with shifts up to 50 cm'1 observed upon formation of strong hydrogen bonds [46]. In the Amide l range (1680-1620 cm'1), strong signals at 1684, 1662, 1646 and 1634 cm'1 are attributed to changes in the carbonyl stretching modes of the polypeptide backbone. These modes occur below 1700 cm'1 due to the strong hydrogen bonding experienced by the backbone. The relatively small amplitude of these signals indicates subtle structural rearrangements of individual peptides upon the redox process. Additionally, contributions from the formyl and propionate groups of heme a and heme a3 [29, 47-50], and from individual amino acid side chains (Asn, Gln, and Arg) are expected in this region [39, 46]. A number of peaks are observed in the region from 1580 to 1515 cm'1 (Amide lI range). Signals due to the coupled CN stretching and NH bending modes of the protein backbone, the Amide Il modes, are expected in this region [51]. In addition to the Amide Il modes, vibrational modes of heme C=C groups [29, 47-50] and aromatic amino acid side chains (Tyr, Phe, and Trp). as well as asymmetric COO' modes of Asp, Glu or heme propionates, also contribute in this range [39, 46]. In particular, a coupled C=C stretching and C-H bending mode of 157 Tyr occurs near 1517 cm'1 [46]. Additionally, Lys residues contribute a weak asymmetric deformation vibration of the NH3+ group near 1526 cm'1 [46]. The deprotonated Asp, Glu and heme acid symmetric carboxyl modes are expected between 1450 and 1400 cm“. Figure 6-4 Illustration of Characteristic Mid-IR Bands A schematic depiction of some characteristic absorption bands observed in the mid-IR range (1800 - 1000 cm'1) of protein F T/R spectra. Bending modes are indicated by the symbol 6 and stretching modes by the symbol v. Asymmetric and symmetric modes indicated by as and s, respectively. Asp. Glu. heme-COOH vS Lys. 1525 (red) 1522 (red) Amide II a This work. bFrom reference [54]. CFrom reference [36], not observed in this work. A mode at 1537 cm'1 is observed in both bacterial spectra (1538 cm”1 in P. denitrificans), but is absent in bovine electrochemical difference spectra (this work and [54]). However, this mode has been observed for bovine CO photolysis difference spectra [29, 33]. It is possible that the different experimental 162 procedures result in structurally different forms of the enzyme, or that the extinction coefficient of this mode changes. A comparison of the carbonyl spectral region above 1700 cm"1 for the CO photolysis and electrochemical difference procedures also reveals differences, suggesting that subtle differences in enzyme structure result from the different approaches (this work and [33, 54]). Based on the observed results from this work and those in the literature, as well as from characteristic group IR frequencies for amino acids and protein vibrations, several assignments for the mode at 1537 cm'1 are possible. The asymmetric COO' stretching mode of Asp, Glu, the C-terminus, or heme propionates, the Amide ll mode, and the symmetric NH3+ deformation mode of Lys or the N-terminus are possible candidates (Figure 6-4). At first thought, the large peak amplitude, and the 16 cm'1 downshift observed upon exchange with 15N, suggest it is due to a contribution from the Amide ll mode. However, Hellwig and coworkers have shown that a peak at 1538 cm'1 in the P. denitrificans cco enzyme was sensitive to 13C heme labeling and have assigned it to 3 COO” asymmetric stretch [42]. The presence of a corresponding COO" symmetric stretching mode near 1416 cm'1 further supported their assignment. Comparisons of bovine cco resonance Raman data to that of Cu-bound imidazole complexes, suggested that a mode at 1537(8) cm’1 would arise from a Cu-His coupled vibration [29, 33]. The observed 15N isotope shift in this work would support an assignment of this peak to a N-containing His vibration. It is likely that the true assignment of the peak at 1537 cm'1 is a combination of the 163 possibilities listed above, rather than that of a single mode. On account of the relatively large amplitude of this mode, it is possible that there are several smaller modes overlaid at the same frequency. Noguchi and coworkers observed this situation for light-induced FTIR difference spectra of the photosystem II enzyme [59]. Isotope labeling experiments (13C and 15N) and the presence of the corresponding symmetric mode supported the assignment that a mode near 1550 cm'1, in the Amide Il region, was actually an asymmetric COO“ stretching mode, overlaid by an Amide ll mode [59]. 1 A peak at 1517 cm' in the reduced form of both the bovine and R. sphaeroides is attributed to a tyrosine ring stretching mode, confirming the previous assignments by Hellwig in P. denitrificans [60] and Rich and Breton in bovine [36]. An interesting question then arises, as structural data have shown a conserved tyrosine that is covalently linked to His 333 of CuB located near the binuclear center [4] that has been implicated in both the binding of intermediates [61] and in proton transfer [62, 63]. Could the mode at 1517 crn'1 be attributed to this unique residue? This cross-linked tyrosine would be influenced by the redox transitions of the binuclear center, and would, therefore, contribute to an electrochemical difference FTIR spectrum. Measurements in 020 indicate that this mode is sensitive to H/D exchange, suggesting an exchangeable protein residue. Comparison of characteristic bands of the tyrosine side chain to those of a model tyrosine-histidine compound are helpful in determining if peaks originate from the cross-linked residue. 164 The neutral tyrosine side chain exhibits strong characteristic modes at 1518 cm'1 and at 1249 cm'1 for a coupled ring CC stretching mode, CH bending mode, and a coupled CO stretching-COH bending mode, respectively [46]. The CC ring stretching mode has been observed at 1518 cm'1 and the CO stretching and COH bending mode as a broad peak at 1282 cm'1 in the FTIR spectrum of a synthetic model of the cross-linked histidine tyrosine side chain, 2-imidazole-1-yl- 4-methylphenol [60]. From these examples, it is clear that the mode at 1517cm’1 is not strongly influenced by the presence of the cross-linked imidazole, but that by examining the region from 1250-1300 cm'1 one may distinguish between the traditional and the cross-linked forms of tyrosine. It seems that the problem is not so easily solved in the complex oxidase difference spectrum. However, site-directed mutants of the cross-linked tyrosine, combined with the known IR spectra of tyrosine and the synthetic cross-linked tyrosine have led Hellwig and coworkers to assign a feature at 1268 cm’1 in the P. denitrificans cco spectrum to the cross-linked Tyr284 residue [60]. In this work peaks at 1264 and 1269 cm'1 are observed for the oxidized forms of R. sphaeroides and bovine cco, respectively. It is likely that these modes correspond to that observed by Hellwig in the P. denitrificans cco spectrum, but without further site-directed mutagenesis data, it is not possible to say so definitively. There are a few striking discrepancies observed in Figure 6-5, with respect to Hellwig and coworkers’ study. For instance, the tentative assignment 165 of a reduced mode at 1588 cm"1 to the asymmetric carboxyl stretch of an exclusive (i.e., not conserved in bacterial sources) mammalian residue, Asp51, [54] seems unlikely, since this mode is present at 1586 cm'1 in both spectra shown in Figure 6-5. Rich and Breton also point out that it would be unusual for this residue to be deprotonated in the reduced form [36]. 6.2.3 Isotopically-Labeled Spectra The complexity of assigning vibrational bands can be reduced with global isotopic labeling experiments of the cco protein. Figure 6-6 presents the oxidized minus reduced FTIR difference spectra from R. sphaeroides (Figure 6-6A) and bovine (Figure 6-6B) oxidase samples following deuterium buffer exchange, and Figure 6-7 the R. sphaeroides oxidase sample following global 15M labeling. In both cases the labeled spectra (dashed lines) are overlaid with the corresponding unlabeled spectra (solid lines). The deuterium labeling reveals a number of small spectral changes as a consequence of the H/D exchange. One region to change 1 significantly upon deuteration is the C=O region at 1740-1730 cm” for R. sphaeroides. However, relatively few H/D changes occur within the strong IR bands that overlay the Amide I and Amide ll regions, 1689-1660 and 1550 - 1500 cm'1, respectively. 166 Figure 6—6 Comparison of the FTIR Difference Spectra of Cytochrome c Oxidase Measured in 020 and H20 Electrochemical oxidized minus reduced FTIR difference spectra (1800 - 1000 cm'1) of cco from (A) Rhodobacter sphaeroides and (8) bovine heart. The spectra of the global deuterium labeled enzyme (dashed line) are overlaid with the unlabeled enzyme (solid line). To aid the comparison, the spectra from the deuterated samples were normalized to those from the non-deuterated samples based on the intensity of the 1540 cm'1 band. Spectra were recorded at 10 0C using BaF2 windows and a 10 pm pathlength. Potentials of +0. 50 V and -—0. 50 V (vs. Ag/AgCl) were applied to obtain the oxidized and reduced states, respectively. M15x10'4 will L l l l 1 l l 1800 1700 1600 1500 1400 1300 1200 1100 1000 Frequency (cm-l) In another isotope approach, global 15N labeling of the R. sphaeroides enzyme was performed to elucidate the involvement of N modes. This method facilitates the assignment of N-coupled vibrations potentially arising from lysine, arginine, asparagine, glutamine, and histidine amino acid side chain modes, as well as heme modes. Figure 6-7 shows the 15N (dashed line) and unlabeled 167 (solid line) oxidized minus reduced spectra from R. sphaeroides. A list of selected 15N sensitive peaks and associated frequency shifts are presented in Table 6-2. The labeled minus non-labeled (double difference) spectrum (Figure 6-7 inset) indicates a number of 15N sensitive modes in the 1615 -1600 cm‘1 region. However, the most obvious changes appear in the 1550 - 1500 cm'1 region, which are coincident with the Amide ll bands Figure 6-7 Effect of Global 15M Labeling on Mid-IR FTIR Spectra of Cytochrome c Oxidase Electrochemical oxidized minus reduced FTIR difference spectra of Rhodobacter sphaeroides cco unlabeled (solid line) and 15N globally labeled (dashed line). The insert shows the double difference spectrum (15N minus 14N) from 1625 - 1470 cm'1. Spectra were recorded at 10 OC using BaFg windows and a 10 pm pathlength. Potentials of +0. 50 V and —0. 50 V (vs. Ag/AgCI) were applied to obtain the oxidized and reduced states, respectively. , Mlsxm" 15100 ‘ A 1151001 I I I l 1 I L I 1 1 1 I 4 I i 1800 1600 1400 1200 1000 -1 Frequency (cm ) 168 Table 6—2 15N Sensitive Peaks in Rhodobacter Sphaeroides Cytochrome c Oxidase FTIR Spectra Redox Frequency in H20 Frequency in 15N State spectrum (cm'1) spectrum (6'04) AV (cm'1) Ox 1560 ? Red 1548 1534 14 Ox 1538 1522 16 Red 1525 151 1 14 OX 1498 1482 16 Ox ? 1428 OX 1 105 1095 10 Figure 6-8 presents an enlargement of the Amide Il region showing the effect of 15N labeling in greater detail. However, as discussed above, the HID exchange experiments shown in Figures 6-6 and 6-8, and reported elsewhere [54] show relatively little change to the Amide ll region and no change to the large 1550 cm'1 mode. The Amide lI modes are expected to downshift ~100 cm'1 upon D exchange, while the Amidel modes are less sensitive to deuteration [51]. The Amide l signals, which are attributed to reorganization of 01 helical and [3 sheet secondary structure elements, are expected to undergo small shifts (2- 10 cm”) after H/D exchange, largely arising from changes in the hydrogen bonding status [52]. This finding, from Figures 6-6 and 6-8, that there is little H/D effect on the Amide bands has also been observed in previous FTIR difference measurements of cco from a variety of sources [40, 64]. However, the 169 result is somewhat surprising when compared to the FTIR difference spectra of other membrane proteins [59] and synthetic peptides [65] that have shown more substantial (2 20 cm'1) H/D effects to the Amide II bands. Figure 6-8 Isotope Exchange Effects in the Amide lI Region Electrochemical oxidized minus reduced FTIR difference spectra (1600 - 1400 cm'1) of Rhodobacter sphaeroides cco unlabeled (solid line), deuterium labeled (dashed-dotted line) and 15N globally labeled (dashed line). Spectra were recorded at 10 0C using Ban windows and a 10 pm pathlength. Potentials of +0. 50 V and —0. 