ADVANCEMENTS IN ELECTROCHEMISTRY FOR THE SENSING OF AQUEOUS COMPOUNDS AND PERSISTENT COMPOUND DESTRUCTION By Mary Ensch A DISSERTATION Michigan State University in partial fulfillment of the requirements Submitted to for the degree of Chemical Engineering – Doctor of Philosophy 2020 ABSTRACT ADVANCEMENTS IN ELECTROCHEMISTRY FOR THE SENSING OF AQUEOUS COMPOUNDS AND PERSISTENT COMPOUND DESTRUCTION By Mary Ensch A significant research question to address is if existing electrochemical methods can be used to advance the sensing of biological compounds and toxic heavy metals, and to destroy persistent toxic pollutants to benefit the planet and humankind. Using the methods of cyclic voltammetry (CV), square wave voltammetry (SWV), square wave stripping voltammetry (SWSV), and electro- chemical oxidation (EO), the sensing of isatin and manganese, and the destruction of perfluoroalkyl compounds was investigated. Each task had the main focus of advancing our understanding of the process and making significant process improvements. Isatin is an endogenous indole that has potential pharmaceutical applications based on its anti- convulsant abilities through reaction pathways in the human body. Due to its low concentrations in blood, serum, and tissues, it is typically detected using gas chromatography mass spectroscopy (GC/MS) which requires a specialized individual and lengthy measurement times. In order to combat this, researchers have begun investigating electrochemical detection methods for this com- pound. Here, the investigation of isatin detection on a 3-in-1 boron-doped diamond (BDD) sensor was studied to simplify the electrochemical system requirements and set the ground work for a potential plug and play sensor for isatin. Both sensor configurations tested showed applicability to this application with the lowest limit of detection obtained to be 0.04 µM. Not only was isatin able to be detected in a complex sample matrix in a short measurement time with this simplified technical approach, it was done so with the elimination of extensive electrode pretreatment processes that have been required for other electrode materials in the past. This study paves the way for a fully developed BDD sensor that can expand the understanding of isatin reaction pathways and advance its ability to be used in pharmaceuticals. Manganese is an important nutrient but has toxic eects in humans making it imperative to be able to detect it in human blood, urine, and serum as well as drinking water. Indium tin oxide (ITO) has been a typical electrode material for spectroelectrochemical measurements and metal detection experiments on transparent substrates. This study goes through a few process parameters for an ITO film on a non-transparent silicon substrate. It was found that heating and annealing the film created the best film structure for an electrochemical sensing application towards manganese with a limit of detection of 0.1 parts per billion (ppb). This obtained lower detection limit allows for improved accuracy in human serum measurements with no added membranes to the electrode’s surface and uses an electrode material that already has the capabilities to be mass produced with good repeatability. This makes ITO an economical sensing solution. Per- and polyfluoroalkyl substances (PFAS) are toxic to human health and are wide spread in the environment due to their numerous consumer and industrial applications. Most current remediation techniques take advantage of filtration or adsorption technologies to remove these compounds from a water source. However, since they do not destroy the compounds, these remediation techniques yield waste products. This study, utilizing electrochemical oxidation for the destruction of PFAS, looks at tackling waste products from an ion exchange process. By utilizing BDD, a laboratory study showed that PFAS completely mineralize in these complex solutions. Overall, it was found that ammonia based solutions allow for the best mineralization of the most common PFAS using electrochemical oxidation for an ion exchange regenerate solution. This study provides valuable knowledge to ion exchange manufacturers who are looking to solve their residual PFAS waste problem. All three studies have shown that existing electrochemical techniques can be used to further sensing and remediation knowledge in today’s world and provide safer, cleaner technology options for the betterment of human health and the environment. Copyright by MARY ENSCH 2020 To Mom, Dad, Hannah, and Michaelangelo. The best support system I could have asked for. v ACKNOWLEDGEMENTS First, I would like to thank Dr. Thomas Schuelke for providing financial support through the Fraunhofer USA, Inc. Center for Coatings and Diamond Technologies, which has allowed me to complete my research as well as enable me to present the results at various scientific conferences. I would also like to thank him for his ongoing support and guidance through my PhD studies. Additionally I would like to thank both Michael Becker and Dr. Cory Rusinek for providing day- to-day mentorship and support. I am also very grateful to Dr. Scott Calabrese-Barton, Dr. Greg Swain, Dr. Elias Garratt, and Dr. Richard Lunt for serving as committee members and oering valuable insights and advice. Finally, I want to thank all of the dedicated researchers at Fraunhofer USA as well as sta members, interns, and MSU graduate and undergraduate students for all of their help along the way. vi TABLE OF CONTENTS . . . . . . . x . . . . . . . . . . . . . . . . xi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LIST OF FIGURES . CHAPTER 1 MOTIVATION OF WORK . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2 1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.2 Electrochemical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.2.1 Cyclic Voltammetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 Square Wave Voltammetry . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.2 1.2.3 9 Square Wave Stripping Voltammetry . . . . . . . . . . . . . . . . . . . . . 1.2.4 Electrochemical Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . 10 1.3 Background on Electrochemical Sensing . . . . . . . . . . . . . . . . . . . . . . . 11 Structure-property Connections . . . . . . . . . . . . . . . . . . . . . . . 12 1.3.1.1 Boron-doped Diamond . . . . . . . . . . . . . . . . . . . . . . . 12 1.3.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.3.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Prior State of the Art 1.3.3 Deficiencies of Technology . . . . . . . . . . . . . . . . . . . . . . . . . . 17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Indium Tin Oxide 1.3.1 . . . . . . . . BIBLIOGRAPHY . CHAPTER 2 2.3 Results and Discussion . Introduction . 2.1 . 2.2 Experimental Section . . . . . . . . . ISATIN DETECTION USING A BORON-DOPED DIAMOND 3-IN-1 SENSING PLATFORM . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.2.1 Chemicals and Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 Instrumentation . 2.2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3-in-1 Electrode Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.2.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.3.1 Electrode Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.3.2 BDD as a Quasi-Reference Electrode . . . . . . . . . . . . . . . . . . . . . 36 2.3.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.3.4 Urine Simulant Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 2.3.5 Repeatability Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 2.3.6 Key Advancements Obtained from Study . . . . . . . . . . . . . . . . . . 47 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Isatin Detection . . . . . . . . . . . . . . . . . INDIUM TIN OXIDE FILM CHARACTERISTICS FOR CATHODIC STRIPPING VOLTAMMETRY . . . . . . . . . . . . . . . . . . . . . . . . 56 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 Introduction . 3.1 . 3.2 Experimental Section . . . . . . . 2.4 Summary . . BIBLIOGRAPHY . CHAPTER 3 vii . . . . . . Instrumentation . 3.3 Results and Discussion . 3.2.1 Chemicals and Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.2.3 Electrode Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 3.3.1 Physical and Optical Characterization: Atomic Force Microscopy . . . . . 61 3.3.2 Physical and Optical Characterization: XRD . . . . . . . . . . . . . . . . . 63 3.3.3 Physical and Optical Characterization: UV-Vis Spectroscopy . . . . . . . . 65 3.3.4 Electrical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.3.5 Electrochemical Characterization . . . . . . . . . . . . . . . . . . . . . . . 68 3.3.6 CSV of Manganese . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.3.7 Calibration and Figures of Merit . . . . . . . . . . . . . . . . . . . . . . . 72 3.3.8 Possible Metals Interference . . . . . . . . . . . . . . . . . . . . . . . . . 74 3.3.9 Key Advancements from Study . . . . . . . . . . . . . . . . . . . . . . . . 75 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.4 Summary . . BIBLIOGRAPHY . CHAPTER 4 ELECTROCHEMICAL OXIDATION FOR THE MINERALIZATION . . . . . . . . . . . . . . . . . . 4.3.3 Introduction . PFOS . PFOA . OF PER- AND POLYFLUOROALKYL SUBSTANCES . . . . . . . . . . . 84 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 4.2 Electrode Stack Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 4.3 Constant Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 4.3.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 . 4.3.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 4.3.2.1 4.3.2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 4.3.2.3 Comparison of ISE Results to LC/MS Results . . . . . . . . . . . 96 Fluorination of BDD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.4 Combined Current Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 4.4.1 Comparison of ISE and LC/MS . . . . . . . . . . . . . . . . . . . . . . . 104 4.4.2 Remediation of Simulated Ion Exchange Regenerate Solutions . . . . . . . 105 4.4.2.1 Comparison of 50 mA/cm2 to 50&5 mA/cm2 . . . . . . . . . . 