50 V (vs. Ag/AgCI) were applied to obtain the oxidized and reduced states, respectively. AAI2x10‘ ii I i i 1600 1550 1500 1450 1400 Frequency (cm'1) The virtual absence of shift in the Amide ll bands upon deuteration has lead earlier work to concede that there was little or no significant Amide contribution to cco FTIR difference spectra. Hellwig and coworkers have assigned many of the modes in the region from 1560 to 1520 cm'1 to asymmetric stretches of carboxyl groups of Asp or Glu residues and heme propionates, based on the small shifts observed upon deuteration [39]. Alternate explanations 170 for the small observed shifts include a lack of solvent accessibility to the protein backbone due to a hydrophobic environment or a tightly ordered structure (a helix or [3 sheet) [51]. This effect is observed for the core region of bacteriorhodopsin, which, in contrast to the peripheral region, is largely inaccessible to HID exchange [66-68]. This raises an interesting question as to whether significant deuterium exchange is experienced in the core of the cco enzyme following lengthy 020 incubation. Based on measurement of the Amide II, and deuterated Amide ll, Amide Il', bands in the absolute spectra (results not shown), the D20 incubation significantly reduced the intensity of the Amide II band at ~ 1550 cm'1 and was replaced with an Amide II' band at ~ 1450 cm‘1. This suggests that there was significant global deuterium labeling of the cco enzyme. It is estimated that > 80% of the enzyme was deuterated. However, precise estimation of protein in these regions because of the closely overlaying H-O-H and D-O-H water modes. As the 15N data reveal an indisputable 10 cm'1 downshift of the 1550 cm"1 band (and neighboring bands), it seems likely that in cco, deuterium exchange of the inner core is kinetically limited beyond 3 days It is probable that this inner core is responsible for the Amide l and Amide lI changes observed in the difference spectra. A low level of exchange (~ 10%) occurred for the inner core, despite lengthy incubation and concentration efforts, which took the best part of 48 h. An interesting deuterium effect is observed for the peripheral and core regions of bacteriorhodopsin. The peripheral region undergoes rapid HID exchange of backbone peptide groups. However, reconstitution and refolding in 171 D20 is required to observe a complete downshift of the Amide II modes from the core region, which consists of primarily d-helical structure largely buried in the lipid bilayer [69]. Therefore, it is conceivable that a similar process, in which the inner core protons are shielded from exchange, is occurring in cco. Such shielding would result in non-exchange of the core peptide backbone, which gives rise to the Amide ll modes. As discussed above, based on shifts observed upon 15N exchange, it is clear that the Amide ll modes do contribute to the oxidized minus reduced electrochemical difference spectra of cco. Nitrogen isotope exchange experiments are novel for this enzyme and they are critical to providing validation for tentative mode assignments. For example, a marked (14 cm'1) downshift was observed for the reduced mode around 1525 cm’1 upon 15N labeling, suggesting that the previous assignment of this mode exclusively to a C00' asymmetric stretching mode of a heme propionate [54] is unlikely. It is more conceivable that this mode is instead a combination of an Amide ll vibration of the peptide backbone and a weakly intense asymmetric stretching mode of the heme of nearly the same frequency that is overlaid by the large Amide ll mode. Additionally, a peak around 1634 cm'1 present in the reduced form of all three cco samples was assigned to a symmetric stretch of an arginine side chain [54] and in a recent report, it has been assigned to the 8(NH2) mode of asparagine [60]. However, the 15N labeling experiment revealed no shift of this peak, indicating that it is not attributable to a nitrogen-containing vibrational mode. 172 Both of these modes would be very sensitive to deuteration, with expected shifts of greater than 10 cm'1, which are not observed [46]. An alternate explanation is that the reduced peak around 1634 cm’1 is due to an Amide l backbone mode. 6.3 Discussion 6.3.1 Carbonyl Region The FTIR carbonyl region (1750 - 1700 cm'1) of cco reveals a number of changes arising from protonated Asp and Glu carboxylic acid residues. Figure 6-9 shows the spectral expansion of some of the potential carbonyl vibrations in cco spectra. The R. sphaeroides spectra obtained in H20 (solid line) and D20 (dashed line) are presented in Figure 6-9A and the H20 (solid line) and D20 (dashed line) bovine spectra in Figure 6-98. A difference signal at 1746/1735 cm‘1 is observed in the bacterial H20 spectrum that shifts to 1740/1730 cm"1 in DZO. A similar, deuterium-sensitive peak is observed in the oxidized form of the bovine sample at 1748 cm"1 and the corresponding reduced peak is absent. An additional positive peak at 1734 cm'1 is observed in the bovine spectrum. Small shifts between the two forms of the oxidase are not surprising, and are indicative of subtle perturbations of the environment surrounding the carbonyl group. The reversal of a mode, such as that observed for the peak near 1735 cm'1, can be explained as the contribution of an additional Asp or Glu residue or a change in the extinction coefficient as a 173 result of a different environment. Figure 6-9 Deuterium Exchange Effects in the Carbonyl Region Electrochemical oxidized minus reduced FTIR difference spectra carbonyl region (1760 - 1715 cm”) of cco from (A) Rhodobacter sphaeroides and (8) bovine heart. The spectra of the global deuterium labeled enzyme (dashed line) are overlaid with the unlabeled enzyme (solid line). To aid the comparison the spectra from the deuterated samples were normalized to that of the non-deuterated samples based on the intensity of the 1540 cm'1 band. Spectra were recorded at 10 0C using Ban windows and a 10 pm pathlength. Potentials of +0. 50 V and -0. 50 V (vs. Ag/AgCI) were applied to obtain the oxidized and reduced states, respectively. -1740 ~11746 o ’’’’’ ..1734 - .."".,."'.. ...U 1748 1741 1755 1745 1735 _1 1725 Frequency (cm ) Hellwig and coworkers made the first tentative assignment of an amino acid mode in the carbonyl region of a cco FTIR electrochemical difference spectrum. Based on mutagenesis experiments, this residue, Glu 286, which is 174 situated at the end of the D-channel proton pathway, has been assigned at 1746 cm'1 in the oxidized enzyme and to a neighboring counter band at 1734 cm'1 in the reduced enzyme [40]. A similar difference pattern at 1745/1735 cm'1 (oxidized/reduced) has been identified in a related ubiquinol oxidase, cytochrome bo3 oxidase, from Escherichia coli by Liibben and coworkers [38]. The difference bands at 1745/1735 cm’1 were lacking in studies of site- directed mutant samples in which the glutamic acid residue was replaced by an aspartic acid or a glutamine at position 286, indicating they were clearly related to Glu 286 [38]. In photolysis-induced difference FTIR studies of bovine cco, Rich and coworkers observed a 1749/1741 cm’1 feature in the CO photolysis spectrum and a related 1752/1740 cm'1 feature in the CN photolysis spectrum, which also contained features at 1725/1715 and 1708/1698 cm’1 [36]. As mentioned earlier, it is conceivable that the two different approaches (photolysis and electrochemistry) result in slightly different enzyme structures, which are manifested as variations in the amino acid environment, thus yielding different peak frequencies. The question of assignment seems simple, as all three oxidase forms contain a conserved glutamic acid residue (Glu 286), which has been proposed to play an integral role in the proton pumping process and is, therefore, linked to the electron transfer steps. However, the task of definitively assigning these features to Glu 286 has yet to be resolved. The reported deuterium-induced 175 shifts are consistent with assignment to carboxylic acid species, although, in many cases, the magnitudes of the changes are small, if observed at all [36, 40, 54, 64]. The complexities observed in this region may be attributed to multiple carboxylic acids, or the same residue in different states. However, all authors confidently maintain that the conserved Glu 286 is a major contributor in both ligand and redox difference FTIR spectra. Questions remain regarding whether Glu 286 exists in several conformations that are affected differently by H/D exchange, and in identifying other residues that would contribute carbonyl stretching modes in this region. Likely candidates include heme propionic acids and carboxylic acids associated with the magnesium site above heme a, all of which are in close proximity to the redox-active metal centers. However, more remote residues may be considered, especially those implicated in the proton pumping function, such as those associated with the D channel [4, 70] or the conforrnationally flexible Asp 51 [5]. 6.3.2 Histidine Modes Two technical improvements over previously published methods have extended the accessible frequency range to include the region below 1200 cm'1. First, using protein solutions prepared in HEPES instead of phosphate buffer removed the large phosphate bands around 1100 cm'1 that were obscuring the spectrum in this region. Secondly, BaF2 windows were implemented instead of Can, which increased the accessible range from 1100 to 800 cm'1. Amino acid modes expected to be present in the 1200 - 1000 cm‘1 region include the C-N 176 stretch of histidine, which has been proposed to be an important residue in the catalytic function of the enzyme. Bioinorganic complexes with histidine ligands have been reported to exhibit characteristic FTIR signals of the imidazole ring around 1100 om‘1 [71-78]. Figure 6-10 Closer Examination of a Possible Histidine Mode Electrochemical oxidized minus reduced FTIR difference spectra (1150 - 1050 cm'1) of cco from (A) Rhodobacter sphaeroides and (B) bovine heart. The spectra of the global 15N labeled enzyme (dashed line) and the global deuterium labeled enzyme (dashed- dotted line) are shown with the unlabeled enzyme (solid line). To aid in the comparison, the spectra from the deuterated samples were normalized to that of the non-deuterated samples based on the intensity of the 1540 cm’1 band. Spectra were recorded at 10 0C using BaFg windows and a 10 pm pathlength. Potentials of +0.50 V and —0.50 V (vs. Ag/AgCI) were applied to obtain the oxidized and reduced states, respectively. 1140 1120 1100 1080 1060 Frequency (cm-i) 177 Figure 6-7 depicts a 15N sensitive mode around 1100 cm‘1 in the R. sphaeroides sample. Figure 6-10 shows a magnified view of the region from 1150 - 1050 cm’1 for the mammalian cco samples (H20 and 020) compared to the three bacterial samples (H20, D20 and 15N). There is a slight (4 cm'1) difference in the position of this mode between the bovine and R. sphaeroides samples. This is not surprising, as similar results (i.e., small frequency discrepancies) were obtained in this work and others [54, 58], for spectral comparisons of different types of cco. Therefore, this small difference is attributed to slight changes in the environment of the structural element giving rise to this mode. More important are the shifts due to isotope exchange, as they provide insight into the source of this mode. Noguchi and coworkers have performed similar isotope exchange experiments in the photosystem II enzyme and model compounds of the histidine side chain (4- and 5-methyl imidazoles), in an effort to identify histidine modes [73, 74]. The results from this work are consistent with their findings, in that FTIR measurements revealed a mode around 1100 cm"1 that undergoes a 1 cm'1 downshift upon deuteration and a 10 cm'1 downshift upon 15N labeling. Therefore, this mode is attributed to a C-N stretching mode of a histidine imidazole ring. In a study of histidine modes, Noguchi’s group determined that 1 this C-N stretching mode around 1100 cm' was a useful IR marker band of the protonation state of the histidine imidazole ring [73]. 178 Figure 6-11 Structures of the Two Neutral Histidine Tautomers Representative structures and atom numbering for the two neutral forms of 4-methy/ imidazole, a histidine side chain analog. (A) The Na—protonated or 4-methyl imidazole form. (8) The Nt‘iprotonated or 5—methy/ imidazole form. (C) Structure of a metal-bound N6-protonated or 5-methyl imidazole form. Frequency data is provided based on vibrational spectroscopic (resonance Raman and F T/R) studies and OF T calculations for the HON) mode, and from resonance Raman studies and DFT calculations for the t(MN) mode. (A) (B) - HN 5 8 V" 4-Melm 5-Melm Ns protonated His N6 protonated His (C) \‘(CC) ~ 1600 cm'1 ( v(CN) ~ 1100 om'1 E HN8 Err. V (Metal v(MN) ~ 250 on1 The imidazole side chain of the histidine residue can exist as either of two neutral tautomers, depending on the protonation state of imidazole N. The model compounds 4- and 5-methyl imidazoles correspond to the Ne and N6 protonated forms of DL-Histidine. The structures of these molecules are shown in Figure 6-11. The two different forms of the side chain exhibit unique FTIR spectra. The spectrum of the 4-Melm form consists of a characteristic C-N stretching mode centered at 1087 - 1093 cm'1 depending on temperature, which 179 upshifts by 5 - 10 cm'1 upon N-deuteration. The 5-Melm form, however, exhibits a C-N stretching mode at slightly higher frequency (1103 - 1106 cm“) and is relatively unaffected by deuteration (-1 cm'1) [73]. Studies of FePP(4-Melm)2 [76], cytochrome b559 of PSII [76] and non-heme iron of PSII [75, 77, 78] demonstrate similar results for cases where the imidazole ring is coordinated to a metal. The data from these and related studies are presented in Tables 6-3 and 6-4. The observed N-deuteration effect can be explained by normal mode analysis. Majoube and coworkers have assigned the bands around 1100 cm’1 of 4- and 5-Melm to the C5-N1 stretching vibration [79]. From the structures presented in Figure 6-11 it is clear that this vibration will be sensitive to deuteration at the N1 site in the Ne protonated form, but will be rather unaffected by deuteration at the N3 in the N6 protonated form. It is proposed that the upshifting observed in the C5-N1 band upon N1 deuteration results from decoupling of the mixed N-H bending vibration upon deuteration [79]. The C4C5 stretching vibration has also been used as a vibrational marker of the protonation state and metal binding status of histidine and 4-Melm molecules [71, 79-84]. Tables 6-5 and 6-6 list the C405 frequencies found in the literature for metal-free and metal-bound histidine and 4-Melm, respectively. The values presented are those corresponding to the N6 protonated neutral forms of 4-MeIm and histidine, as they are most relevant to the present discussion. 