106 4.4.2.2 Comparison of four dierent regenerate solutions . . . . . . . . . 108 4.4.2.3 Byproduct Formation Investigation . . . . . . . . . . . . . . . . 114 4.4.3 Energy Usage and Cost Comparison . . . . . . . . . . . . . . . . . . . . . 115 4.5 Key Advancements from Bench Top PFAS Studies . . . . . . . . . . . . . . . . . . 118 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 . . 4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 . . BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 CHAPTER 5 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 . . CHAPTER 6 FUTURE OUTLOOK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 6.1 Future Outlook . 5.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii Isatin Detection . Indium Tin Oxide Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 6.1.1 6.1.2 . . . . . . . . . . . . . . . . . . . . . . 129 6.1.3 Mineralization of Per- and Polyfluoroalkyl Substances . . . . . . . . . . . . 129 . . ix LIST OF TABLES Table 1.1: Summary table outlining the prior art for the electrochemical detection of isatin . 15 Table 1.2: Summary table outlining the prior art for the electrochemical detection of manganese in simulated buer solutions . . . . . . . . . . . . . . . . . . . . . . 16 Table 1.3: Summary table outlining the prior art for the simulated PFAS solutions . . . . . 17 Table 1.4: Summary table outlining the prior art for the complex solutions containing PFAS 17 Table 2.1: Figures of merit for isatin detection using the 3-in-1 MAC and MEA sensors compared to other reported literature studies on dierent detection methods. Standard deviations calculated at lowest concentration detected for this study . . 43 Table 3.1: Comparison of each of the figures of merit for each of the four ITO films that were investigated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 Table 3.2: Fitting parameters for the band gap determination from the Tauc plot of each film. 67 Table 3.3: Redox potentials for other metals that are potentially in environmental and biological samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 Table 4.1: Comparison of remaining concentration of PFOA obtained from the ISE anal- ysis and LC/MS analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 Table 4.2: Rate constants and their linear fits for the three tested combined current densities 104 Table 4.3: Comparison of 1 h PFOA degradation and 12 h PFOA degradation . . . . . . . . 111 Table 4.4: Comparison of 1 h PFOA degradation and 12 h PFOA degradation . . . . . . . . 113 Table 4.5: Energy requirements of each PFOA study to reach 75% defluorination or at 1 hour116 x LIST OF FIGURES Figure 1.1: (left) diagram of a beaker setup for electrochemical measurements and (right) a clamp cell setup that can also be used alternatively. Each has (A) a working electrode, (B) a reference electrode, and (C) a counter/ auxiliary electrode . . . Figure 1.2: Setup used for an electrochemical oxidation experiment including an (A) anode, (B) cathode, and (C) stir bar . . . . . . . . . . . . . . . . . . . . . . . . Figure 1.3: (left) Potential waveform for cyclic voltammetry and (right) the corresponding current response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 1.4: (left) Potential waveform and (right) typical forward, reverse, and dierence current response for square wave voltammetry . . . . . . . . . . . . . . . . . . 4 4 5 8 Figure 2.1: Molecular structure for isatin . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Figure 2.2: Optical micrograph of the MEA (left) and MAC (right) 3-in-1 sensors with exposed BDD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 Figure 2.3: Figure 2.4: (A) An SEM micrograph of the polycrystalline BDD surface that was used to fabricate the two 3-in-1 sensors with no defects or large cracks in the films structure and (B) the corresponding Raman spectrum analysis for the BDD wafer 33 (A) Cyclic voltammograms of 1 mM Fe(CN)63- and 1 mM Ru(NH3)6Cl on the MAC 3-in-1 BDD sensor oth dissolved in 1 M KCl and (B) of a 1:1 ratio of Fe(CN)63-/4- and 1 mM of Ru(NH3)6Cl , both dissolved in KCl, on the 3-in-1 BDD MEA sensor. Scan rate: 0.1 V/s. . . . . . . . . . . . . . . . . . . . 34 Figure 2.5: The BDD’s reference potential in 0.1M phosphate buered saline at various pH values measured against a Ag/AgCl reference electrode for 12 hours. . . . . 37 Figure 2.6: The BDD’s reference potential for the extended 28 hours in 0.1M phosphate buered saline at a pH of 7.4 measured against a Ag/AgCl electrode. . . . . . . 38 Figure 2.7: Current versus voltage cyclic voltammogram curves for 350 µM isatin in 0.1 M PBS at pH 7.4 on the 3-in-1 (A) MAC sensor and (B) MEA sensor. Scan rate of 0.1 V/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 . . . Figure 2.8: Current versus voltage cyclic voltammogram curves of 350 µM isatin in 0.1 M PBS at pH 7.4 (red) and after rinsing the 3-in-1 MAC electrode and rescanning the same electrode in clean electrolyte solution (black). Scan rate of 0.1 V/s. . . 40 xi Figure 2.9: Squarewave voltammograms for various concentrations of isatin on the 3-in-1 (A) MAC and (B) MEA sensor configurations in 0.1 M phosphate buered saline with pH of 7.4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Figure 2.10: Calibration curves for (A) the MAC and (B) the MEA in 0.1 M phosphate buered saline with a pH of 7.4 . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Figure 2.11: UV-vis absorbance spectra for isatin in 0.1 M of PBS pH 7.4 . . . . . . . . . . 44 Figure 2.12: Squarewave voltammograms for the 3-in-1 MAC sensor (A) and MEA sensor (B) in Surine Negative Urine Control with no additional N2 bubbling . . . . . . 45 Figure 2.13: Calibration curves for the (A) MAC and (B) MEA for isatin in Surine Negative Urine Control with no N2 bubbling . . . . . . . . . . . . . . . . . . . . . . . . 46 Figure 2.14: Repeatability data for the 3-in-1 (A) MAC and (B) MEA sensors with the discussed electrochemical pretreatment and without any addition of a pre- treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 . . . . . . Figure 3.1: Tapping mode atomic force microscopy (AFM) images for the four analyzed ITO films (A) HA, (B) HNA, (C) NHA, and (D) NHNA. . . . . . . . . . . . . 62 Figure 3.2: XRD scans for the (A) (222) peak and (B) (400) peak for each ITO film investigated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 Figure 3.3: (left) ITO crystal quality (I(222)/I(400)) ratio obtained from XRD analysis, and (right) the full width half max (FWHM) data for each film from XRD analysis pertaining to either the (222) or (400) peak . . . . . . . . . . . . . . . . . . . . 64 Figure 3.4: ITO crystal size derived using the Scherrer equation and the (222) and (400) peak of each film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 . . Figure 3.5: Transmittance spectra measured with UV/Vis for each of the four films de- posited on a transparent microscope slide . . . . . . . . . . . . . . . . . . . . . 66 Figure 3.6: Tauc graph used to determine the direct optical band gap for the four ITO films studied . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 Figure 3.7: Graph showing the current vs scan rate study of each of the ITO films used to determine their corresponding capacitance values . . . . . . . . . . . . . . . 69 Figure 3.8: (A) CV curves of 0.1 mM ferrocyanide in 1.0M KCl with a scan rate of 0.1 V/s. (B) Potential windoe study performed in acetate buer of pH 5.0. (C) CSV od 25 ppb of Mn in acetate buer of pH 5.0, deposition time of 3 minutes and a deposition potential of +1.2 V. . . . . . . . . . . . . . . . . . . . . . . . 69 xii Figure 3.9: Plots for the Ohm’s resistance of each of the four ITO films investigated . . . . 71 Figure 3.10: Mn stripping voltammograms for the full concentration range on HA (A1) and the smaller range (A2), for the full range on HNA (B1) and the smaller range (B2), and for the full range on NHA (C1) and smaller range (C2). Performed in pH 5.0 acetate buer with a deposition potential of 1.2 V and a deposition time of 3 min . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 . . Figure 3.11: Calibration curves for Mn on HA (A), HNA (B), and NHA (C) . . . . . . . . . 73 Figure 4.1: Molecular structure of (top) PFOA and (bottom) PFOS . . . . . . . . . . . . . 86 Figure 4.2: Scanning electron microscopy (SEM) image of the BDD surface . . . . . . . . 88 Figure 4.3: Raman spectroscopy of BDD electrodes . . . . . . . . . . . . . . . . . . . . . 88 Figure 4.4: Current versus scan rate plots to determine the capacitance of BDD . . . . . . . 89 Figure 4.5: Cyclic voltammogram of both ferrocyanide and ruthenium hexamine on the BDD surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Figure 4.6: PFOS Temperature, pH, and voltage trends for the constant current density studies 91 Figure 4.7: PFOS reduction overtime with each current density tested . . . . . . . . . . . . 92 Figure 4.8: PFOS 1st order rate constants . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 Figure 4.9: PFOA Temperature, pH, and voltage trends for the constant current density studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 Figure 4.10: PFOA degradation over time at each current density value . . . . . . . . . . . . 95 Figure 4.11: 1st order rate constants for PFOA based on LC/MS measurements for each current density study. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 Figure 4.12: Comparison of the PFOA concentration obtained by the F- ISE and LC/MS . . . 97 Figure 4.13: 1mM ferrocyanide in 1 M KCl measured on both the used and unused portion of the anode that was used for the constant current density PFOA degradation studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 . . . . . . . Figure 4.14: CV in 1 M sulfuric acid for both the used and unused portion of the anode . . . 100 xiii Figure 4.15: (A) CV comparison in1 mM ferrocyanide in 1 M KCl for the used anode, unused anode, and used anode post hydrogen termination and (B) CV in 1 M sulfuric acid for the used, unused, and used electrode post hydrogen termination 100 Figure 4.16: Temperature, pH, and voltage trends for the combined current density studies . . 102 Figure 4.17: PFOA reduction overtime with each combined current densities tested to- gether as well at the standard 50 mA/cm2 for comparison . . . . . . . . . . . . 102 Figure 4.18: First order rate constant fits for the first 0.25 h, k1, and the remainder of each test, k2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 Figure 4.19: Comparison of the PFOA concentration obtained by the F- ISE and LC/MS . . . 105 Figure 4.20: Temperature, pH, and voltage trends for the comparison of 50 and 50&5 in IX simulated solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 Figure 4.21: Comparison of the PFOA (left) and PFOS (right) degradation using 50 mA/cm2 and 50&5 mA/cm2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 Figure 4.22: Temperature, pH, and voltage trends for the comparison of the four IX simu- lated solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 Figure 4.23: Comparison of the PFOA (left) and PFOS (right) degradation using the four tested simulated IX solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 Figure 4.24: 1st order rate constant kinetics for PFOA in each simulated solution . . . . . . . 110 Figure 4.25: Sequential 1st order rate constant kinetics for PFOS in each simulated solution . 112 Figure 4.26: Oxygen evolution reaction performed for all four solutions . . . . . . . . . . . . 113 Figure 4.27: Calibration curve for formaldehyde detection . . . . . . . . . . . . . . . . . . . 115 Figure 4.28: Extrapolation plots used to determine the time to 75% defluorination for (A) 1 mA/cm2, (B) 5 mA/cm2, and (C) 10 mA/cm2 . . . . . . . . . . . . . . . . . 117 Figure 4.29: Bar graph comparing the conversion eciency values for the simulated IX regenerate solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 xiv CHAPTER 1 MOTIVATION OF WORK 1 1.1 Overview An important question to consider is if electrochemical techniques can be advanced to improve upon biological compound sensing, toxic compound sensing, and toxic compound destruction to benefit the planet and humankind. The ability to improve upon biological sensing is important to advance pharmaceuticals based on how compounds react in the body and to understand the levels at which they can be manipulated to better treat or prevent disease. Isatin’s first detection was in 1991 with a gas chromatographic mass spectroscopy (GC/MS) method and a few years later by high performance liquid chromatography measurements (HPLC). [1, 2] UV detection methods were being coupled with HPLC to detect isatin, but only were successful at obtaining limits of detection in the mg/L ranges. [3] These are not relevant to concentrations seen in the human body. [4, 5] Application of electrochemical detection for HPLC was utilized to gain insight on how isatin eected the levels of neurotransmitters in rats that had Parkinson’s disease, and although it provided promising results, the three electrode setup comprised of a chemically modified working electrode material that required a multiple step preparation procedure. [6] Isatin’s detection would be greatly advanced by removing the need of multiple sample pre-treatments, providing a measurement technique that is accurate and portable without extensive electrode processing, and obtaining limits of detection that allow understanding of isatin’s reactions in the human body. Toxic heavy metals are persistent in the environment and have been shown to cause harm to humans when above a required threshold. If these metals go undetected in drinking and other water sources, they can become a substantial health issue. Electrochemical heavy metal analysis dates back to 1996 with the development of a microfabricated analysis system. [7] Since this time, electrochemical sensors and biosensors have been developed with dierent materials, configurations, and styles. For manganese detection specifically, electrode materials that have been used include carbon, lithium doped or polymer coated indium tin oxide, bismuth, copper based palladium, or platinum. [8–16] The main advancement that the field is relying on is the ability to mass produce reliable sensors, that require minimal pretreatment, at a low cost without hindering 2 on the obtained limits of detection. Finally, the use of electrochemical methods to destroy organic compounds that are toxic to humans are of utmost importance to mineralize the contaminants once and for all and prohibit a contamination cycle from continuing. Using electrochemistry aimed at remediating organic contaminants has been around since the early 1960s. These contaminated solutions include textile dyes, petrochemicals, paper industry euents, tannery euents, and other industrial wastewater. [17] As industry advances and develops new chemicals, the risks associated with them, and their need for remediation, also increases. To improve upon this technique for modern times, a system would need to be adapted to allow for the removal of ultra low concentrations of contaminants, in a true environmental matrix, that are unable to be broken down by competing technologies in a way that remains energy ecient. These three important tasks can be accomplished using the electrochemical methods of cyclic voltammetry, square wave voltammetry, square wave stripping voltammetry, and electrochemical oxidation. 1.2 Electrochemical Methods The main electrochemical methods used in this dissertation include those of cyclic voltammetry (CV), square wave voltammetry (SWV), square wave stripping voltammetry (SWSV), and elec- trochemical oxidation (EO). The next few sections outline the basic underlying principles that are important in further chapters. Each of the voltammetry measurements utilizes a three-electrode setup comprised of a working, reference and auxiliary electrode shown either of the configurations in Figure 1.1. The three electrodes are connected to a potentiostat. Electrochemical oxidation utilizes a two-electrode setup with the use of an anode and a cathode that are connected to a powersupply. The setup for an electrochemical oxidation experiment is shown in Figure 1.2. 3 Figure 1.1: (left) diagram of a beaker setup for electrochemical measurements and (right) a clamp cell setup that can also be used alternatively. Each has (A) a working electrode, (B) a reference electrode, and (C) a counter/ auxiliary electrode Figure 1.2: Setup used for an electrochemical oxidation experiment including an (A) anode, (B) cathode, and (C) stir bar 1.2.1 Cyclic Voltammetry CV is performed by scanning the potential between the working and reference electrodes in a negative or positive direction and measuring the resulting current response. Typically, an experiment would start at one potential, scan out in either direction to a further potential in a linear fashion, 4 and return to the starting point to complete one cycle. During this time, the current is measured. A figure depicting the potential waveform used in CV and the corresponding current response curve is shown in Figure 1.3. In order to understand repeatability and monitor change over time, typically 4 to 6 cycles are performed. The rate at which the scan is performed is also an important experimental parameter. Figure 1.3: (left) Potential waveform for cyclic voltammetry and (right) the corresponding current response There is a relationship between the scan rate and the current that is measured on the working electrode that is known as the capacitance, Equation 1.1. [18] i = C ⇤ dV dt (1.1) Where i is the current (A), V is the potential in V, t is the time in seconds, and C is the capacitance (F). By selecting a potential and recording the current output at various scan rates, a plot of the current versus the scan rate can be generated which provides the capacitance as the slope (A*s/V or F). CV is also important to determine at which potentials chemical reactions take place and how fast they occur on the working electrode’s surface. Since most electrolytes are aqueous, the workable potential window of a specific electrode material can be determined by measuring the potentials between the water oxidation and reduction reactions. This will also provide knowledge on if the electrode can be oxidized or reduced and thus hinder its ability to detect other compounds in solution. 5 How fast a reaction occurs on a given electrode can also be determined using CV. This allows for the determination of the heterogenous rate constant for the material and allows characterization and comparison of dierent electrode materials. Using a common redox couple, such as potassium ferrocyanide, a CV experiment can be performed and the Randles-Sevcik equation can be used that relates the peak heights measured, the area of the working electrode exposed in solution, the scan rate used, and the diusion coecient of the redox couple as seen in Equation 1.2. ip = (2.69x105)⇤ n 3 2 ⇤ D 1 2 ⇤ C ⇤ A ⇤ v 1 2 (1.2) Where ip is the peak current (A), n is the number of electrons in the reaction, D is the diusion coecient (cm2/s), C is the concentration of the redox compound (mol/cm3), A is the exposed area of the electrode (cm2), and v is the scan rate (V/s). Once the diusion coecient is determined from this equation it can be used in two dierent methods to determine the rate constants for a redox couple on an electrode material. The first method is known as the Nicholson Shain method and can be utilized when the separation between the oxidation and reduction peaks is less than 212 mV. [19] The equation that describes this method is provided as Equation 1.3. ko = '(⇡Doxv nF RT) ↵ ( Dox 2 Dred) 1 2 (1.3) Where ko is the heterogeneous rate constant (cm/s), ' is a kinetic parameter that is tabulated based on peak potential separation at a set temperature for a one electron process and can be found in a table in reference [19], Dred and Dox are the diusion coecients for the reduced and oxidized form of the redox couple, respectively (cm2/s), and ↵ is the transfer coecient that usually has a value of 0.5. The type of diusion that a redox couple has on a working electrode material can be determined based on graphing the peak currents obtained from multiple scan rates against the square root of the scan rate. If this relationship is linear, there is semi-infinite linear diusion. 