180 Table 6—3 C5N1 Stretching Modes for Histidine and 4-Melm Frequency (cm'1) Method Reference L-Hisa 1106 RR [71] DL-His (H/D)b 1093/1098 FTIR [73] 4-Me|m (H/D)C 1087/1097 FTIR [73] 5-Me.m (H/D)C 1103/1104 FTIR [73] 4-Melm (HID) 1075/1079 DFT calc. [80] 5-Melm (H/D) 1126/1125 DFT calc. [80] BA.a_Hisd 1 105 RR [72] aL-histidine in water (pH 11.02, 41 OC). bCrystalline DL-histidine measured as a KBr pellet. CIn aqueous solution at room temperature, 4-Melm and 5-Melm in tautomeric equilibrium. quueous solution of B-alanyI-L-histidine (pH 8.0). *‘5-Melm is equivalent to N6 protonated 4-Melm. Table 6—4 C5N1 Stretching Modes for Metal-Bound Histidine and 4-Melm Frequency (cm'1) Method Reference Co(lI)-(4-Melm)GClza 1112 RR [71] Cu(lI)-(BAla-His)b 1112 RR [72] Mn (psu OEC)C 1114-1113 FTIR [73] Fe(II)/Fe(|||) (Ps||)d 1111-1110/1103-1102 FTIR [75, 77, 78] Fe(lll)PP(4-Melm)29 1103 FTIR [76] Fe(lll) (cyt b559)’ 1104 FTIR I76] Zn(II)-(4-Melm)g 1108 DFT calc. [81] aCrystalline hexakis(4-methylimidazole)cobalt(Il)ch|oride. bCrystalline (b-alanyl-L-histidino) copper(ll)dihydrate. CThe Mn cluster of photosystem II. The non-heme iron of photosystem II. eBis(4-methylimidazole) complex of iron protoporphyrin in water. fOxidized form of cytochrome b559 of photosystem II. g(4-methylimidazole) zinc(ll) trihydrate complex with N1 metal coordination and N3 protonated. *‘5-Melm is equivalent to N6 protonated 4-Melm. 181 Table 6—5 C4C5 Stretching Modes for N6 Protonated Histidine and 4-Melm Frequency (cm'1) Method Reference L-Hisa 1585 RR [71] DL-Hisb 1588 RR [72] 5-Melmc 1594 FTIR [80] 5-Me|mc 1593 RR [80] 5-Melm 1593 DFT calc. [80] His (zinc finger)d 1585 RR [72] liAla-His" 1588 RR [72] aL-histidine in water (pH 11.02, 41 OC). bAqueous solution of DL-histidine (pH 8.0). CIn aqueous solution at room temperature, 4-Melm and 5-Melm in tautomeric equilibrium. quueous solution of a 27mer zinc finger peptide (pH 7.8). eAqueous solution of B-alanyl-L-histidine (pH 8.0). "5 Melm is equivalent to N6 protonated 4-Melm. Table 6—6 C405 Stretching Modes for Metal-Bound N6 Protonated Histidine and 4-Melm Frequency (cm'1) Method Reference Co(lI)-(4-Melm)eClza 1593 RR [71] Zn(ll)-(4-Melm)b 1615 DFT calc. [81] Zn-ABC 1604 RR [72] Zn-(zinc finger)d 1505 RR [72] Cu(II)-(8AIa-His)e 1594 RR [72] aCrystalline hexakis(4-methylimidazole)cobaIt(ll)chloride. 6(4-methylimidazole) zinc(ll) trihydrate complex with N1 metal coordination and N3 protonated. CInsoluble precipitate from a mixture of Zn(ll) and the human amyloid B-peptide. d27mer Zn finger peptide in the presence of ZnClz (6.25 mM). eCrystalline (b-alanyl-L-histidino) copper(ll)dihydrate. **5-Me|m is equivalent to N6 protonated 4-Melm. 182 The N6 protonated form contributes a C4C5 band at 1594 cm'1 to the aqueous solution FTIR spectrum of 4-MeIm [80, 84]. Free histidine in solution shows a band in the range of 1588 - 1580 cm’1 that has been attributed to the N6 protonated form [80, 84]. In both cases, the modes due to the N6 protonated form occur at higher frequency (15 - 20 cm'1) than those due to the Na protonated form [80, 84]. Raman spectroscopic [82] and DFT computational [81] studies report that metal binding to histidine upshifts the C4C5 frequency by as much as 21 cm'1. The C4C5 stretching mode is coupled to the ring NH bending mode, suggesting it would be sensitive to global 15N labeling. Our work shows evidence of 15N sensitivity in the region from 1605 - 1615 cm'1, which is where one would expect to observe a band due to the C405 stretching mode of a metal-bound N6 protonated histidine. However, this region also potentially contains a number of N-containing modes from the heme moieties. Once again, the complex nature of these spectra requires one to obtain additional information before definitively assigning a mode in the 1605 - 1615 cm'1 region to the C405 stretching vibration of a metal-bound N6 protonated histidine. Proposed experiments to measure the FTIR spectra of R. sphaeroides samples containing specifically labeled 15N histidine residues would assist in such an assignment. Global 15N labeling of cco of R. sphaeroides together with deuterium exchange has identified a histidine signal in oxidized minus reduced FTIR spectra. Examination of Figure 6-10 revealed a positive band at 1105 cm'1, 183 which was downshifted to 1095 cm"1 upon 15N exchange and to 1104 cm'1 upon deuteration, and was assigned to the C-N stretching mode of a histidine side chain. The structure of the histidine mode was determined using the above criteria. The relatively high frequency at 1105 cm'1 and the observed insensitivity to deuteration supported the assignment that the imidazole ring of this histidine had a N6 protonated form. We acknowledge that in the above experiments the N atoms were globally labeled with 15N, therefore, the possibility of shifted bands arising from structural moieties including N but not imidazole cannot be excluded. Definitive assignment of the mode at 1105 cm'1 to a C-N stretching vibration of a histidine side chain can only be achieved by a further experiment to measure the FTIR spectra of R. sphaeroides samples containing specifically labeled 15N histidine residues. For our preliminary analysis to be correct, we would expect to see the same result observed herein for the 1150-1050 cm"1 region and no shift observed around 1550 cm'1 (Amide ll). Investigation of the high frequency N-H stretching modes of the histidine side chain will provide further information regarding the H bonding status of the imidazole ring. When hydrogen bonded, the N-H stretching mode occurs as a broad band around 2800 cm'1 of weak intensity, but without hydrogen bonding, this mode is observed as a sharp band at 3500 cm'1 [79, 85-88]. Frequency regions above 2850 cm'1 are difficult to access, as strong water absorbance causes very high noise levels. 184 Studies of site-directed mutants of key histidine residues will also aid in the assignment of this mode to a specific amino acid. The histidine ligands of Cue, residues numbered His 333, His 334, and His 284, are believed to be important in the catalysis. Attempts to obtain site-directed mutants for these residues have been unsuccessful (i.e., the organism was severely disrupted and, therefore, did not grow) with the exception of His334Asn. 6.4 Implications Histidine is one of the most common ligands to metal ions in metalloproteins [89, 90]. This is due to the characteristics of the imidazole ring, which comprises the histidine side chain. The two nitrogens of the imidazole ring participate in metal binding, as well as protonation and deprotonation reactions. In the neutral form, either the Ne or N6 is protonated, and the other is free to act as a ligand for metal coordination. The metal-free states and the metal-bound states provide at least four possible forms that neutral histidine is capable of adopting. Not discussed here are several other charged states of histidine, both coordinated and not. With so many possible structures, it is not surprising that histidine often functions as a versatile ligand that is a key residue in a wide variety of catalytic reactions. For example, histidine is an integral part of the copper binding site of human serum albumin [91, 92], it is required for the formation of the zinc finger peptide nucleic acid-binding structural motif [93], it is involved in the reactions of enzymes such as the Zn, Cu superoxide dismutase enzyme [94], the photosystem II enzyme [73], and, as a proposed ligand to Cue, 185 it has been implicated in the proton translocation mechanism of the heme-copper oxidases [27]. Thus, it is important to know the coordination and protonation state of a histidine ligand, in order to elucidate the reaction mechanisms of many metalloenzymes. Table 6—7 Metal-Histidine Stretching Modes Frequency (cm'1) Method Reference Cu—N(His) 250 RR [95, 96] Fe-N(His) 220-250 RR [97, 98] Zn—N(His) 260 DFT calc. [81] As discussed above, experiments in the mid-IR have been successful at determining the protonation state of histidine. It is clear from Tables 6-3 and 6-4, however, that the histidine C5N1 vibrational mode does not provide a clear distinction between metal—free and metal-bound forms of histidine. Tables 6-5 and 6-6 indicated that such information might be obtained from the C4C5 stretching mode of histidine; however, at this time, it is not possible to assign such a mode in our spectra. Direct information regarding the metal ligation state of histidine is contained in His-metal modes, which occur at frequencies much less than 1000 cm'1, in the far-IR region. Table 6-7 lists His-metal modes observed in resonance Raman studies and normal mode calculations of bioinorganic model compounds. As indicated in Table 6-7, experimental access to low-frequency vibrations of metalloproteins has been limited to resonance Raman spectroscopy methods. 186 Resonance Raman spectroscopy is a powerful technique for studying many biological systems, especially cco. However, this technique is not capable of addressing the questions at all active sites in metalloproteins. For example, the Cue site of cco is “invisible" to resonance Raman methods due to interference attributed to the nearby hemes. Therefore, low-frequency electrochemical difference data obtained by IR for this metal center would be used to provide valuable information for answering mechanistic questions involving the role of Cue, in the catalytic cycle. As indicated in Chapter 5 (Table 5-1), this method would be useful to determine and directly monitor substrate or metal-ligand modes such as His-Cue or HO-CuB, that can provide insight into the chemical details of cco’s catalytic function. 6.5 Conclusions High quality, electrochemical difference mid-frequency (1800 - 1000 cm'1) FTIR spectra were obtained for several cco samples. A comparison of a bacterial and a mammalian form of the enzyme, as well as isotope exchange experiments involving D20 and 15N were carried out in an effort to simplify mode assignment in the complex spectra. Evaluation of the measured spectra for oxidases from bovine heart and R. sphaeroides sources revealed variations, due to both subtle and significant changes in enzyme active site structure. Many of the observed variations were consistent with a published comparison of bovine heart enzyme with that of a bacterial form from the P. denitrificans organism [54]. Although the oxidase structure and mechanistic function is highly conserved 187 between the bovine heart and R. sphaeroides forms of the enzyme, the observed spectroscopic deviations provide an indication of the subtle interactions the amino acid residues provide in facilitating electron transfer to the oxidase cofactors. Two important insights were gained through isotope global-exchange studies via solvent exchange with DZO buffers and organism (R. sphaeroides) growth in 15N media. First, these experiments revealed that the H/D exchange of the R. sphaeroides enzyme core was kinetically limited beyond three days, thus, it was very shielded from exchange; and, second, the results aided in the assignment of a mode at 1105 cm'1 to a histidine, which may be a residue that is integral in the cco proton pumping mechanism. The fact that the H/D exchange data show a lack of shifts observed for the Amide ll region (1580 -1520 cm"), despite expected shifts of up to 100 cm'1 upon deuteration, has led others to believe that the Amide II modes do not contribute to electrochemical difference FTIR spectra of cco [39]. However, this work clearly shows that upon 15N labeling, significant changes occur, as expected, in the Amide Il region, indicating strong contributions from the N-containing Amide ll modes to this region. Therefore, it was concluded that the Amide modes do contribute to the spectra in the region from 1580 to 1520 cm'1, and that the small observed shifts are instead due to a lack of solvent accessibility to the protein backbone caused by a hydrophobic environment or a tightly ordered structure (a helix or 8 sheet). Without further isotope data for the 188 bovine sample, it was assumed that the similarly small shifts observed upon H/D exchange are due to a similar shielding process of the exchangeable core protons of the bovine oxidase. The novel 15N global labeling experiments of this enzyme provided further insight into several previously assigned modes. Most importantly, it also facilitated in the identification of an unidentified histidine amino acid side chain. This particular residue is one that is linked to the electron transfer reaction of cco, and is potentially important in the proton pumping mechanism. Questions remain regarding whether or not this histidine is a direct ligand to the CUB of the active site. Further experiments are required involving 15His specifically labeled R. sphaeroides enzyme, in order to unequivocally assign this mode to a histidine side chain. Site-directed mutagenesis will aid in assigning this mode to a specific His of the protein structure, and low-frequency FTIR measurements of the cco enzyme will provide answers to questions involving its ligation to the Cue, and its involvement in the enzymatic mechanism. 189 6.6 [1.] [2.] [3-1 [4-1 [5-1 [6-1 [7-1 [8-1 [9-1 [10.] [11.] [12.] References S. Ferguson-Miller and G. T. Babcock Chem. Rev. 1996, 96, 2889-2907. H. Michel, J. Behr, A. Harrenga and A. Kannt Annual Review of Biophysics and Biomolecular Structure 1998, 27, 329-356. M. Wikstrdm Nature 1989, 338, 776-778. T. Tsukihara, H. Aoyama, E. Yamashita, T. Tomizaki, H. Yamaguchi, K. Shinzawaltoh, R. Nakashima, R. Yaono and S. Yoshikawa Science 1996, 272, 1136-1144. S. Yoshikawa, K. Shinzawa-Itoh, R. Nakashima, R. Yaono, E. Yamashita, N. Inoue, M. Yao, M. J. Fei, C. P. Libeu, T. Mizushima, H. Yamaguchi, T. Tomizaki and T. Tsukihara Science 1998, 280, 1723-1729. S. lwata, C. Ostermeier, B. Ludwig and H. Michel Nature 1995, 376, 660- 669. C. Ostermeier, A. Harrenga, U. Ermler and H. Michel Proc. Natl. Acad. Sci. U. S. A. 1997, 94, 10547-10553. J. L. Anderson, T. Kuwana and C. R. Hartzell Biochemistry 1976, 15, 3847-3855. N. A. Schroedl and C. R. Hartzell Biochemistry 1977, 16, 4961-4965. J. K. Cullison, F. M. Hawkridge, N. Nakashima and S. Yoshikawa Langmuir 1994, 10, 877-882. J. H. Li, G. J. Cheng and S. J. Dong J. Electroanal. Chem. 1996, 416, 97- 104. A. S. Haas, D. L. Pilloud, K. S. Reddy, G. T. Babcock, C. C. Moser, J. K. Blasie and P. L. Dutton J. Phys. Chem. B 2001, 105, 11351-11362. 190 [13.] [14.] [15.] [16.] [17.] [18.] [19.] [20.] [21.] [22.] [23.] [24.] [25.] [26.] T. Kitagawa and T. Ogura In Progress in Inorganic Chemistry, Vol45, 1997; Vol. 45, 431-479. M. Wikstrém Biochim. Biophys. Acta 1998, 1365, 185-192. P. Adelroth, H. Sigurdson, S. Hallén and P. Brzezinski Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 12292-12297. A. A. Konstantinov, S. Siletsky, D. Mitchell, A. Kaulen and R. B. Gennis Proc. Natl. Acad. Sci. U. S. A. 1997, 94, 9085-9090. T. V. Vygodina, C. Pecoraro, D. Mitchell, R. Gennis and A. A. Konstantinov Biochemistry 1998, 37, 3053-3061. D. A. Mills, L. Florens, C. Hiser, J. Qian and S. Ferguson-Miller Biochim. Biophys. Acta 2000, 1458, 180-187. S. Riistama, A. Puustinen, A. Garcia-Horsman, S. lwata, H. Michel and M. Wikstrbm Biochim. Biophys. Acta 1996, 1275, 1-4. I. Hofacker and K. Schulten Proteins 1998, 30, 100-107. S. Riistama, A. Puustinen, M. l. Verkhovsky, J. E. Morgan and M. Wikstrom Biochemistry 2000, 39, 6365-6372. B. Schmidt, L. Florens, L. Qin, J. McCracken and S. Ferguson-Miller Biophys. J. 2002, 82, 1391. L. Florens, B. Schmidt, J. McCracken and S. Ferguson-Miller Biochemistry 2001, 40, 7491-7497. Y. Georgievskii, E. S. Medvedev and A. A. Stuchebrukhov Biophys. J. 2002, 82, 2833-2846. H. Michel Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 12819-12824. H. Michel Biochemistry 1999, 38, 15129—15140. 191 [27.] M. Wikstrbm Biochim. Biophys. Acta 2000, 1458, 188-198. [28.] M. Wikstrbm Biochemistry 2000, 39, 3515-3519. [29.] S. Park, L. P. Pan, S. I. Chan and J. O. Alben Biophys. J. 1996, 71, 1036- 1047. [30.] A. Puustinen, J. A. Bailey, R. B. Dyer, S. L. Mecklenburg, M. Wikstrom _ and W. H. Woodruff Biochemistry 1997, 36, 13195-13200. E [31.] E. D. Dodson, X. J. Zhao, W. S. Caughey and C. M. Elliott Biochemistry 1996, 35, 444-452. [32.] Y. J. Zhen, B. Schmidt, U. G. Kang, W. Antholine and S. Ferguson-Miller i Biochemistry 2002, 41, 2288-2297. [33.] D. Heitbrink, H. Sigurdson, C. Bolwien, P. Brzezinski and J. Heberle Biophys. J. 2002, 82, 1-10. [34.] J. A. Bailey, F. L. Tomson, S. L. Mecklenburg, G. M. MacDonald, A. Katsonouri, A. Puustinen, R. B. Gennis, W. H. Woodruff and R. B. Dyer Biochemistry 2002, 41, 2675-2683. [35.] B. Rost, J. Behr, P. Hellwig, O.-M. H. Richter, B. Ludwig, H. Michel and W. Mantele Biochemistry 1999, 38, 7565-7571. [36.] P. R. Rich and J. Breton Biochemistry 2001, 40, 6441-6449. [37.] P. R. Rich and J. Breton Biochemistry 2002, 41 , 967-973. [38.] M. Li'lbben and K. GenNert FEBS Lett. 1996, 397, 303-307. [39.] P. Hellwig, S. Grzybek, J. Behr, B. Ludwig, H. Michel and W. Mantele Biochemistry 1999, 38, 1685-1694. 192 [40.] [41.] [42.] [43.] [44.] [45.] [46.] [47.] [48.] [49.] [50.] [51.] [52.] P. Hellwig, J. Behr, C. Ostermeier, O.-M. H. Richter, U. Pfitzner, A. Odenwald, B. Ludwig, H. Michel and W. Mantele Biochemistry 1998, 37, 7390-7399. P. Hellwig, B. Rost, U. Kaiser, C. Ostermeier, H. Michel and W. Mantele FEBS Lett. 1996, 385, 53-57. J. Behr, P. Hellwig, W. Mantele and H. Michel Biochemistry 1998, 37, 7400-7406. S. Horie and M. Morrison J. Biol. Chem. 1963, 238, 2859-2865. J. P. Hosler, J. Fetter, M. M. J. Tecklenburg, M. Espe, C. Lerma and S. Ferguson-Miller J. Biol. Chem. 1992, 267, 24264-24272. C. Zscherp and A. Barth Biochemistry 2001, 40, 1875-1883. A. Barth Prog. Biophys. Mol. Biol. 2000, 74, 141-173. P. V. Argade, Y. C. Ching and D. L. Rousseau Biophys. J. 1986, 50, 613- 620. S. Choi, T. G. Spiro, K. C. Langry, K. M. Smith, D. L. Budd and G. N. Lamar J. Am. Chem. Soc. 1982, 104, 4345-4351. S. Choi, J. J. Lee, Y. H. Wei and T. G. Spiro J. Am. Chem. Soc. 1983, 105, 3692-3707. S. Choi, T. G. Spiro, K. C. Langry and K. M. Smith J. Am. Chem. Soc. 1982, 104, 4337-4344. B. Stuart Biological Applications of Infrared Spectroscopy; John Wiley & Sons, New York, 1997. J. L. Arrondo, A. Muga, J. Castresana and F. M. Goni Prog. Biophys. Mol. Biol. 1993, 59, 23-56. 193 [53.] [54.] [55.] [56.] [57.] [58.] [59.] [60.] [61.] [62.] [63.] [54.] [65.] J. L. R. Arrondo and F. M. Goni Prog. Biophys. Mol. Biol. 1999, 72, 367- 405. P. Hellwig, T. Soulimane, G. Buse and W. Mantele FEBS Lett. 1999, 458, 83-86. G. E. Heibel, P. Hildebrandt, B. Ludwig, P. Steinrucke, T. Soulimane and G. Buse Biochemistry 1993, 32, 10866-10877. P. Hellwig, T. Mogi, F. L. Tomson, R. B. Gennis, J. lwata, H. Miyoshi and W. Mantele Biochemistry 1999, 38, 14683-14689. A. Kannt, U. Pfitzner, M. Ruitenberg, P. Hellwig, B. Ludwig, W. Mantele, K. Fendler and H. Michel J. Biol. Chem. 1999, 274, 37974-37981. P. Hellwig, C. Ostermeier, H. Michel, B. Ludwig and W. Mantele Biochim. Biophys. Acta 1998, 1409, 107-112. T. Noguchi, T. A. Ono and Y. Inoue Biochim. Biophys. Acta 1995, 1232, 59-66. P. Hellwig, U. Pfitzner, J. Behr, B. Rost, R. P. Pesavento, W. von Donk, R. B. Gennis, H. Michel, B. Ludwig and W. Mantele Biochemistry 2002, 41, 9116-9125. R. B. Gennis Biochim. Biophys. Acta 1998, 1365, 241-248. D. A. Proshlyakov, M. A. Pressler, C. DeMaso, J. F. Leykam, D. L. DeWitt and G. T. Babcock Science 2000, 290, 1588-1591. D. A. Proshlyakov, M. A. Pressler and G. T. Babcock Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 8020-8025. M. Linben, A. Prutsch, B. Mamat and K. Genrvert Biochemistry 1999, 38, 2048-2056. Y. Maeda, T. Nakamura and I. lkeda Macromolecules 2001, 34, 8246- 8251. 194 [66.] [67.] [68.] [69.] [70.] [71.] [72.] [73.] [74.] [75.] [76.] [77.] [78.] [79.] T. N. Earnest, J. Herzfeld and K. J. Rothschild Biophys. J. 1990, 58, 1539— 1546. N. W. Downer, T. J. Bruchman and J. H. Hazzard Biophys. J. 1985, 47, A105—A105. N. W. Downer, T. J. Bruchman and J. H. Hazzard J. Biol. Chem. 1986, 261 , 3640-3647. T. Kluge, J. Olejnik, L. Smilowitz and K. J. Rothschild Biochemistry 1998, 37, 10279-10285. J. R. Fetter, J. Qian, J. Shapleigh, J. W. Thomas, A. Garcia-Horsman, E. Schmidt, J. Hosler, G. T. Babcock, R. B. Gennis and S. Ferguson-Miller Proc. Natl. Acad. Sci. U. S. A. 1995, 92, 1604-1608. M. Itabashi and K. ltoh Chem. Lett. 1979, 1331-1334. T. Miura, T. Satoh, A. Hori-i and H. J. Takeuchi J. Raman Spectrosc. 1998, 29, 41-47. T. Noguchi, Y. Inoue and X. S. Tang Biochemistry 1999, 38, 10187-10195. T. Noguchi, Y. Inoue and X. S. Tang Biochemistry 1999, 38, 399-403. T. Noguchi and Y. Inoue J. Biochem. 1995, 118, 9-12. C. Berthomieu, A. Boussac, W. Mantele, J. Breton and E. Nabedryk Biochemistry 1992, 31, 11460-11471. R. Hienerwadel and C. Berthomieu Biochemistry 1995, 34, 16288-16297. C. Berthomieu and R. Hienerwadel Biochemistry 2001, 40, 4044-4052. M. Majoube, P. Millie and G. Vergoten J. Mol. Struct. 1995, 344, 21-36. 195 [80.] [81.] [82.] [83.] [84.] [85.] [86.] [87.] [88.] [891 [90.] [91.] [92.] [93.] [94.] K. Hasegawa, T. Ono and T. Noguchi J. Phys. Chem. B 2000, 104, 4253- 4265. K. Hasegawa, T. Ono and T. Noguchi J. Phys. Chem. A 2002, 106, 3377- 3390. S. Hashimoto, K. Ono and H. Takeuchi J. Raman Spectrosc. 1998, 29, 969-975. A. Toyama, K. Ono, S. Hashimoto and H. Takeuchi J. Phys. Chem. A 2002, 106, 3403-3412. I. Ashikawa and K. ltoh Biopo/ymers 1979, 18, 1859-1876. H. Wolff and E. Wolff Spectrochim. Acta A 1971, 27, 2109-2118. C. Perchard and A. Novak J. Chem. Phys. 1968, 48, 3079-3084. A. M. Bellocq, C. Perchard, A. Novak and M. L. Josien J. Chim. Phys. 1965, 62, 1334-1343. H. Zimmermann Z. E/ektrochem. 1961, 65, 821-840. R. J. Sundberg and R. B. Martin Chem. Rev. 1974, 74, 471-517. L. Regan Annu. Rev. Biophys. Biomolec. Struct. 1993, 22, 257-281. T. Peters, Jr. and F. A. Blumenstock J. Biol. Chem. 1967, 242, 1574-1578. N. Camerman, J. K. Fawcett, T. P. A. Kruck, B. Sarkar and A. Camerman J. Am. Chem. Soc. 1978, 100, 2690-2693. J. M. Berg and Y. (3. Shi Science 1996, 271, 1081-1085. D. J. Wang, X. J. Zhao, M. Vargek and T. G. Spiro J. Am. Chem. Soc. 2000, 122, 2193-2199. 196 [95.] Thamann, P. Frank, L. J. Willis and T. M. Loehr Proc. Natl. Acad. Sci. T. J. U. S. A. 1982, 79, 6396-6400. [96.] L. Nestor, J. A. Larrabee, G. Woolery, B. Reinhammar and T. G. Spiro Biochemistry 1984, 23, 1084-1093. [97.] K. Nagai, T. Kagimoto, A. Hayashi, F. Taketa and T. Kitagawa Biochemistry 1983, 22, 1305-1311. [98.] J. Teraoka and T. Kitagawa J. Biol. Chem. 1981, 256, 3969-3977. 197 CHAPTER 7 7 Spectroelectrochemical Studies Using a Novel Optically Transparent Electrode: Boron-Doped Diamond 7.1 Introduction Diamond is something magnificent, the peerless "king of gems" that glitters, dazzles, and symbolizes purity and strength. It is the optical properties of diamond that have generated the awe surrounding this material. These properties, present in both high-quality synthetic and naturally-occurring diamond, include the highest values of reflectivity (0.17), index of refraction (2.42), and dispersion (0.057) of any mineral, giving a well-cut diamond stone a marvelous brilliance and fire [1, 2]. The exceptional scintillation properties of this hard-to-find mineral have caught the eyes of the wealthy, the powerful, and the general public for centuries. Aside from imparting beauty to a gemstone, the Optical properties of diamond have been exploited in more functional roles. Gemologists routinely evaluate the quality of a diamond gemstone based on its optical properties. Spectroscopists have coveted diamond for its superb optical properties, finding it suitable for application as optical windows for IR and UV/vis spectroscopic measurements and as optical components of high-power lasers Additionally, its electronic, thermal, structural, and acoustic properties have captured the imagination of electrochemists, physicists, and engineers. 198 Chapters 3 and 4 described applications of the electronic properties of boron- doped diamond (BDD), a synthetic diamond material that is doped with boron atoms to render it highly conductive. The focus of this chapter will be the investigation of the combined electronic and optical properties of BDD for spectroelectrochemical measurements. The use of electrically conductive diamond as an optically transparent electrode (OTE) is a new area of research [3-5]. BDD possesses attractive qualities as both an electrode and an optically transparent material, making it an obvious choice for utilization as an OTE in spectroelectrochemical measurements. Diamond OTEs offer several advantages over other materials including: (i) the possibility of transmission measurements from the near-UV into the far-IR (0.25 -100 pm), (ii) a low background current, (iii) a wide working potential window, (iv) stability in both aqueous and nonaqueous solution environments during both cathodic and anodic polarization, and (v) resistance to foufing. An added advantage of BDD OTEs relies on the ease and versatility of fabrication, as BDD thin-films have been grown on a variety of substrates (e.g., CVD white diamond, quartz and undoped Si), and also exist as freestanding disks. Each type of OTE possesses characteristics that can be exploited for a specific application. This work is a product of our long-term goal to develop BDD as an optically transparent electrode for electroanalytical applications and to study biological electron transfer mechanisms. At present, our group has investigated the detection capabilities of BDD OTEs in the UV/vis 199 for analytes such as chlorpromazine [5]. We are also working to develop BDD OTEs for use in the far-IR detection of single metal-ligand modes (e.g. Cu-N) in complex metalloenzymes, such as cytochrome c oxidase. Such spectroelectrochemical studies of biomolecules will provide experimental evidence for the various models that exist to describe their mechanism of action. An important part of the material development involves manipulation and control of the optical properties through adjustments in the deposition conditions. The first-generation diamond OTEs developed by our multi-investigator team demonstrated 40 - 80% transmissivity from the UV into the far infrared. Optimization of the diamond OTEs is achieved through transmission spectroelectrochemical measurements of both aqueous and nonaqueous redox systems in the UV/vis, and IR regions of the electromagnetic spectrum. There are many factors that influence the transmissivity of BDD. Many of these can be controlled, to some extent, by manipulation of the diamond growth conditions. Disruption of the lattice symmetry by spz-bonded nondiamond carbon or introduction of chemical impurities (intentional or not), such as boron and nitrogen, will decrease transmission, as each imperfection gives rise to its own characteristic effect [1, 6-9]. This creates an interesting problem, as boron doping is required to make the diamond films conductive. Scattering losses contribute to the overall decreased light throughput via structural defects and large grain size. Losses due to scattering become very important when the defect or grain size approaches that of the incident wavelength [7]. The 200 thickness of the film affects the transmission, as transmission decreases with increasing film