6 The second method used for rate constant determination is the Klinger and Kochi method that utilizes fast scan CV for compounds that traditionally have peak separations larger than 212 mV or those that show a steady state response. [20] Fast scan CV refers to using scan rates that are >100 V/s compared to traditional scans in the mV/s range. The equation for this method is provided as Equation 1.4. [20] ko = 2.18(D↵nv F RT ) 1 2 e(↵2n F RT )(Eox p Ered p ) (1.4) Where Epox and Epred are the peak potentials for the oxidation and reduction peaks, respectively. This method will also provide information on the diusion of the process and again shows semi- infinite linear diusion when the forward peak currents have a linear relationship with the square root of the scan rate. The method of CV provides the tools to characterize and compare dierent working electrode materials. It is being utilized in the following chapters to advance knowledge on the properties that boron-doped diamond has as a quasi-reference material, what factors change the electrochemical responses seen with indium tin oxide, and if there are surface adsorption reactions occurring with fluoride in the use of BDD for PFAS oxidation. To further use these materials for sensing and oxidation experiments, more advanced methods need to be utilized. 1.2.2 Square Wave Voltammetry Square wave voltammetry (SWV) is similar to CV in that a potential range is scanned and an output current is measured, however, this method utilizes a staircase waveform. This allows for background noise suppression, high sensitivity, and the diagnostic ability of normal pulse voltammetry. [21] The current is recorded at the first pulse per cycle and at the end of the second pulse allowing for a dierence current between the two points to be calculated. [21] A figure depicting this waveform in provided in Figure 1.4 as well as the corresponding graph depicting a typical current response. This pause after the potential step, before the current is measured, allows for non-faradaic current to dissipate so that there is a lower background current and, subsequently, lower limits of detection for 7 the specified compound can be obtained. Non-faradaic current is the current that is not associated with a the direct consumption of reactants to form products describes by Faraday’s law such as the current from the formation of the double layer at the electrode’s surface. [18] Due to the sensitivity of this technique care must be taken to ensure the cleanliness of the analysis equipment to avoid interference from other compounds. Figure 1.4: (left) Potential waveform and (right) typical forward, reverse, and dierence current response for square wave voltammetry This method can be utilized in combination with the standard addition method to obtain the limit of detection (LOD) for a given compound when using a specific working electrode material. In order to do this, a scan is performed with the desired working electrode in a pure electrolyte solution. Next a set concentration of analyte is added to the solution and the scan is repeated. After this, an additional concentration is added and the scan is performed again. The cycle of adding analyte and scanning the potential range continues until there are enough scans that a graph can be generated for the current obtained versus the added concentration. This graph should provide a linear relationship between the concentration and the measured current. The slope of this line provides the sensitivity (m) that the electrode has towards the specific compound detection. In order to find the LOD, the scan for the lowest concentration of analyte is repeated 3 times. [22] The standard deviation of the peak height (s) is then multiplied by 3 and divided by the obtained sensitivity to find the LOD, Equation 1.5. In order to get statistical data, the experiment can be repeated multiple times in order to obtain a standard deviation. This method is utilized in the 8 following chapters to show that full BDD sensors, comprising of a working, auxiliary, and quasi reference electrode, can be developed to reach lower detection limits for a biological compound is a shorter amount of time than other techniques. LOD = (3s) m (1.5) This method is sucient for a variety of biological compounds, but in order to do heavy metal analysis an additional step is needed. This leads to the method of square wave stripping voltammetry. 1.2.3 Square Wave Stripping Voltammetry Stripping voltammetry advances SWV by adding in a pre-concentration or deposition step. In this pre-concentration/deposition step of anodic stripping voltammetry (ASV), the metal ion is reduced from its ionic state to its elemental state onto the working electrode’s surface. In cathodic stripping voltammetry (CVS), the metal ion is oxidized from its ionic state to a higher oxidation state and plates onto the electrode’s surface as an oxide. After sucient pre-concentration, the analyte is eectively removed, or stripped, from the electrode surface by a potential sweep in the opposite direction from the pre-concentration potential. In general, these methods are known as square wave stripping voltammetry (SWSV) when they use the same square wave waveform as previously mentioned. The solution is stirred during the preconditioning step where a sucient potential is applied to plate the analyte on the electrodes surface, the solution settles during a few second rest, then the analyte is removed from the electrode’s surface through scanning the potential in the opposite direction. The important parameters for this method include the deposition potential and time. These must be carefully chosen to ensure adequate deposition of the metal onto the electrode’s surface without sacrificing on measurement time. Not only can this method be used to determine the concentration of ions in solution, but it can also provide information about the electrochemical activity of an electrode based on where the ions are depositing onto the surface. For example, a metal can be plated onto the surface using 9 the pre-conditioning/ deposition step before removing the electrode from solutions and imaging the surface to determine where the metal was actually deposited. When using anodic or cathodic stripping voltammetry, the same approach used for SWV to determine the LOD of a compound can be used. This method is utilized in the following chapters to advance the understanding of indium tin oxide (ITO) as an electrode material for sensing applications in a way that lowers costs of sensor fabrication with the ability to mass produce sensors from this material and allows for repeatable and reliable results with minimal pretreatments. 1.2.4 Electrochemical Oxidation Electrochemical oxidation (EO) has been used for wastewater treatment studies and commercial applications throughout history. [23–30] This method works by direct or indirect oxidation at atmospheric conditions with no additional chemicals. The oxidation occurs with the transfer of one electron from the species in solution to the anode or the generation of oxidizing agents at the anodes surface that can then oxidize the contaminant in the bulk solution. There is no need for a reference electrode in this experimental setup. An anode and a cathode are directly connected to a power supply that controls the applied current or voltage. Oxidation occurs at the anode and involves the transfer of an electron from the species of interest in the electrolyte solution to the electrode, subsequently oxidizing the species in solution through a direct electron transfer process. In order for this exchange to occur in a direct process, the species is required to be within an appropriate distance from the electrode surface before diusing away. Therefore, both mass transport and electron transfer are important. This oxidation can also happen by having other compounds in solution transform into oxidizing agents, such as hydroxyl radicals, that can oxidize the compound of interest away from the electrode’s surface. This is known as indirect oxidation. This method is utilized in the following chapters to advance the management of PFAS waste from ion exchange processes in order to stop the contamination cycle from continuing to endanger the population, while exploring energy conscious techniques. 10 1.3 Background on Electrochemical Sensing The three classifications of electrochemical sensing are conductimetric, potentiometric, and voltammetric. The work here focused on voltammetric sensing which is based on recording a current-voltage curve over the course of time. Advantages of these electrochemical sensors include the choice of electrode material having an influence on the ability to separate two similarly oxidizable or reducible species, having high selectivity and specificity due to species undergoing a reaction at a specific potential, having the ability to apply complex potential waveforms leading to sensors with low detection limits, and the ability to miniaturize these sensors. [31] The first instance of voltammetric sensing came about after the invention of the dropping mercury electrode (DME) in 1922. [31] This electrode allowed for the study of reducible species due to its wide negative potential window. Shortly after in 1925, a polarograph was invented that allowed for a current- voltage curve to be recorded photographically, but it was not until 1939 that the first commercial instrument became availible. [32] In order to advance sensing from the negative potential range to cover both the negative and the positive potential range, more advanced solid electrode materials began to be investigated. These included materials such as platinum, gold, or carbon based films. In the 1950s more advanced polarography appeared, including the use of square wave, when circuits became available. [31] Once the development of the instrumentation became widely available, work in stripping analysis began to be investigated and compared to the standard voltammetric analysis that was previously performed. [32] Both linear scan voltammetry and cyclic voltammetry grew in popularity in the 1960s. [32] Microelectrodes began to make an appearance after this time and the first comprehensive review of them was written in 1981. [32] It was found that microelectrodes have advantageous properties such as immunity to solution resistance, an increase in current density, and a lower double- layer capacitance. [32] Additionally, in the 1980s fast-scan linear sweep voltammetry and cyclic voltammetry were added to high performance liquid chromatography (HPLC) to extend the range of analytes that could be detected. [32] All these advancements require the understanding of the 11 material being used as the working electrode. The ability for a material to be applicable as a good electrode material for voltammetric sensing can rely on its structure. Two newer electrode materials are that of boron-doped diamond (BDD) and indium tin oxide (ITO). Discussed further in Section 1.3.1 are the structures of the electrode materials of BDD and ITO and how they play a role in the properties that each material possesses. 