thickness. Work remains to completely understand the relationship between the growth conditions and optical throughput. We continue to investigate the effect of the boron doping level on electrode conductivity and transmission. Highly doped films (B concentration~1020 cm'3) have been used for measuring P apparent rate constants for ferrocene [10] and for cyclic voltammetric measurements of a variety of aqueous systems [11, 12]. Such films are highly conductive but not transparent. To achieve maximum light throughput, less L conductive films, with lower doping levels (B concentration ~ 1018 cm'3), are used for spectroelectrochemical measurements. Table 7-1 demonstrates the general effect of several controllable growth parameters on the electrical and optical properties of BDD OTEs. Table 7—1 Effect of Selected Growth Parameters on the Optical and Electronic Properties of Boron-Doped Diamond Optically Transparent Electrodes . . Conductivity Growth Parameter OTE Property Transmrssron -1 -1 (0 cm ) micro vs. nano BDD T surface roughness i no change 1 growth time 1 film thickness I no change 1‘ conc B2H6 gas 1 B doping level i i micro, microcrystalline; nano, nanocrystalline We previously reported on the development of freestanding BDD for transmission UV/vis spectroelectrochemical measurements. The electrode and 201 optical responses toward ferri/ferrocyanide and methyl viologen were used to demonstrate the potential usefulness of this type of OTE [4]. We expand upon that work, presently, by reporting the first measurements of the spectroelectrochemical responsiveness of optically transparent diamond electrodes toward a nonaqueous redox analyte, ferrocene. We also exploit the IR optical transparency of BDD. Martin and Morrison recently demonstrated the application of a BDD OTE for attenuated total reflection (ATR) IR measurements [3]. Our work, however, demonstrates the utility of BDD for IR transmission spectroelectrochemistry. Fabrication of the BDD OTEs was previously described in Sections 2.1.3 (freestanding BDD disk) and 2.1.4 (BDD film on undoped Si). The novel optically transparent thin-layer electrochemical (OTTLE) cell designs incorporating these BDD OTEs for use in UV/vis and IR were discussed in Sections 2.4.2 (UV/vis OTTLE cell), 2.4.3 (IR OTTLE cell -— Model I), and 2.4.4 (IR OTTLE cell - Model II). The electrochemical characterization of these thin- layer cells, as well as their utility in spectroelectrochemical measurements of ferrocene, ferrocyanide, and cytochrome c are reported in this chapter. The main advantage of using a thin-layer cell design for spectroelectrochemical measurements lies in the very short times required (typically 20 - 120 s) for complete electrolysis. In order to exhaustively electrolyze the bulk solution, as is desired in spectroelectrochemistry, a large area (A) to volume (V) ratio is required. One method for achieving a large AN ratio without convective mass transfer involves confining a very small volume (i.e., a few uL) to a thin layer (2 - 250 pm) at the electrode surface. 202 According to the random-walk model of diffusion, the diffusion layer thickness, x, is defined by x: 20L in which D is the diffusion coefficient (cmzls) and t is the experimental time (s) [13]. Mass transfer within the cell can be neglected when the solution thickness, I, is much less than the diffusion layer thickness, or I< g 0.0 0.2 0.4 -05 -0.8 l r V V I V 11*1 red L Ep 1 l -0.2 -O.1 0.0 0.1 0.2 Potential (V) Figure 7-1 shows a reversible cyclic voltammogram for the thin-layer case, in which diffusion is neglected, and the current is instead limited by the exhaustive electrolysis of the thin-layer solution. The experiment begins at the initial potential (E1) of —0.20 V and is scanned toward the switching potential (EA) 204 of 0.20 V, at which point the direction of the scan is reversed. The current remains minimal until ~-0.10 V, beyond which, the current rises rapidly with increasingly positive values of Eapp. as the oxidized form is generated. The oxidative current reaches a maximum (igx) at the oxidation peak potential (ng ). Beyond the E8", the current drops rapidly to the baseline until the scan reaches the EA. At this point, the direction of the scan is reversed and the shape of the reverse scan mirrors that of the fonlvard scan with the reduction peak current ( ‘fed lp ) having equal magnitude but opposite sign to that of the 10", and the reduction peak potential (Eged) at the same value as that observed for the oxidative process. The shape of the i-E curve presented in Figure 7-1 infers exhaustive transformation of the reduced form to the oxidized form upon the forward scan and the complete regeneration of the reduced form during the reverse scan [13- 16]. Theoretically, if the rate of electron transfer is faster than the scan rate, then the peaks should be symmetric about EO', and due to the exponential dependence of the current with the potential, Gaussian in shape. It is important to note that the peak currents, and therefore, peak potentials occur at the formal reduction potential, E". The peak current is given by _r12F2vI»'('; P 4RT ’ i 205 in which the scan rate, v (V/s), the cell volume, V (cm3), and the bulk Ir concentration, (2,, (moI/cm3), yields current in amperes [13]. From this relationship, it is seen that the current varies linearly with the scan rate and the concentration. The integrated area under the cyclic voltammetric peak yields a value of the faradaic charge, Q, passed during the anodic and cathodic processes. For exhaustive electrolysis, the charge can be calculated from the number of moles of the oxidized form, NO, or from the cell volume and concentration according to Q = nFNU = nFl'CZ. Examination of the above equation reveals that the total charge passed is independent of the scan rate [13]. The electrochemical performance of thin-layer spectroelectrochemical cells is evaluated based on the above criteria. Knowledge of the charge also enables one to compare the computed cell volume with that expected from the cell dimensions or injected volume. Thin-layer spectroelectrochemical measurements are effective, because, in a relatively short time, one can obtain electrochemical and spectral information from a single experiment. 7.2 Results - UVIvis Spectroelectrochemistry 7.2.1 Optical Properties of BDD OTEs - UVIvis Optically pure white diamond (i.e., chemically pure and low in defects) has one of the widest optical windows of any material, extending from the band gap 206 absorption edge at 225 nm well into the far-IR region [7]. Imperfections (impurities and defects) in diamond, like those that exist in boron-doped polycrystalline films, give rise to multiple absorption and scattering phenomena that reduce the light throughput [1, 6, 7]. Such defects can be structural or chemical in nature. Structural defects include grain boundaries, stacking faults, dislocations, twin boundaries and point defects, such as vacancies or divacancies. Nitrogen and boron are common chemical impurities that may be intentionally introduced, or naturally occurring. These impurities give rise to point defects or defect complexes [1, 6, 7, 17-19]. Figure 7-2 shows transmission spectra in the UV/visible region of the electromagnetic spectrum for two BDD OTEs with different levels of boron doping, compared to the spectrum for a piece of optically-pure white diamond. It is clear from Figure 7-2 that boron doping results in an absorption continuum beginning at about 500 nm and extending into the near-IR. Some of the decreased transmittance is due to increased reflectance with increasing doping level. This continuum (absorption from 0.37 -2 eV) is due to the creation of holes as electrons populate the degenerate boron acceptor levels. High concentrations of boron result in strong absorption in the red, making heavily doped films appear blue [1]. Nitrogen introduces yellow coloration due to electronic absorption in the UV and blue regions of the visible spectrum [1, 6, 17- 19]. In general, boron-doped films are more opaque than their non-doped counterparts. Increased absorption due to nitrogen impurities or non-diamond 207 carbon and increased reflectance or scattering may also contribute to the overall decrease in the transmission of BDD films. Figure 7-2 UV/vis Transmission Spectra of Freestanding Boron-Doped and CVD White Diamond Disks: Effect of Boron Doping Level Transmission spectra demonstrating the optical transparency of a CVD white (type Ila) diamond window (solid line) and a freestandirztg BDD disk OTE from 200 - 800 nm. Spectra are presented for a heavily doped (~ 10 B/cm3) disk (dashed line) and a lightly doped (~ 1018 B/cm3) disk (dashed-dotted line). % Transm'ssion ‘ Q Q . ~ .- Q in- Figure 7-3 compares the optical throughput in the UV/vis region for a traditional OTE, ITO deposited on quartz, to that of different types of BDD OTEs. Some of the overall decreased transmittance observed for the BDD OTEs, compared to the ITO film on quartz, is likely a result of the differing thicknesses. The BDD freestanding OTEs were ca. 380 pm thick, and the BDD film on quartz was ca. 1 pm, compared to the 20 nm nominal thickness of the ITO film. 208 Although the BDD films offer a small (~ 75 nm) gain in the accessible range of the UV/vis region, the most distinct advantages over ITO lie in the extended transparency of BDD into the IR region [3, 20], as well as in the superior electrochemical properties and mechanical stability of diamond [5]. Figure 7-3 UV/vis Transmission Spectra of Optically Transparent Electrodes: Boron-Doped Diamond compared to Indium-Doped Tin Oxide A comparison of the UV/vis transmission spectrum of (A) an ITO thin-film on quartz to the spectra of several BDD OTES: (B) a BDD thin-film on quartz, (C) a polished BDD thin-film on CVD white diamond, and (D) a freestanding BDD disk. 80 . .- A .8 E B L '— . o\° W'RTr-s C D l 2004 300R400 L500; 500‘7001900 Wavelength(nm) The spectrum of the BDD film deposited on quartz (Figure 7-3B) does not show the sharp cut-off at 226 nm possessed by the other BDD films (Figure 7-3C and D), this is likely due to absorption by aggregates of unintentionally-added nitrogen atoms, which decrease the transmittance below 300 nm [5, 21]. The 209 adventitious impurity comes mainly from atmospheric leaks into the reactor during deposition. The freestanding disks were grown by our collaborator, Prof. James E. Butler at the Naval Research Laboratory (NRL), under conditions that are pristinely shielded from commonly encountered nitrogen impurities arising from source gas contamination or atmospheric leaks into the CVD reactor. An ideal BDD OTE for UV/vis spectroelectrochemical measurements would consist of an ultrathin film with a nominal grain size less than 500 nm, in an effort to minimize light absorption and scattering losses, respectively. Additionally, such films would possess a minimal electrical resistivity (< 0.05 0 cm) and high optical throughput (50 - 80%). The nanocrystalline BDD film deposited on quartz approaches these figures of merit. These films can be reproducibly prepared to possess low electrical resistivity (< 0.03 0 cm) and high optical throughput (40 - 50% from 300 to 800 nm) [5]. However, at the time the work reported in this chapter was performed, only the freestanding BDD disks were available for use in the UV/vis region. The BDD freestanding disk was chosen over the BDD on CVD diamond disk due to higher conductivity, better electrode performance characteristics, and the fact that its usefulness had been demonstrated for aqueous spectroelectrochemical measurements [4]. It is important to note that the properties of these new BDD OTEs do not represent the limiting case; but, instead, a starting point from which an understanding of the required deposition conditions for the optimal balance of electrical conductivity and optical transparency can be established. 210 7.2.2 Electrochemical Characterization of the UVIvis OTTLE Cell The cuvette-based OTTLE cell used for spectroelectrochemical experiments in the UVIvis region was previously described in Chapter 2.4.2 and shown in Figure 2-3. An important step in OTTLE cell design is the characterization of the thin-layer electrochemical properties of the cell. As discussed above, there are a number of criteria that a cell must meet in order to demonstrate thin-layer behavior. Only when the cell behaves as a thin-layer case can one be assured that complete electrolysis is occurring. A series of CV measurements were performed to test the linearity of the peak current variance, and the relationship of the faradaic charge with the scan rate. Figure 7-4 Ferrocene Cyclic Voltammetric Response Measured in the UV/vis Optically Transparent Thin-Layer Electrochemical Cell (A) Cyclic voltammetric i-E curve for 0.1 mM ferrocene in acetonitrile containing 0.2 M TBAC/O4 at 0. 002 V/s. (B) Reduction (squares) and oxidation (circles) peak currents versus the scan rate plot for 0.1 mM ferrocene in acetonitrile containing 0.2 M TBA ClO4. Data are shown for scan rates from 0. 