1.3.1 Structure-property Connections 1.3.1.1 Boron-doped Diamond The process parameters used when depositing a BDD film can have an influence on its physio- chemical properties. These parameters include the level of boron dopant. The boron doping level of a diamond film can have an influence on the surface and physical properties such as the electrical conductivity, sp2 carbon content, surface termination, and surface area. [33] Boron content influences the grain size of the material with films that have a higher boron concentration have smaller grain sizes. [33, 34] The wetting ability of the electrode’s surface is also an influence of the boron content of the film. It has been shown that film’s with a lower boron concentration have a hydrophobic surface whereas those with a higher boron content show characteristics of an oxygen terminated film. [33, 34] A lower boron concentration allows for electrodes with an increased electrochemical potential window due to their higher resistivity. [34, 35] In addition, a higher boron content allows for a higher carrier mobility, a necessary parameter for electrochemical sensing. [35] One can also relate the boron content of a diamond film to its ability to be used in certain electrochemical applications. For example, electrochemical nitrate reduction has been shown to have a larger peak potential and occur at a more positive potential for a BDD film with a higher boron content that has been attributed to a higher conductivity value. [33]. The highly doped film ultimately had a higher sensitivity and lower detection limit for nitrate compared to the other that was not as heavily doped. The boron content has also been related to the detection of uric acid on BDD films. An increase in the linear detection range and a decrease in the detection limit is 12 obtained with an increase in boron concentration. [34] For the oxidation of sulphamerazine on BDD, a higher concentration of boron to carbon was more eective. [35] Similarly, the eciency of the electrochemical oxidation of azo dyes was influenced by the boron concentration of the film. An increase the doping level increased the eciency of the process. [36] Indium Tin Oxide 1.3.1.2 Similar to BDD, the deposition parameters used when depositing ITO through sputtering can aect the structural, electrical, and optical properties of ITO that can come into play when utilizing the material as an electrode for a sensing application. The first process dependent report on the use of ITO for electrochemical applications was published in 1992. [37] More recently, a report detailing how the microstructure and doping of the film can have an influence on it’s electrochemical properties was published in 2002. [38] An electrochemical comparison between an as grown ITO film and one with the application of vacuum annealing has been made. It was found that more reversible behavior was obtained on the film that was vacuum annealed compared to the as-grown film. [37] This is suggested to be due to the fact that the resistivity decreases with annealing and increases the free carrier concentration. [37] The vacuum annealed films also allowed for a lower peak separation in peak potential than those obtained on as-grown films due to their increase in oxygen vacancies. [37] Although this study had very limited film characterization, it showed that the film that went through post-deposition annealing in a vacuum had a superior performance than the as-grown film. ITO can have an amorphous or crystalline microstructure. The shift to crystallinity comes from the application of heat during deposition. In amorphous films, the carrier concentration is low with a high mobility due to the absence of crystal defects, but upon crystallization, the carrier density increases and the mobility decreases. [38] It has also been shown that the tin in an amorphous state is not electrically active in contrast to the crystalline material [38] The electron transfer rates for three common redox couples were lower for the amorphous film with lower peak currents and larger peak separations. [38] This study provided evidence that the amorphous ITO film was 13 not an adequate electrode for the oxidation of guanine. This study provides evidence that the microstructure of the film has an influence on its ability to be used in electrochemical applications such as voltammetric sensing. 1.3.2 Prior State of the Art Since this report will cover three distinct topics related to elecrochemistry, it is important to provide the prior state of the art for each. When considering the electrochemical detection of isatin, there are three primary works prior to this investigation. Two of which cover the electrochemical detection of isatin after an initial system such as high-performance liquid chromatography (HPLC) or flow injection analysis (FIA), whereas, the third looks directly at the first attempt to detect isatin solely based on electrochemical methods. [6, 39, 40] Table 1.1 shows each study and the figures of merit that were obtained. The dierential pulse study obtained similar limits of detections (LOD) to both the electrodes used after FIA and the glassy carbon electrode used after HPLC. The lowest obtained LOD was from using the chemically modified palladium hexacyanoferrate electrode after HPLC. However, it is important to simplify the electrode requirements for this process to make it more accessible. Manganese detection using electrochemistry is not a new topic. To consolidate the prior state of the art research, the detection of only manganese specifically using cathodic stripping voltammetry in the last 25 years was researched. Working electrodes without any use of membranes or additional techniques, such as ultrasonic or prior column methods, were focused on that had no other interfering metals. The prior art scope was also limited to manganese detection in simulated solutions so that it would be more applicable to the study performed here. Electrode materials in this prior art research include modified carbon paste, glassy carbon, carbon film, graphite composite, boron-doped diamond, palladium, carbon nanotube, indium tin oxide, graphite, and platinum.[41, 42, 8, 43–46, 14, 47–49] Table 1.2 compares the obtained limit of detection and linear range for each study as well as the pretreatment the working electrode went through It is important to consider an economical working and the electrochemical parameters used. 14 Table 1.1: Summary table outlining the prior art for the electrochemical detection of isatin electrode material for this application with minimal pretreatments to be able to advance this sensing technology to a single use sensor for the medical or environmental field. The prior art research focused on perfluoroalkyl substance (PFAS) oxidation with electrochem- istry focused on the studies that looked at using BDD as the anode material. There are five categories included in this review: those involving industrial waste waters, those in natural water matrix or ground water, simulated solutions with only perfluorooctanoic acid (PFOA), simulated solutions starting with perfluorooctane sulfonate (PFOS) or perfluorobutane sulfonic acid (PFBS), and studies that start with a mixture of PFAS. These categories are split into two tables, Table 1.3 that covers the simulated solutions studies and Table 1.4 that shows the PFAS studies performed in industrial or ground water solutions. Table 1.3 covers four studies with just PFOA, one study with just PFOS, one study with just 15 Table 1.2: Summary table outlining the prior art for the electrochemical detection of manganese in simulated buer solutions PFBS, and one study with a mixture of a few chain lengths. [50–55, 27] Some of the removal percentages were estimated based on the provided figures in the literature. Most studies achieved greater than 50 % degradation for their perspective compound in under 10 hours. Table 1.4 shows the studies that were performed in complex solutions. These include those of industrial wastewater and contaminated groundwater. [56–61] All of those obtained removal percentages of greater than 90 %. In terms of advancing the prior state of the art, it is important to look at minimizing the treatment time needed for these solutions while also reducing the energy consumption that is required. 16 Table 1.3: Summary table outlining the prior art for the simulated PFAS solutions Table 1.4: Summary table outlining the prior art for the complex solutions containing PFAS 1.3.3 Deficiencies of Technology Although electrochemical voltammetric sensing has advanced substantially over the years, there are still a few deficiencies in the technology. Noble metals and advanced diamond materials can be costly. In addition, in order to be more convenient, electrodes themselves need to be miniaturized 17 as well as the necessary equipment. For example the development of sensors that comprise of the working, counter, and auxiliary electrodes all on a single platform. Both cost eective materials and miniaturization of the sensing technology will allow for more accessible sensors and more single use sensors. In addition, the reduction in time and energy for PFAS remediation will play an important role for it to be used in industry. 