002 to 0.010 V/s. Current (uA) A 4”0.2 ‘ 03 ‘ 074 05 ‘ 06 0mo ‘ 0.005 0310 Paertial (V vs. Ag wire QRE) Scan rate (V/s) 211 Figure 7-4A shows a cyclic voltammetric i-E curve for the diamond OTE exposed to 0.1 mM ferrocene in 0.2 M TBACIO4/CH3CN. The measurement was made in the thin-layer spectroelectrochemical cell. The potential sweep rate was 0.002 V/s, in an effort to minimize ohmic distortion. The E0 is 0.394 V, as calculated from the average of the oxidation and reduction peak potentials. AEp is 0.033 V and Qox/Qred is 1. The oxidation and reduction current peaks are nearly symmetric with no evidence for any diffusion effects, as expected for a thin-layer cell. The ferrocene reduction (R2 = 0.998) and oxidation (R2 = 0.999) peak currents increase linearly with the scan rate in the range from 0.002 to 0.020 V/s, as shown in Figure 7-4B. The charge (Q())( = 25.7 (:1: 2.0) uC, Qred = 24.6 (i' 1.7) uC) passed at the electrode due to the ferrocene faradaic process is independent of the scan rate over the same range. From the average charge, the value calculated for the cell volume is 2.52 uL. The expected cell volume was ca. 3 uL. The voltammogram is slightly distorted with an upward shift to the right. This is likely due to the changes in the background capacitive current due to slow solution leakage around the electrode/mica spacer seal. For ideal thin-layer behavior, the theoretical peak splitting is zero when the electron transfer kinetics are rapid compared to the potential sweep rate. Some peak splitting is observed, due to ohmic cell resistance. 7.2.3 Ferrocene Figure 7-5 shows a series of ferrocene absorbance spectra between 225 and 350 nm collected at increasingly positive potentials. UVIvis absorbance 212 spectra for 1 mM ferrocene in 0.5 M TBACIO4/CH30N were obtained with the freestanding diamond OTE in the thin-layer cell. The potential was stepped intermittently from 0.20 to 0.60 V and spectra were collected after a one-minute equilibration period at each potential. Each spectrum was corrected by subtracting the reference spectrum obtained at 0.10 V. Figure 7-5 UV/vis Spectroelectrochemical Absorbance Spectra of Ferrocene Oxidation at a Freestanding Boron-Doped Diamond Optically Transparent Electrode A series of U V/vis absorbance spectra for the generation of ferricenium via electrooxidation of ferrocene. The potential was stepped intermittent/y from 0. 20 V to 0. 60 V. In all cases, the analyte solution consisted of 1 mM ferrocene in acetonitrile containing 0.5 M TBA CIO4 as the supporting electrolyte. Each spectrum was corrected for background absorption by subtracting a reference spectrum acquired at 0.10 V. The series includes spectra acquired after a 1 min equilibration period at (A) 0. 20 V, (B) 0.34 V, (C) 0.38 V, (D) 0.42 V, (E) 0.46 V, and (F) 0.60 V. The optical pathlength was 80 pm. 0.14 - » 252 nm 0.12 - . F 0.10 e 08 » E 0.08 - . D OTB e 0.04 - i C 0.02 f B 0.00 - ’ 1 . 1A A 1 t r . 1 g 225 250 275 SC!) 325 350 Wavelength (nm) 213 As the potential is made more positive, the oxidation of ferrocene to ferricenium occurs and the absorbance at 252 and 285 nm increases. These two absorbance maxima are due to ligand to metal charge transfer processes of the ferricenium ion [22]. The absorbance was constant beyond 0.55 V, indicating complete electrolysis of ferrocene in the thin-layer cell. The spectral features were completely reversible with repeated anodic and cathodic steps. In addition to obtaining the absorption spectrum of the oxidized ferricenium species, one can extract the values of n and E°' from such data via the slope and x-intercept of a Nernst plot. As discussed in the introduction to this chapter, the Nernst equation defines the relationship between the applied potential and the ratio of the oxidized-to-reduced concentrations in the thin solution layer. Application of Beer’s Law, A zabC , relates concentration to absorbance, in which, e is the molar absorptivity(M'1cm'1), b is the optical pathlength (cm), and C is the analyte concentration (M). Therefore, one can construct a so-called Nernst plot from a series of absorbance spectra acquired at different potentials, as for each value of Eappl, the corresponding concentration ratio can be calculated from the observed absorbance. Assuming that that the pathlength and molar absorptivity remain constant allows one to construct a Nernst plot directly from the measured absorbance changes. 214 Figure 7-6 Nernst Plot for the UV/vis Spectroelectrochemistry of Ferrocene at a Freestanding Boron-Doped Diamond Optically Transparent Electrode A Nernstian plot overlaid on Figure 7-4A demonstrating the linear relationship between the log of the ratio of the oxidized and reduced product absorption and potential. Data were acquired from the potential step experiments presented in Figure 7-5. In all cases the analyte solution consisted of 1 mM ferrocene in acetonitrile containing 0.5 M TBA C/O4 as the supporting electrolyte. -0. - . - 1 - 1 . 1 -1.5 13.2 0.3 , 0.4 0,5 0.6 Potential (V vs. Ag Wire QRE) Figure 7-6 shows a linear relationship in the Nernstian plot, which indicates the system reaches equilibrium and the absorbance increase correlates with a one-electron reaction. The slope of the linear fit (R2 = 0.999) yields a value of 1.1 for n. The x-intercept results in a E°' value of 0.397 V, which is in excellent agreement with that calculated from the CV (0.394 V), as emphasized by the double plot in Figure 7-6. For the data presented in Figure 7-4, the ratio of Aon/AAred was obtained using the following equation AA A AA 0." A A red max _ 215 where A represents the absorbance at 252 nm at a given potential, and Amax is the maximum absorbance at 252 nm observed when Eapp. >> EO' Figure 7-7 Absorbance-Potential (Voltabsorptometry) Profiles of Ferrocene UV/vis Spectroelectrochemistry at a Freestanding Boron-Doped Diamond Optically Transparent Electrode Absorbance-potential profiles for electrogenerated ferricenium absorbance peaks at 252 and 285 nm. In both cases, the data were acquired simultaneously with a cyclic voltammogram at 0. 002 V/s. The measurement was made in the cuvette OTTLE cell for a solution of 1 mM ferrocene in acetonitrile containing 0.5 M TBACIO4 at the freestanding BDD OTE. The absorbance presented is an absolute value, uncorrected for the background absorbance. The optical pathlength was 80 pm. 1.95 Absorbance 1.75 ~ 170 " 285 nm 160 4 4 1 1 1 1 0.4 0.5 0.1 0.2 A 0.3 0.6 1 0.7 Potential (v vs. Ag wire QRE) Figure 7-7 shows absorbance-potential (voltabsorptometry) profiles measured during one complete cyclic voltammetric cycle through the oxidation and reduction peaks. The scan was initiated at 0.10 V at a rate of 0.002 V/s and 216 reversed at 0.70 V. The absorbance profiles at 252 and 285 nm were collected as a function of time. For presentation purposes, conversion from time to potential was accomplished via knowledge of the experimental time over which the cyclic voltammogram was performed and the scan rate. The shape of the Abs-E profiles are diagnostic of exhaustive electrolysis of the thin-layer solution [14-16]. Upon initiation of the scan, the absorbance remains constant until ~ 0.30 V, at which point the absorbance begins to increase as the potential becomes more positive, due to the electrogeneration of the ferricenium ion. Beyond 0.55 V, the absorbance again becomes constant despite the increasingly positive potentials. This constancy demonstrates both the completion of the oxidation of ferrocene to the ferricenium ion, as well as the retention of the redox species within the restricted volume of the thin-layer. The reverse sweep yields a mirror image of the forward scan. The absorbance should theoretically return to its original value. The magnitude of the absorbance increase (AA252 = 0.094, AA235 = 0.062) of both peaks compares well with that measured during the potential step measurements shown in Figure 7-5 (AA252 = 0.110, AA285 = 0.072). At both wavelengths, the absorbance increases and decreases with the same rate of change, and the magnitudes of the initial and final absorbance are nearly the same. The relative absorbance at 252 to that 285 nm is 1.54, in good agreement with that expected from the ratio of the extinction coefficients (1.49). From the 217 Abs-E data, the reduction potential for the ferrocene/ferricenium was found to be ~ 0.40 V. Figure 7-8 Concentration Dependence for the Electrogenerated Ferricenium UV/vis Absorbance Peaks (A) U V/vis spectra obtained for varying concentrations of ferrocene in acetonitrile containing 0.2 M TBA CIO4. Spectra are presented for 2 mM (dashed-dotted line), 3 mM (dotted line), and 5 mM (solid line) ferrocene. The measurements were performed in the UV/vis OTTLE cuvette cell using a freestanding BDD OTE with a 150 pm spacer. For each measurement, the potential was stepped from 0. 20 V to 0. 60 V and a spectrum was acquired after a 1 min equilibration period. Each spectrum was corrected for the background absorbance by subtracting a reference spectrum obtained at 0.10 V. (B) Linear dependence of the absorbance at 252 (squares) and 285 (circles) nm on the concentration for the data presented in (A). 1.2 - B 0.9 ~ 0.6 L O I Abs 252 nm 0-3 ' o Ab5285 nm 1.0 ‘ 2.0 3.0 4.0 5.0 A 6.0 Concentration(rrM) Absorption spectra for three different ferrocene concentrations are shown in Figure 7-8A. The spectra were collected using a 150 pm spacer at 5 mM, 3 mM, and 2 mM ferrocene in acetonitrile containing 0.2 M TBACIO4 supporting electrolyte. The potential was stepped from 0.20 to 0.60 V and spectra were collected after a 1 min equilibration period at each potential. Each spectrum was corrected by subtracting the reference spectrum obtained at 0.10 V. Figure 7-8B 218 demonstrates that the linear dependence of absorbance on concentration of ferrocene for both peaks. The linear correlation coefficients are 0.999 and 0.990 for the peaks at 252 and 285 nm, respectively. From the slope of the absorbance versus concentration plots, the extinction coefficients are calculated 1 as 15,110 111'1 cm“1 at 252 nm and 10,845 ltli‘1 cm‘ at 285 nm, which is in agreement with independently reported values [22]. In summary, the results from the diagnostic tests discussed above establish that the OTTLE cell design permits rapid, accurate, and reproducible control of the reduction state of the molecules in the thin layer, which are the necessary conditions for obtaining high-quality UV/vis electrochemical difference spectra. It is also important to note that despite the use of nonpolar solvents, voltammetry was not plagued by large values of inherent cell resistance, further confirming the choice of cell geometry used in these designs. 7.2.4 Cytochrome c A series of reduced minus oxidized UVIvis difference spectra of cytochrome c are shown in Figure 7-9. The insert shows the absolute UVIvis absorption spectra of the fully oxidized (black line) and fully reduced (grey line) states for comparison. When fully oxidized, the Soret maximum occurs at 410 nm, and when fully reduced, the Soret maximum shifts to 414 nm and a prominent d-band at 550 nm appears. The largest amplitude difference signal is due to the spectrum collected at —0.20 V. The other spectra represent intermediate states of the protein collected at 0.05 V intervals from 0.30 to -0.20 V. The arrows in the figure depict that positive peaks, corresponding to the 219 reduced state, increase in amplitude with the application of increasingly negative potentials. Figure 7-9 UV/vis Spectroelectrochemical Absorbance Spectra of Cytochrome c at a Freestanding Boron-Doped Diamond Optically Transparent Electrode U V/vis reduced minus oxidized (at 0. 40 V) difference absorbance spectra for horse heart cytochrome c collected at 0.050 V intervals from 0. 30 to —0. 20 V. The inset depicts the absolute reduced (grey line) and oxidized (black line) spectra, which were recorded after one-minute equilibration at -0. 30 V and 0.40 V, respectively. The protein solution contained 3.5 mM cytochrome c in 100 mM phosphate buffer at pH 7.4. Measurements were performed in the UV/vis OTTLE cell using a freestanding BDD OTE. The optical pathlength was 40 pm. 0.10 - 414 _o, 300 350‘4m‘450 500 550‘600 Wavelength(nm) The UVIvis spectra of heme-containing proteins such as cytochrome 0 contain two major bands, the Soret and o-bands. The most intense of these two is the Soret band, which occurs between 400 - 450 nm, depending on the nature of the heme iron and the state of the enzyme. Both bands are attributed to the 220 n—m“ transitions of the heme, and are, therefore, a very sensitive indicator of the heme iron oxidation state [23]. The spectra presented in Figure 7-9 (from 400 - 600 nm) are the same as reported earlier [24] and are indicative that the direct electrochemical titration of this enzyme is achievable. It is important to note that the response is stable in time and reproducible with repeated cycling. The stable isobestic points further indicate that no protein denaturation occurred. One additional feature that should be pointed out from Figure 7-9 is the accessible wavelength range possible with a BDD OTE. Conventional methods for spectroelectrochemical analysis of cytochrome 0, including indirect methods relying on electrochemical titrant compounds [25] have not demonstrated access below 400 nm due to interferent absorption. Figure 7-9 shows that the region between 200 - 400 nm contains two broad absorption bands, the N and L bands, which, like the Soret and d-bands, are due to rr—->rr* transitions of heme Fe [27]. Additionally, many characteristic amino acid absorption phenomena occur below 400 nm, such as tyrosine absorption at 280 nm [25]. Although the large Soret bands report on the oxidation state of the heme, the absorbance of the d-band is routinely used to calculate cytochrome 0 concentration and to monitor heme reduction [25]. This is due to the fact that absorption of 550 nm is characteristic to reduced cytochrome 0, therefore, it is not prone to overlapping bands, as is the case in the Soret region where the oxidized and reduced forms give rise to strong, broad absorption bands at nearly equal energies. The absorbance data at 550 nm obtained from the direct redox titration of cytochrome c is shown in Figure 7-10. The solid line through the data 221 points in Figure 7-10 shows a near perfect fit to the Nernst equation. The formal reduction potential of cytochrome c from this plot is 0.072 V, which is In agreement (3: 0.01 V) with the accepted value of 0.06 V (vs. Ag/AgCI) [28, 29]. The peak amplitudes remained constant negative of —0.10 V and positive of 0.30 V, as indicated by the plateaus in Figure 7-10. Figure 7-10 Direct Electrochemical Redox Titration of Cytochrome c at a Freestanding Boron-Doped Diamond Optically Transparent Electrode Plot of the amplitude of the a band at 550 nm versus the applied potential. Spectra acquired at discrete steps from 0.40 to —0. 