18 BIBLIOGRAPHY 19 BIBLIOGRAPHY [1] JM Halket, PJ Watkins, A Przyborowska, BL Goodwin, Angela Clow, Vivette Glover, and M Sandler. Isatin (indole-2, 3-dione) in urine and tissues: Detection and determination by gas chromatography—mass spectrometry. Journal of Chromatography B: Biomedical Sciences and Applications, 562(1-2):279–287, 1991. [2] Shigeo Manabe, Qiang Gao, Juan Yuan, Toshie Takahashi, and Akira Ueki. Determina- tion of isatin in urine and plasma by high-performance liquid chromatography. Journal of Chromatography B: Biomedical Sciences and Applications, 691(1):197–202, 1997. [3] Zeid A Alothman, Masoom R Siddiqui, and Saikh M Wabaidur. Estimation of isatin in spiked plasma samples using high performance liquid chromatography. Chemical Industry and Chemical Engineering Quarterly, 19(4):605–614, 2013. [4] Shigeo Manabe, Qiang Gao, Juan Yuan, Toshie Takahashi, and Akira Ueki. Determina- tion of isatin in urine and plasma by high-performance liquid chromatography. Journal of Chromatography B: Biomedical Sciences and Applications, 691(1):197–202, 1997. [5] Alexei E Medvedev, Angela Clow, Merton Sandler, and Vivette Glover. Isatin: a link between natriuretic peptides and monoamines? Biochemical pharmacology, 52(3):385–391, 1996. [6] Haihong Xu, Dan Wang, Wen Zhang, Wei Zhu, Katsunobu Yamamoto, and Litong Jin. Deter- mination of isatin and monoamine neurotransmitters in rat brain with liquid chromatography using palladium hexacyanoferrate modified electrode. Analytica chimica acta, 577(2):207– 213, 2006. [7] Richard J Reay, Anthony F Flannery, Christopher W Storment, Samuel P Kounaves, and Gre- gory TA Kovacs. Microfabricated electrochemical analysis system for heavy metal detection. Sensors and Actuators B: Chemical, 34(1-3):450–455, 1996. [8] Christopher M.A. Brett Olga M.S. Filipe. Cathodic stripping voltammetry of trace mn(ii) at carbon film electrodes. Talanta, 61(5):643–650, 2003. Junwei Di and Feng Zhang. Voltammetry determination of trace manganese with pretreatment glassy carbon electrode by linear sweep voltammetry. Talanta, 60(1):31–36, 2003. [9] [10] Yeo May Ching, Tan Wee Tee, and Zulkarnain Zainal. Electrochemical studies of mn (ii) mediated by li+ doped indium tin oxide (ito) electrode. Int. J. Electrochem. Sci, 6:5305–5313, 2011. [11] Preetha Jothimuthu, Robert A Wilson, Josi Herren, Erin N Haynes, William R Heineman, and Ian Papautsky. Lab-on-a-chip sensor for detection of highly electronegative heavy metals by anodic stripping voltammetry. Biomedical microdevices, 13(4):695–703, 2011. 20 [12] He Xu, Qiaoli Zheng, Ping Yang, Jianshe Liu, and Litong Jin. Sensitive voltammetric detection of trace heavy metals in real water using multi-wall carbon nanotubes/nafion composite film electrode. Chinese Journal of Chemistry, 29(4):805–812, 2011. [13] Cory A Rusinek, Wenjing Kang, Keaton Nahan, Megan Hawkins, Cooper Quartermaine, Angela Stastny, Adam Bange, Ian Papautsky, and William R Heineman. Determination of manganese in whole blood by cathodic stripping voltammetry with indium tin oxide. Electroanalysis, 29(8):1850–1853, 2017. [14] Cory A Rusinek, Adam Bange, Mercedes Warren, Wenjing Kang, Keaton Nahan, Ian Pa- pautsky, and William R Heineman. Bare and polymer-coated indium tin oxide as working electrodes for manganese cathodic stripping voltammetry. Analytical chemistry, 88(8):4221– 4228, 2016. [15] Sophia Karastogianni and Stella Girousi. Electrochemical behavior and voltammetric de- termination of a manganese (ii) complex at a carbon paste electrode. AnAlyticAl chemistry insights, 11:ACI–S32150, 2016. [16] Mahmoud Roushani, Zahra Saedi, Fatemeh Hamdi, and Behruz Zare Dizajdizi. Preparation an electrochemical sensor for detection of manganese (ii) ions using glassy carbon electrode modified with multi walled carbon nanotube-chitosan-ionic liquid nanocomposite decorated with ion imprinted polymer. Journal of Electroanalytical Chemistry, 804:1–6, 2017. [17] Sergi Garcia-Segura, Joey D Ocon, and Meng Nan Chong. Electrochemical oxidation reme- diation of real wastewater euents—a review. Process Safety and Environmental Protection, 113:48–67, 2018. [18] Thomas F Fuller and John N Harb. Electrochemical engineering. John Wiley & Sons, 2018. [19] Richard S Nicholson. Theory and application of cyclic voltammetry for measurement of electrode reaction kinetics. Analytical Chemistry, 37(11):1351–1355, 1965. [20] RJ Klingler and JK Kochi. Electron-transfer kinetics from cyclic voltammetry. quantitative description of electrochemical reversibility. The Journal of Physical Chemistry, 85(12):1731– 1741, 1981. [21] Allan J. Bard and Larry R. Faulkner. Electrochemical Methods Fundamentals and Applica- tions. John Wiley sons, Inc., United States of America, second edition, 2011. [22] Daniel C. Harris. Quantitative Chemical Analysis. Clancy Marshall, New York, 8th edition, 2010. [23] Angela Anglada, Ane Urtiaga, and Inmaculada Ortiz. Contributions of electrochemical oxidation to waste-water treatment: fundamentals and review of applications. Journal of chemical technology and biotechnology, 84(12):1747–1755, 2009. [24] Minghua Zhou, Heikki Särkkä, and Mika Sillanpää. A comparative experimental study on methyl orange degradation by electrochemical oxidation on bdd and mmo electrodes. Separation and Purification Technology, 78(3):290–297, 2011. 21 [25] Qiongfang Zhuo, Meiqing Luo, Qingwei Guo, Gang Yu, Shubo Deng, Zhencheng Xu, Bo Yang, and Xiaoliang Liang. Electrochemical oxidation of environmentally persistent perfluorooctane sulfonate by a novel lead dioxide anode. Electrochimica Acta, 213:358–367, 2016. [26] Qiongfang Zhuo, Shubo Deng, Bo Yang, Jun Huang, and Gang Yu. Ecient electrochemical oxidation of perfluorooctanoate using a ti/sno2-sb-bi anode. Environmental science technol- ogy, 45(7):2973–2979, 2011. [27] Qiongfang Zhuo, Shubo Deng, Bo Yang, Jun Huang, Bin Wang, Tingting Zhang, and Gang Yu. Degradation of perfluorinated compounds on a boron-doped diamond electrode. Elec- trochimica Acta, 77:17–22, 2012. [28] Marek Trojanowicz, Anna Bojanowska-Czajka, Iwona Bartosiewicz, and Krzysztof Kulisa. Advanced oxidation/reduction processes treatment for aqueous perfluorooctanoate (pfoa) and perfluorooctanesulfonate (pfos)–a review of recent advances. Chemical Engineering Journal, 336:170–199, 2018. [29] BV Sarada, Tata N Rao, DA Tryk, and A Fujishima. Electrochemical oxidation of histamine and serotonin at highly boron-doped diamond electrodes. Analytical chemistry, 72(7):1632– 1638, 2000. [30] MA Rodrigo, P Cañizares, A Sánchez-Carretero, and C Sáez. Use of conductive-diamond electrochemical oxidation for wastewater treatment. Catalysis Today, 151(1-2):173–177, 2010. [31] Christopher MA Brett and Ana Maria Oliveira-Brett. Electrochemical sensing in solu- tion—origins, applications and future perspectives. Journal of Solid State Electrochemistry, 15(7-8):1487–1494, 2011. [32] Allen J Bard and Cynthia G Zoski. Voltammetry retrospective., 2000. [33] JT Matsushima, WM Silva, AF Azevedo, MR Baldan, and NG Ferreira. The influence of boron content on electroanalytical detection of nitrate using bdd electrodes. Applied surface science, 256(3):757–762, 2009. [34] Peng Yu, Jiawei Zhang, Tong Zheng, and Ting Wang. Influence of boron doped level on the electrochemical behavior of boron doped diamond electrodes and uric acid detection. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 494:241–247, 2016. [35] Aleksandra FabiaÒska, Robert Bogdanowicz, Patrycja Zi˛eba, Tadeusz Ossowski, Marcin Gnyba, Jacek Ryl, Artur Zielinski, Stoel D Janssens, Ken Haenen, and Ewa M Siedlecka. Electrochemical oxidation of sulphamerazine at boron-doped diamond electrodes: Influence of boron concentration. physica status solidi (a), 210(10):2040–2047, 2013. [36] Robert Bogdanowicz, Aleksandra FabiaÒska, L Golunski, Micha≥ Sobaszek, Marcin Gnyba, Jacek Ryl, Kazimierz Darowicki, Tadeusz Ossowski, Stoel D Janssens, Ken Haenen, et al. Influence of the boron doping level on the electrochemical oxidation of the azo dyes at si/bdd thin film electrodes. Diamond and related materials, 39:82–88, 2013. 22 [37] MA Martinez, J Herrero, and MT Gutierrez. Electrochemical stability of indium tin oxide thin films. Electrochimica acta, 37(14):2565–2571, 1992. [38] Natasha D Popovich, Sze-Shun Wong, Brian KH Yen, Hyo-Young Yeom, and David C Paine. Influence of microstructure on the electrochemical performance of tin-doped indium oxide film electrodes. Analytical chemistry, 74(13):3127–3133, 2002. [39] Romana Jaroöová, Simon Sanchez, Lars Haubold, and Greg M Swain. Isatin analysis using flow injection analysis with amperometric detection–comparison of tetrahedral amorphous carbon and diamond electrode performance. Electroanalysis. [40] Victor C Diculescu, Sunita Kumbhat, and Ana Maria Oliveira-Brett. Electrochemical be- haviour of isatin at a glassy carbon electrode. Analytica chimica acta, 575(2):190–197, 2006. [41] Soo Beng Khoo, Ming Koon Soh, Qiantao Cai, Miriam R Khan, and Si Xuan Guo. Dierential pulse cathodic stripping voltammetric determination of manganese (ii) and manganese (vii) at the 1-(2-pyridylazo)-2-naphthol-modified carbon paste electrode. Electroanalysis, 9(1):45– 51, 1997. [42] John Stanislaus Roitz and KW Bruland. Determination of dissolved manganese (ii) in coastal and estuarine waters by dierential pulse cathodic stripping voltammetry. Analytica Chimica Acta, 344(3):175–180, 1997. [43] Ji-Ye Jin, Fan Xu, and Tomoo Miwa. Cathodic stripping voltammetry for determination of trace manganese with graphite/styrene-acrylonitrile copolymer composite electrodes. Elec- troanalysis: An International Journal Devoted to Fundamental and Practical Aspects of Electroanalysis, 12(8):610–615, 2000. [44] Craig E Banks, Jaanus Kruusma, Ryan R Moore, Peter Tomík, Judith Peters, James Davis, äe- bojka Komorsky-LovriÊ, and Richard G Compton. Manganese detection in marine sediments: anodic vs. cathodic stripping voltammetry. Talanta, 65(2):423–429, 2005. [45] Wenjing Kang, Xing Pei, Adam Bange, Erin N Haynes, William R Heineman, and Ian Pa- pautsky. Copper-based electrochemical sensor with palladium electrode for cathodic stripping voltammetry of manganese. Analytical chemistry, 86(24):12070–12077, 2014. [46] Wei Yue, Adam Bange, Bill L Riehl, Bonnie D Riehl, Jay M Johnson, Ian Papautsky, and William R Heineman. Manganese detection with a metal catalyst free carbon nanotube electrode: anodic versus cathodic stripping voltammetry. Electroanalysis, 24(10):1909–1914, 2012. [47] M Christine Welch, E Craig Banks, äebojka Komorsky-LovriÊ, and G Richard Compton. Electroanalysis of trace manganese via cathodic stripping voltammetry: Exploration of edge plane pyrolytic graphite electrodes for environmental analysis. 