30 V after a 1 min equilibration in the cuvette OTTLE cell. The line represents the sigmoidal fit of the Nernst equation (R2 = 0. 998). The protein solution contained 0.5 mM cytochrome c in 100 mM phosphate buffer at pH 7. 4. The optical pathlength was 40 pm. 0. 080 I [I Ab5550nm] 0.025 — 0.020 I AAbsorbanoe .0 E? 01 I 0.010 I 0. (1)5 e 0.1 l 0.0 A 0.1 L 0.2 A 0.3 A 0.4 Potential (V vs. Ag/AgCI) AL 0.000 ‘ ' ‘ —0.3 —0.2 222 7.3 Results - IR Spectroelectrochemistry 7.3.1 Optical Properties of BDD OTEs — IR Figure 7-11 The Effect of Boron Doping Level on the IR Transmission of Diamond Optically Transparent Electrodes IR transmission spectra from 3000 - 700 crnA1 of (A) a BDD freestanding disk with low dopant concentration (~ 1018 B/cm3), (B) a highly boron doped (~ 1020 B/cm3) freestanding BDD disk, and (C) a CVD white (type Ila) diamond disk. Distinct regions of the diamond IR spectrum are represented by the numbers across the top axis: ( 1) one-phonon vibronic transitions, (2) two-phonon electronic transitions, and (3 CH) three-phonon electronic and CH” stretching vibronic transitions. All spectra are the average of 32 scans collected at 4 cm1 resolution. 3 40H-.. 2 “‘1 k 100» l o 80.. .6 . r C 3’) EGO- 8 ‘8 1.. 5°40- 20- l L 3000 2500 2000 A 1500 A 1000 Frequenqr(cm“) Figure 7-11 demonstrates the effect of boron doping on the transmission properties of diamond in the IR region. The IR absorption spectrum of diamond is unique, in that, with the exception of the one-phonon region, the transitions 223 involved are electronic rather than vibronic in nature [1]. The CVD white diamond spectrum (Figure 7-11C) is characteristic for type Ila diamond, which is relatively free of structural defects and chemical impurities [30]. The IR absorption of optically pure diamond is comprised of intrinsic multi-phonon (i.e., lattice vibration) modes and CHn symmetric and asymmetric stretching vibrations. Studies of diamond-like carbon films attribute modes at < 3000 cm'1 to spa-bonded CH groups and those at > 2950 cm'1 to spz-bonded CH [1, 31, 32]. The absorbance bands in the 2700 - 1400 cm'1 region are due to the tvvo-phonon absorption [1 , 2, 7, 30]. These modes are intrinsic to all diamond [8]. Due to the high degree of lattice symmetry in type Ila diamond, the one-phonon mode is prohibited; therefore, it is not observed in Figure 7-11C. When boron is introduced into the lattice, creating a conductive OTE material (type llb diamond), the transmittance decreases, as shown in Figures 7-11A and B. Comparison of these figures reveals that the IR absorption is proportional to the impurity content of the diamond film. Introduction of boron breaks the lattice symmetry, resulting in increased absorption due to impurities and one-phonon modes [1]. Specifically, boron gives rise to absorption bands at 2460 and 2790 cm’1 due to the electronic transitions from the ground to the first and second excited states of the dopant atoms, respectively [8, 9, 33]. Figure 7-11B shows that in highly doped films a broad continuum develops beyond 2000 cm'1 as a result of the interaction of the boron centers, resulting in 224 boron acceptor level degeneracy. The mode at 1290 cm'1 is the one-phonon vibronic absorption induced by the presence of boron acceptor centers [8, 9, 33]. The spectra in Figure 7-11 have not been normalized for B doping level or optical pathlength. Therefore, to make comparisons of the relative transmittance is difficult. Although the optical pathlengths for BDD freestanding disks are larger than the CVD white diamond disk, they possess increased transmittance below ~1800 cm‘1. This suggests that processes that decrease the transmission properties of the material, are less influential in the BDD freestanding films. It is likely that the refractive indices of the BDD films are lower than that of the CVD white diamond, as the transmissivity in this region is limited by reflectance losses. Despite the marked decrease in transmittance observed for the region above 2000 cm'1 due to boron incorporation, it is important to note that the BDD films retain high levels of throughput in the region below 1000 cm'1. Ideally, a BDD OTE for IR spectroelectrochemical measurements would consist of an ultrathin film deposited on a highly transparent substrate to aid in the mechanical stability. Such OTEs would also possess a minimal electrical resistivity (< 0.05 0 cm) and high optical throughput (50 - 80%) over the IR range of the spectrum. It is our opinion that a BDD thin-film on CVD white diamond or a BDD freestanding disk of minimal thickness would represent the best-case scenario for BDD OTEs. However, the unavailability and the high cost associated with the fabrication of such electrodes resulted in the decision to use BDD thin-films deposited on Si for the studies reported herein. Figure 7-12 shows the IR absorption spectra for three different types of BDD OTEs: (A) a 225 freestanding BDD disk, (B) a BDD thin-film on Si, and (C) a BDD thin-film on CVD white diamond. It is important to note that these spectra have not been normalized for boron doping level or thickness. Figure 7-12 IR Transmission Spectra of Different Types of Boron-Doped Diamond Optically Transparent Electrodes IR absorption spectra for three different types of BDD OTEs: (A) a freestanding BDD disk, (B) a BDD thin film on Si (4 h growth), (C) a BDD thin-film on CVD white diamond, ' and (D) a BDD thin film on Si (10 h growth). All spectra are the average of 32 scans collected at 4 cmA1 resolution. % Transm'ssion _. 8 8 8 8 B Figure 7-12A shows that below 2300 cm"1 the freestanding disk transmits greater than 50% of the IR radiation. This film also possesses a sufficient level of conductivity for transferring electrons during potential step measurements. However, it is clear from Figure 7-12A that this film is of a critical thickness that 226 gives rise to interference fringes below 2300 cm'A, rendering it unsuitable for our purposes. The IR spectrum of the BDD thin-film on CVD white diamond (Figure 7-12C) shows very low throughput over the entire range; this is likely due to the high boron content of this film. As mentioned earlier, these freestanding films were fabricated by our collaborators at the NRL; therefore, it was not ideal nor time-efficient to try to understand the relationship between growth conditions and optimal OTE properties via these films. Instead, we chose to focus our preliminary studies on OTEs that could be fabricated in-house. Figures 7-128 and 7-12D show the IR absorption for such BDD OTEs. The OTE represented in Figure 7-12B possessed adequate conductivity and ample transparency below -1 . . . . 2300 cm for use in spectroelectrochemical measurements. Therefore, it is important to note that the properties of these new BDD OTEs do not represent the limiting case, but the current status in our quest to understand the necessary deposition conditions for obtaining optimal OTE properties. Investigation of the BDD thin-film OTE spectra, presented in Figures 7-12B and D, reveals the effect of the growth time on the throughput. The growth conditions for the films were identical except for the growth time. As discussed above, boron doping contributes modes at 1285, 2460 and 2790 cm'1. The boron features above 2400 cm'1 drastically affect the transmission properties of the boron-doped films. In Figures 7-128 and D, these features absorb intensely, leading to a reduction in throughput to < 30% in the 4h film (Figure 7-1ZB) and < 1% in the 10 h film (Figure 7-1ZD). The resistivities were 0.4 and 0.01 0 cm for the 4 h and 10 h films, respectively. Thus, the 10 h film 227 was highly conductive, although, not viable for use as an OTE, due to low Optical throughput. The 4 h film possessed sufficient conductivity and optical throughput, rendering it useful for the potential step, spectroelectrochemical studies of ferrocene, ferrocyanide, and cytochrome 0 described later. The highly conductive, 10 h film was used for the electrochemical characterization of the thin-layer spectroelectrochemical cell. 7.3.2 Ferrocene The spectroelectrochemical cell used to measure FTIR difference spectra of ferrocene was described in Section 2.4.3 and shown in Figure 2-4. In the mid-IR region, the spectrum for ferrocene has been measured using a number of thin-layer cells [34-39]. Figure 7-13 shows the IR electrochemical difference (oxidized minus reduced) spectrum for 20 mM ferrocene and 0.1 M TBAPFe in dichloromethane. In good agreement with published data, a ferrocene C-H bending mode at 1004 cm'1 shifts to 1011 cm'1 upon oxidation to ferricenium ion [34-38]. A cyclopentadienyl ring-breathing mode at 1107 cm'1 is also observed for ferrocene. The oxidized and reduced spectra were collected after a 2 min equilibration time at 0.60 V and 0.20 V, respectively. 228 Figure 7-13 IR Spectroelectrochemical Absorbance Spectra of Ferrocene/Ferricenium Measured at a Boron-Doped Diamond Optically Transparent Electrode - Mid-IR Oxidized minus reduced electrochemical FTIR difference spectrum for the electrooxidation of 20 mM ferrocene in dichloromethane containing 0.1 M TBAPFG. The spectrum was acquired in the OTTLE cell (Model I) using a BDD thin-film on Si OTE and a AgCI window. The optical pathlength was 250 pm. Spectra were collected after a 2 min equilibration at 0.20 V (reduced) and 0. 60 V (oxidized). Each spectrum is the average of 200 scans measured at 15 0C with 4 cm1 resolution. -3 AA]1.5x10 1014 Q) U C (U '9 O (D Q < <3 1004 1107 1150 A 1100 A 1050 A 1000 A 950 Frequency (cm'1) The low-frequency FTIR difference spectrum (oxidized minus reduced) for 10 mM ferrocene plus 0.1M TBACIO4 in acetonitrile is shown in Figure 7-14. The shift from 823 to 854 cm’1 upon oxidation is clearly observed. This result is in agreement with a previous report of ferrocene oxidation in this region of the spectrum [34]. The spectra were collected at 0.20 V and 0.60 V after a 2 min equilibration. This experiment was an important test for our design, as we 229 eventually intend to use the method to measure low-frequency (<1000 cmA1) metal-ligand vibrations in protein active sites and inorganic model complexes. Figure 7-14 IR Spectroelectrochemical Absorbance Spectra of Ferrocene/Ferricenium Measured at a Boron-Doped Diamond Optically Transparent Electrode — Far-IR Oxidized minus reduced electrochemical F T/R difference spectrum of 10 mM ferrocene in acetonitrile containing 0.1 M TBACIO4. The spectra were acquired in the OTTLE cell (Model I) using a BDD thin-film on Si OTE and a AgCI window. The optical pathlength was 250 pm. Spectra were collected after a 2 min equilibration at 0. 20 V (reduced) and 0. 60 V (oxidized). Each spectrum is the average of 200 scans measured at 15 0C with 4 cm1 resolution. 354 740.1 A Absorbance 823 1000 900 8CD 700 Frequency (cm'1) 7.3.3 Electrochemical Characterization of the IR OTTLE Cell The OTTLE cell that was used for the spectroelectrochemical experiments of the aqueous samples (ferrocyanide and cytochrome c) in the IR region was previously described in Chapter 2 (Section 2.4.4) and shown in Figure 2-5. A 230 series of CV measurements were performed to test the linearity of the peak current variance, and the relationship of the faradaic charge with the scan rate, in an effort to demonstrate thin-layer electrochemical behavior within the OTTLE cell. Figure 7-15 Ferrocyanide Cyclic Voltammetric Response Measured in the IR Optically Transparent Thin-Layer Electrochemical Cell (Model II) (A) Cyclic voltammetric i—E curves for 1 mM ferrocyanide in 1 M K CI at 0. 002, 0. 005, 0.007, 0.010, and 0. 020 V/s. (B) Plot of the oxidation (circles) and reduction (squares) peak currents versus the scan rate for the CVs shown in (A). . B 9 . 6 L" I red 3 3 P o ax 75 l .. 0 5 . g .3 v -6 c -9 _ -12 A i L ‘ . l + i - 1 A _1 A 1 . 1 . J . 1 0.0 0-1 0-2 Q3 0-4 0-5 0-6 02.000 0.005 0.010 0.015 0.020 Potertial (V vs. AgIAgCI) Scat Rate (Vls) Figure 7-15A shows a series of cyclic voltammetric i-E curves for the diamond OTE exposed to 1 mM ferrocyanide in 1 M KCI. The measurements were made in the thin-layer spectroelectrochemical cell with potential sweep rates from 0.002 to 0.02 V/s. As the potential is swept from 0 to 0.60 V, ferrocyanide is oxidized to the Fe(III) ferricyanide form. The reduction of ferricyanide to the Fe(ll) ferrocyanide form takes place upon the reverse scan from 0.60 to 0 V. The characteristic thin-layer shapes of the oxidation and 231 reduction CV peaks indicate complete conversion of ferro- to ferricyanide and ferri- to ferrocyanide, respectively. The Ep/2 of 0.272 V, calculated from the average of the oxidation and reduction peak potentials, remains constant over the entire scan rate range. The AEp values range from 0.011 to 0.074 V and Qox/Qred is 1. As expected for a thin-layer cell, the oxidation and reduction current peaks are nearly symmetric with no evidence for any diffusion effects. The oxidation (R2 = 0.999) and reduction (R2 = 0.996) peak currents of the ferri-lferrocyanide couple increase linearly with the scan rate in the range from 0.002 to 0.020 V/s, as shown in Figure 7-158. The charge (Q0x = 67.5 (_+_ 4.2) pC, Qred = 68.9 (i 4.0) pC) passed at the electrode due to the ferrocyanide faradaic process is independent of the scan rate over the same range. From the average charge, the value calculated for the cell volume is 0.71 uL. The injected volume was ca. 1 uL. The thickness of the solution layer was estimated to be 11 pm by dividing the cell volume by the geometric area of the electrode (0.636 cm2). For ideal thin-layer behavior, the theoretical peak splitting is zero when the electron transfer kinetics are rapid compared to the potential sweep rate. Some peak splitting is observed, due to cell resistance, electrode resistance, or slow electron transfer kinetics. 