2006. [48] Miroslav Rievaj, Peter Tomcik, Zuzana Janosikova, Du≥an Bustin, and Richard G Compton. Determination of trace mn (ii) in pharmaceutical diet supplements by cathodic stripping voltammetry on bare carbon paste electrode. Chem. Anal.(Warsaw), 53:153–161, 2008. 23 [49] Wenjing Kang, Cory Rusinek, Adam Bange, Erin Haynes, William R Heineman, and Ian Pa- pautsky. Determination of manganese by cathodic stripping voltammetry on a microfabricated platinum thin–film electrode. Electroanalysis, 29(3):686–695, 2017. [50] Tsuyoshi Ochiai, Yuichi Iizuka, Kazuya Nakata, Taketoshi Murakami, Donald A Tryk, Akira Fujishima, Yoshihiro Koide, and Yuko Morito. Ecient electrochemical decomposition of perfluorocarboxylic acids by the use of a boron-doped diamond electrode. Diamond and Related Materials, 20(2):64–67, 2011. [51] Ane Urtiaga, Carolina Fernández-González, Sonia Gómez-Lavín, and Inmaculada Ortiz. Ki- netics of the electrochemical mineralization of perfluorooctanoic acid on ultrananocrystalline boron doped conductive diamond electrodes. Chemosphere, 129:20–26, 2015. [52] Hanshuang Xiao, Baoying Lv, Guohua Zhao, Yujing Wang, Mingfang Li, and Dongming Li. Hydrothermally enhanced electrochemical oxidation of high concentration refractory perfluorooctanoic acid. The Journal of Physical Chemistry A, 115(47):13836–13841, 2011. [53] NW Khun, E Liu, and XT Zeng. Corrosion behavior of nitrogen doped diamond-like carbon thin films in nacl solutions. Corrosion Science, 51(9):2158–2164, 2009. [54] Kimberly E Carter and James Farrell. Oxidative destruction of perfluorooctane sulfonate using boron-doped diamond film electrodes. Environmental science technology, 42(16):6111–6115, 2008. [55] Zhaohui Liao and James Farrell. Electrochemical oxidation of perfluorobutane sulfonate using boron-doped diamond film electrodes. Journal of applied electrochemistry, 39(10):1993– 1999, 2009. [56] Beatriz Gomez-Ruiz, Sonia Gómez-Lavín, Nazely Diban, Virginie Boiteux, Adeline Colin, Xavier Dauchy, and Ane Urtiaga. Boron doped diamond electrooxidation of 6: 2 fluorotelom- ers and perfluorocarboxylic acids. application to industrial wastewaters treatment. Journal of Electroanalytical Chemistry, 2017. [57] Beatriz Gomez-Ruiz, Sonia Gómez-Lavín, Nazely Diban, Virginie Boiteux, Adeline Colin, Xavier Dauchy, and Ane Urtiaga. Ecient electrochemical degradation of poly-and perflu- oroalkyl substances (pfass) from the euents of an industrial wastewater treatment plant. Chemical Engineering Journal, 322:196–204, 2017. [58] AM Trautmann, H Schell, KR Schmidt, K-M Mangold, and A Tiehm. Electrochemical degradation of perfluoroalkyl and polyfluoroalkyl substances (pfass) in groundwater. Water Science and Technology, 71(10):1569–1575, 2015. [59] Álvaro Soriano, Daniel Gorri, and Ane Urtiaga. Ecient treatment of perfluorohexanoic acid by nanofiltration followed by electrochemical degradation of the nf concentrate. Water Research, 112:147–156, 2017. [60] Charles E Schaefer, Christina Andaya, Aniela Burant, Charles W Condee, Ana Urtiaga, Timo- thy J Strathmann, and Christopher P Higgins. Electrochemical treatment of perfluorooctanoic 24 acid and perfluorooctane sulfonate: Insights into mechanisms and application to groundwater treatment. Chemical Engineering Journal, 317:424–432, 2017. [61] Charles E Schaefer, Sarah Choyke, P Lee Ferguson, Christina Andaya, Aniela Burant, Andrew Maizel, Timothy J Strathmann, and Christopher P Higgins. Electrochemical transformations of perfluoroalkyl acid (pfaa) precursors and pfaas in groundwater impacted with aqueous film forming foams. Environmental science & technology, 52(18):10689–10697, 2018. 25 CHAPTER 2 ISATIN DETECTION USING A BORON-DOPED DIAMOND 3-IN-1 SENSING PLATFORM 26 2.1 Introduction This work has been published in Analytical Chemistry1. Indole-2,3-dione, or isatin, has attractive properties as an endogenous indole that could be applied in the treatment of Parkinson’s Disease, tremors, or epilepsy. [1–3] Its structure is shown in Figure 2.1. Isatin has a large amount of biological activity and is found in the tissues and fluids of mammals. [4–7] In 1988 scientists discovered this compound using the technique of gas chromatography mass spectroscopy (GCMS). [4] Concentrations of isatin range from above 1 µmol/L in blood to 0.1 to 10 µmol/l in tissues and greater than 1 mmol/L in urine samples. [6, 8, 7] This compound has been of interest due to its ability to bring on stress and anxiety, the inhibition of sodium in rat intestines, and the potential link it may provide between a person’s mood, the natriuretic peptide system, and monoamine neurotransmitter levels. [6, 9, 10] Moderate concentrations of isatin, 15-20 mg/kg of body weight, have been found to be anxiogenic but a higher dose, 50 mg/kg of body weight, can induce sedation in rats and larger mammals such as monkeys. [9] To gain a better understanding of all the interactions isatin is capable of and how to take advantage of it as a pharmaceutical drug, a more complete understanding of how it interacts in the human body and the mechanisms it goes through is essential. To achieve this, precise measurement techniques for the detection of isatin are necessary. Since the discovery of isatin using GCMS, liquid chromatography (LC) coupled with UV-vis spectroscopy or an electrochemical voltammetric method have been used as detection methods. [1, 8, 7, 11–14] A downside to GCMS is the long measurement times required. [14] A research group using high performance liquid chromatography (HPLC) with UV-vis measured isatin’s absorbance peak at 242 nm with a limit of detection (LOD) of 2 µmol/L. [14] Another group using LC with electrochemical detection with a palladium hexacyanoferrate modified electrode measured isatin in rat brains with a LOD of 0.03 µmol/L. [1] Diculescu et al. used UV-vis with a found LOD of 0.8 µmol/L and began to study the detection possibilities of using electrochemical techniques without the addition of a chromatography techniques. [11] 1Analytical Chemistry 2018, DOI:10.1021/acs.analchem.7b04045 27 Figure 2.1: Molecular structure for isatin Isatin has both an oxidation and reduction reaction that are irreversible and dependent on pH that can be detected via electrochemistry. [11] When utilizing a glassy carbon (GC) electrode, isatin displays an oxidation peak at +1.08 V and a reduction peak -0.48 V. [7, 11] One group used a GC electrode and multiple voltammetry techniques to study isatin and were able to obtain a LOD of 0.19 µM when using dierential pulse voltammetry. [11] Other groups have also studied isatin on GC electrodes. [3, 7, 12] Boron-doped diamond (BDD) is another electrode material known for its outstanding properties when used for the electrochemical detection of biological compounds. [15–17] BDD has many electrochemical advantages such as its wide potential window, low background current, chemical inertness, and its good chemical and electrochemical stability. [18, 19] The wide potential window that is characteristic of BDD oers the ability to detect compounds that have higher positive or negative redox potentials. BDD electrodes have excellent microstructure stability even at these extreme potential values. [18, 19] With minimal preconditioning steps, BDD displays quasi- reversible redox systems with resistance to deactivation and fouling. [20, 21] Additionally, BDD can be fabricated into dierent electrode microstructures such as micro electrode arrays (MEAs) 28 that can then be used as micro-sensors for free hydroxyl radical generation and neural activity measurements. [21–29] One factor that can produce unwanted results in experiments is the potential drop due to solution resistance which can shift peak potential and increase peak separation. MEAs fabricated out of BDD oer a lower electrical potential dierence, or IR drop, due to their lower electrode surface area compared to typical macro (MAC) electrodes and they can provide steady state current responses that MAC electrodes can not. [23, 24] After a potential step, MEAs require a short time to reach equilibrium. [24] One group reported that an ultramicroelectrode array (UMEA) showed 50-100 times better sensitivity than a traditional MAC electrode with a superior reproducibility. [30] In order to simplify the usability of MEAs and prevent breakage and loss, sensing platforms have been fabricated that consist of the working, auxiliary, and reference electrodes. [30–40] Previously, A 3-in-1 sensor was made from silver (Ag) and gold (Au) electrodes to detect nitrate (NO3-) and another was fabricated with carbon nanotubes for the detection of glucose. [39, 40] Many of these sensors use Ag/AgCl or Ag/Ag+ as a reference or quasi-reference electrode material. [34, 37, 39, 40] One research group fabricated a device with a BDD quasi-reference material but failed to mention how it behaved. [35] However, two groups using sensors made with all BDD components, demonstrated that BDD could behave as an acceptable quasi-reference material. [38, 30] To evaluate BDD’s performance as a reference material, potential controlled experiments were performed to determine its stability for 57 hours and showed that the potential only drifted by +/- 8 mV. [38] For the remainder of this chapeter, two fully BDD 3-in-1 sensors, one MAC and one MEA, were fabricated and characterized. Both sensors were assessed for their ability to detect isatin. The BDD quasi-reference material was characterized based on its potential stability over time compared to a standard Ag/AgCl reference electrode in pH solutions of 4.0 to 7.4. The BDD MAC electrode had a working area of 3.14 x 10-2 cm2. The MEA’s total working area amounted to 1.71 x 10-4 cm2. The BDD material was characterized with Raman spectroscopy and scanning electron microscopy (SEM). The development of such a BDD sensing platform could be used in future 29 work to detect isatin in applications involving liquid chromatography coupled to electrochemical detection methods while removing the need for three separate electrodes and simplifying the required electrode pretreatment. 