232 7.3.4 Ferrocyanide Figure 7-16 IR Spectroelectrochemical Absorbance Spectra of Form/Ferrocyanide Measured at a Boron-Doped Diamond Optically Transparent Electrode Oxidized minus reduced (solid line) and reduced minus oxidized (dashed line) electrochemical difference F TIR spectra for 10 mM ferrocyanide in 1 M KCI. The spectra were acquired in the OTTLE cell using a BDD thin-film on Si OTE and a CaF2 window. The optical pathlength was 10 um. Each spectrum is the average of 200 scans measured at 15 0C with 4 cm'1 resolution. After a 1 min equilibration, oxidized and reduced spectra were collected at 0. 50 V and 0.10 V, respectively. AAbsorbance 2400A2200‘2000A1800 Frequency (cm’1) An experiment in the mid-frequency IR range was performed to evaluate the utility of BDD OTEs in transmission IR spectroelectrochemistry. In this region, the spectra for ferrocyanide and ferricyanide are well documented using a variety of thin-layer cells [34, 36]. Figure 7-16 shows the fonivard (oxidized minus reduced, solid line) and reverse (reduced minus oxidized, dashed line) IR 233 electrochemical difference spectra measured for an aqueous solution of 10 mM ferrocyanide containing 1 M KCI. In good agreement with published data, the CN stretching mode at 2115 cm'1 shifts to 2036 cm'1 upon oxidation to ferricyanide [34, 36, 40]. The oxidized and reduced spectra were collected after a 1 min equilibration time at 0.50V and 0.10 V, respectively. Figure 7-16 also demonstrates that the response measured in the OTTLE cell is highly reversible, in that no signal is lost during the reverse process. 7.3.5 Cytochrome c As discussed above, the results for ferricyanide and ferrocene demonstrated the viability of performing electrochemical difference IR measurements using BDD thin-films on Si as OTEs. One of our ultimate goals was to incorporate BDD OTES into measurements of redox proteins; therefore, we began our investigation with a simple, small molecular weight protein, cytochrome 0. Aside from the simplicity and availability of this protein, the electrochemical difference IR spectrum had been reported using a Au grid OTE [24], enabling us to evaluate the performance of our new OTE with respect to that of a conventional OTE. 234 Figure 7-17 Comparison of the Mid-IR Spectra of Cytochrome c Measured at a Au Grid and at a Boron-Doped Diamond Optically Transparent Electrode Reduced (-0. 20 V) minus oxidized (0. 30 V) electrochemical FTIR difference spectra of cytochrome c acquired at (A) a Au grid OTE and (B) at a BDD OTE (BDD thin-film on undoped Si). A noise spectrum (C) is shown for the cell containing the BDD OTE. All spectra are the average of 800 scans collected at 10 0C with a 10 pm pathlength. The protein solution consisted of 3.5 mM cytochrome c, 0.3 M NaCI, 10 mM Tris HCI at pH 7. Spectrum (A) was acquired using two CaFg windows and the OTTLE cell shown in Figure 2-2. Spectra (B) and (C) were collected in the OTTLE cell shown in Figure 2-5 with a Can window. AAI1X103 1 A l 1 l 1 l A l n l A l L 1800 1700 1600 1500 1400 1300 1200 1100 1000 Frequency (cm '1) Reduced minus oxidized IR difference spectra of cytochrome c are shown in Figure 7-17. This figure presents a comparison between spectra obtained using a conventional Au grid OTE (Figure 7-17A) and a BDD thin-film on undoped Si OTE (Figure 7-17B). Figure 7-17C shows that the noise spectrum, produced by application of an oxidizing potential to an oxidized sample, is flat with weak features in areas of strong background absorbance. Re-oxidation of 235 the reduced sample at the BDD OTE resulted in an oxidized minus reduced spectrum that was almost the exact mirror image of the spectrum shown in Figure 7-17B. Both spectra were found to be fully reproducible over a number of cycles. Comparison of Figures 7-17A and B reveals excellent agreement between the spectra obtained at the Au grid and BDD OTEs, in that all spectral features are retained. Moss and coworkers have reported the same FTIR spectrum for cytochrome c that was observed herein, as well as its corresponding UV/vis difference spectrum obtained for a Au grid OTE between Can windows [24]. Although simultaneous measurement of the optical spectrum was not possible in our case, the agreement of the FTIR spectra with that reported by Moss and coworkers suggests that the protein is intact and not denatured at the BDD surface. The many peaks observed in the FTIR electrochemical difference spectra of proteins are representative of amino acid side chain vibrations, and peptide backbone vibrations. Chapter 6 contained a thorough discussion of the assignments of such modes observed for cytochrome c oxidase spectra. A similar discussion is omitted for the spectrum of cytochrome c because the goal, in this case, was to evaluate the performance of the new BDD OTEs based on previously reported results [24], rather than to attempt to assign modes in the spectrum. Inspection of the regions above 1700 cm'1 and below 1100 cm‘1, demonstrates a marked difference in the noise level of the two spectra. This is primarily attributed to the ~ 20% decrease in transmittance observed for the BDD 236 OTE compared to that of the Au grid OTE. The increased noise level and decreased throughput are concerning, especially when considering low- frequency FTIR measurements. It was demonstrated in Chapters 5 and 6 that measurements of aqueous samples in this region require extremely low noise. in addition, a significant decrease in conductivity was necessary in order to achieve adequate throughput in the BDD thin-film on Si OTEs. Therefore, it is important to investigate the viability of using other types of BDD OTEs, those with higher throughput and the same or better electrical conductivity. “Membrane" BDD thin- films fabricated by our collaborators at the NRL provide one attractive option. These OTEs are based on the same idea used herein, that of depositing a thin- film of BDD onto a Si substrate; however, in this case, the substrate is carefully etched away, leaving a very fragile film of minimal thickness. Yet, these films retain some degree of stability, because only the substrate area in the light path is removed. Additionally, as discussed in Chapter 5, experiments in reflection mode, may be further investigated, as Martin and Morrison have demonstrated the application of a conducting diamond film on Si as a transparent electrode for ATR FTIR spectroelectrochemical experiments [3]. 7.4 Conclusions The spectroelectrochemical responsiveness of two types of optically transparent diamond electrodes was investigated. These novel OTEs possess adequate levels of conductivity and transmissivity for performing spectroelectrochemical measurements in aqueous and nonaqueous media. A freestanding, mechanically polished, BDD disk (0.38 mm thick and 8 mm in 237 diameter) served as the OTE for UV/vis spectroelectrochemical measurements. The OTE for the IR measurements was a BDD thin film (~ 2 pm) deposited on an undoped Si substrate. Specially designed, thin-layer spectroelectrochemical cells were constructed for UVIvis and IR transmission measurements. Transmission spectroelectrochemical measurements were made using ferrocene in acetonitrile and dichloromethane. The results were in good agreement with the literature. Ferrocene was electrooxidized to the ferricenium ion and this product was spectroscopically monitored in the UV/vis (it. = 252, 285 nm) and IR (v = 854, 1014 cm'1) regions of the electromagnetic spectrum. Additionally, the transmission electrochemical difference FTlR spectrum for the electrooxidation of ferrocyanide was measured using a BDD OTE. The conversion of ferrocyanide to ferricyanide was monitored by peaks at 2036 cm'1 and 2115 cm'1 due to the reduced and oxidized forms, respectively. These results are in agreement with independently reported spectra measured using Au minigrid OTES. Lastly, a new application of BDD OTEs was explored. The UVIvis and FTIR electrochemical difference spectra of cytochrome c were measured using BDD OTEs. The results demonstrate that BDD is a useful OTE for spectroelectrochemical measurements of aqueous solutions of proteins. Excellent agreement was observed for the response at BDD with that from conventional OTEs such as ITO and Au minigrids. However, work remains to produce a BDD OTE that can be used for low-frequency FTIR spectroelectrochemical measurements of proteins. Due to decreased 238 transmission, the noise levels below 1000 cm'1 are greater for BDD OTES than for Au grid OTEs. The work described herein, to produce highly conductive and transmissive BDD OTEs, is in the early stages. Therefore, the optical and electronic properties of the BDD OTEs presented in this work do not represent the limiting case. Further attempts to increase conductivity and transmission via manipulation of the BDD growth conditions will be pursued. Additionally, alternate methods of IR spectroelectrochemistry, namely ATR, will be investigated, in efforts to measure high-quality spectra of redox-active proteins. 239 7.5 [1-1 [2-1 [3-1 [4-1 [5-1 [6-1 [7-1 [8-1 [9-1 [10.] [11.] [12.] References J. I. Pankove and C.-H. Qui ln Synthetic Diamond: Emerging CVD Science and Technology; K. E. D. Spear, J.P., Ed.; John Wiley & Sons, Inc., New York, 1994, 401-407. A. M. Zaitsev Optical Properties of Diamond: A Data Handbook; Springer- Verlag, Berlin, 2001. H. B. Martin and P. W. Morrison Electrochem. Solid-State Lett. 2001, 4, E17-E20. J. K. Zak, J. E. Butler and G. M. Swain Anal. Chem. 2001, 73, 908-914. J. M. Stotter, J. K. Zak, Z. Behler, Y. Show and G. M. Swain Anal. Chem. 2002, in press. W. Zhu In DiamondsE/ectronic Properties and Applications; L. S. K. Pan, D.R., Ed.; Kluwer Academic Publishers, Boston, 1995, 175. Y. Tzeng Diam. Film Tech. 1991, 1, 31. K. M. McNamara, B. E. Williams, K. K. Gleason and B. E. Scruggs J. Appl. Phys. 1994, 76, 2466-2472. F. Q. Zhang, E. Q. Xie, B. Yang, Y. M. Cal and G. H. Chen Mater. Lett. 1994,19,115-118. S. Haymond, G. T. Babcock and G. M. Swain Electroanal. 2002, in press. M. Witek, J. Wang, J. Stotter, M. Hupert, S. Haymond, P. Sonthalia, G. M. Swain, J. K. Zak, Q. Chen, D. M. Gruen, J. E. Butler, K. Kobashi and T. Tachibana J. Wide Bandgap Materials 2002, 8, 171-188. M. C. Granger, M. Witek, J. Xu, J. Wang, M. Hupert, A. Hanks, M. D. Koppang, J. E. Butler, G. Lucazeau, M. Mermoux, J. W. Strojek and G. M. Swain Anal. Chem. 2000, 72, 3793-3804. 240 [13.] [14.] [15.] [16.] [17] [18.] [19.] [20.] [21.] [22.] [23.] [24.] [25.] [26.] A. J. Bard and L. R. Faulkner Electrochemical Methods: Fundamentals and Applications; 2nd ed.; John Wiley & Sons, Inc., New York, 2001. J. Niu and S. Dong Electroanal. 1995, 7, 1059-1062. N. J. Simmons and M. D. Porter Anal. Chem. 1997, 69, 2866-2869. J. Zak, M. D. Porter and T. Kuwana Anal. Chem. 1983, 55, 2219-2222. A. Mainwood Phys. Rev. B 1994, 49, 7934-7940. J. A. Freitas, P. B. Klein and A. T. Collins Appl. Phys. Lett. 1994, 64, 2136-2138. A. Cox, M. E. Newton and J. M. Baker J. Phys.-Condens. Matter 1994, 6, 551-563. S. Haymond, J. K. Zak, J. E. Butler, G. T. Babcock and G. M. Swain in preparation 2002. A. T. Collins J. Phys.-Condens. Matter 2002, 14, 3743-3750. Y. S. Sohn, D. N. Hendrickson and H. B. Gray J. Am. Chem. Soc. 1971, 93, 3603-3612. S. Horie and M. Morrison J. Biol. Chem. 1963, 238, 2859-2865. D. Moss, E. Nabedryk, J. Breton and W. Mantele Eur. J. Biochem. 1990, 187, 565-572. R. M. C. Dawson, D. C. Elliott, W. H. Elliott and K. M. Jones Data for Biochemical Research; 3rd ed .; Oxford University Press, Inc., New York, 1986. E. F. Bowden, F. M. Hawkridge, J. F. Chlebowski, E. E. Bancroft, C. Thorpe and H. N. Blount J. Am. Chem. Soc. 1982, 104, 7641-7644. 241 [27.] [28.] [29.] [30.] [31.] [32.] [33.] [34.] [35.] [36.] [37.] [38.] [39.] W. A. Eaton and J. Hofrichter Methods Enzymol. 1981, 76, 175-261. F. M. Hawkridge and T. Kuwana Anal. Chem. 1973, 45, 1021-1027. J. D. Burgess and F. M. Hawkridge In Electroanalytical Methods for Biological Materials; A. Brajter-Toth and J. Q. Chambers, Eds; Marcel Dekker, New York, 2002, 109-142. W. Van Enckevort In Synthetic Diamond: Emerging CVD Science and Technology; K. E. D. Spear, J.P., Ed.; John Wiley & Sons, Inc., New York, 1994, 322-325. L. J. Bellamy The Infrared Spectra of Complex Molecules; Chapman and Hall, New York, 1980; Vol. 2. G. Socrates Infrared Characteristic Group Frequencies; Wiley, New York, 1980. J. Mort, M. A. Machonkin and K. Okumura Appl. Phys. Lett. 1991, 58, 1908-1910. Y. J. Ma, J. M. Zheng and S. M. Zhu Chin. J. Inorg. Chem. 1999, 15, 61- 67. P. A. Flowers and S.-A. Callender Anal. Chem. 1996, 68, 199-202. J. J. Niu and S. J. Dong Electrochim. Acta 1995, 40, 823-828. J. P. Bullock, D. C. Boyd and K. R. Mann Inorg. Chem. 1987, 26, 3084- 3086. M. Krejcik, M. Danek and F. Hartl J. Electroanal. Chem. 1991, 317, 179- 187. H. Visser, A. E. Curtright, J. K. McCusker and K. Sauer Anal. Chem. 2001, 73, 4374-4378. 242 [40.] T. Noguchi and Y. Inoue J. Biochem. 1995, 118, 9-12. 243