2.2 Experimental Section 2.2.1 Chemicals and Materials Potassium ferrocyanide (K4Fe(CN)6) and potassium ferricyanide (K3Fe(CN)6) were purchased from Mallinckrodt Specialty Chemicals and J.T. Baker, respectively. Isatin and hexaamine- ruthenium(III) chloride (Ru(NH3)6Cl) were purchased from Sigma Aldrich. A potassium chloride (KCl) solution of 1.0 M was prepared with DI water and salt from ChemPure Brand Chemicals. The KCl solution was used to prepare stock solutions of Fe(CN)63-/4- and Ru(NH3)63+/2+. A 10x phosphate buered saline (PBS) solution (10.6 mM monobasic potassium phosphate, 1551 mM sodium chloride, 29.7 mM dibasic sodium phosphate) was purchased from Gibco and diluted to 1x using DI water (1.1 mM monobasic potassium phosphate, 155 mM sodium chloride, 3.0 mM dibasic sodium phosphate) to a final pH of 7.4. The 10x PBS was also diluted and pH adjusted with hydrochloric acid from Fisher Scientific to obtain pH values of 4, 5, 6, and 7. The 0.1 M PBS was used to make a stock isatin solution. Surine Negative Urine Control was purchased from a Simga-Aldrich Company, Ceriliant. 2.2.2 Instrumentation A VCA Optima from AST Productions, Inc. was used to perform contact angle measurements with a drop size of 1 µL of water. The systems software determined the contact angles of each water drop on the sample’s surface. This measurement was repeated 3 times to obtain a mean value and a standard deviation. A single compartment glass cell with 5 mL of solution was used for cyclic and square wave voltammetry measurements. A clamp held each sensor in place once immersed in solution. Each 3-in-1 sensor was electrically connected via a micro USB connector to the leads from the potentiostat. Parameters for capacitance measurements using cyclic voltammetry 30 were: scan rates: 100-500 mV s-1, electrolyte: 1.0 M KCl. Square wave voltammetry parameters were: frequency: 15 Hz; Amplitude, 0.05 V; increment E, 0.004 V. A CHI660C analyzer from CH Instruments, Inc. was used. Nitrogen (N2) was bubbled for 5 minutes prior to the start of the isatin experiments unless stated otherwise. Response curves for isatin were generated using the standard addition method. the reproducibility of each electrode was determined from 4 scans at each concentration. The spectrophotometric analysis was performed using a UV-vis-NIR spectrophotoeter (UV-3600) from Simadzu with quartz cuvettes that had a 10 mm path length. 2.2.3 3-in-1 Electrode Fabrication Both the MAC and MEA were fabricated on the same conductive silicon (Si) wafer through multiple processing steps. Microwave plasma-assisted chemical vapor deposition (MP-CVD) was first used to grow the BDD layer on top of the 3 inch diameter Si/ SiO2 wafer. This wafer was p-type with a thickness o 350 µm and a resistivity between 1-5 Ohm cm. This was then coated with an insulating SiO2 top layer that was 1 µm thick. The sensors were patterned using a photolithography mask that allowed for the deposition of an insulating interlayer to be deposited on their surface using plasma- enhanced chemical vapor deposition (PE-CVD). Through all of these steps, the exposed BDD, which could be electrically connected at the top of each sensor, was the only electrochemically active surface on the face of the sensor. A second photoresist mask defined the MAC and MEA configurations and allowed for the three distinct electrodes on each to be patterned. The spacing between each electrode in the MEA was carefully decided. If the micro discs were too close together the advantages of the single small electrodes would be lost as their depletion layers would overlap. [22, 23] Each 3-in-1 sensor was released from the wafer by laser cutting. An earlier publication goes in to more detail on the fabrication process. [22] Figure 2.2 shows optical micrographs of both the MEA (left) and MAC (right) 3-in-1 sensors. The active surface area of the MAC was 3.14 x 10-2 cm2 and the 97 electrodes that comprised the hexagonal patterned MEA had a total active anode area of 1.71 x 10-4 cm2. 31 Figure 2.2: Optical micrograph of the MEA (left) and MAC (right) 3-in-1 sensors with exposed BDD 2.3 Results and Discussion 2.3.1 Electrode Characterization Boron is typically used as a dopant for diamond to charge carrier concentrations greater than 1020 cm-3 to obtain metal-like electrical conductivity. [41–45] The surfaces of the MAC and MEA were characterized using Raman spectroscopy and scanning electron microscopy (SEM). Figure 2.3 shows a Raman spectrum and SEM micrograph taken from the diamond-coated wafer that both sensors were fabricated from. The SEM micrograph, Figure 2.3 (A) shows grain sizes around 1-2 µm for the polycrystalline diamond film that was used for all sensors. The film has no visual evidence of cracks or other defects. The Raman spectrum has large peaks around 500 and 1200 cm-1 that are indicative of highly doped BDD. [41, 46–48] These peaks grow in intensity with higher boron concentrations. The boron concentration of the film was determined by using a fitting program in Matlab; providing a concentration of 0.3 %. Cyclic voltammetry was used to generate a current vs voltage curve to determine the capacitance 32 Figure 2.3: (A) An SEM micrograph of the polycrystalline BDD surface that was used to fabricate the two 3-in-1 sensors with no defects or large cracks in the films structure and (B) the corresponding Raman spectrum analysis for the BDD wafer values of the sensors. The MAC had a capacitance of 721 nF and the MEA had a capacitance of 5.3 nF. The drastic dierence between the two comes from the dierent surface areas that are exposed to solution. If the capacitances are normalized to each corresponding surface area they are 25 µF/cm2 and 31 µF/cm2 for the MAC and MEA, respectively. These capacitance values are in a similar range to those reported for BDD in literature, 20-40 µF/cm. [49] Cyclic voltammetry (CV) was used further to determine the electron-transfer kinetics for the two redox couples of Fe(CN)63-/4- and Ru(NH3)63+/2+. Fe(CN)64- was used for the MAC electrode configuration whereas the MEA used a 1:1 ratio of Fe(CN)63- and Fe(CN)64-. The concentrations of the redox couples were 1.0 mM in 1.0 M KCl. The peaks were analyzed after manual background subtraction. [50, 51] The cyclic voltammograms for both redox couples on each 3-in-1 sensor are shown in Figure 2.4. Each electrode was first conditioned in 1 M sulfuric acid (H2SO4) from -2.0 V to +2.8 V at 0.5 V/s for 30 scans to condition the electrode surface. This was decided based on the investigation 33 Figure 2.4: (A) Cyclic voltammograms of 1 mM Fe(CN)63- and 1 mM Ru(NH3)6Cl on the MAC 3-in-1 BDD sensor oth dissolved in 1 M KCl and (B) of a 1:1 ratio of Fe(CN)63-/4- and 1 mM of Ru(NH3)6Cl , both dissolved in KCl, on the 3-in-1 BDD MEA sensor. Scan rate: 0.1 V/s. into surface chemistry eects as discussed next. For obtaining dierent surface terminations, three dierent pretreatments were used: hydrogen termination (HT), oxygen termination (OT) and electrochemical cycling (ET). The ET was done in 1.0 M H2SO4 using the procedure stated above. For HT, the electrodes were placed in a microwave plasma system in ultra-pure hydrogen (H2) for 10 minutes. For the OT pretreatment, electrodes were placed in an oxygen (O2) atmosphere for 1 minute in a reactive ion etcher (RIE). Contact angle measurements were used to evaluate each pretreatment process based on the obtained angle. Contact angles of 90 +/- 4.5°were measured for HT whereas angles of 0°were measured for the OT pretreatments. These trends align with those seen in literature for BDD surfaces. [52] The contact angles for the ET pretreatment were predominantly oxygen terminated at 13 +/- 3.3°. However, these are slightly more hydrophobic than the OT contact angles, suggesting dierent functional groups and surface coverage. The ET process may be generating a slightly more hydrophilic surfaces due to a stepwise mechanism 34 involving bond breakage and carbon-oxygen bond formation. [53] Because of this, and the cleaning of the surface impurities, the ET process was selected as the pretreatment for the remainder of this study. It is also important to point out that this pretreatment procedure wold be accessible to the users of the sensing technology. Not all users would have access to specialty equipment to perform the hydrogen or oxygen termination procedures. Since the ET uses the same technology as the method its-self, it is accessible to the common user. This pretreatment was only performed on the electrodes before beginning a set of measurements. It was not performed between measurements for the same calibration curve. For the cyclic voltammograms of the redox couples, Figure 2.4, the MAC electrode provided the expected semi-infinite linear diusion response and the typical bell shaped curves. Peak separations, Ep, for Fe(CN)63-/4- and Ru(NH3)63+/2+ were 86 mV and 67 mV, respectively. Ep for Fe(CN)63-/4- on BDD electrodes is sensitive to the surface chemistry and cleanliness whereas Ru(NH3)63+/2+ is insensitive to the surface chemistry, inner sphere or outer-sphere electron transfer, respectively. [53–57] The heterogeneous rate constants for both of these redox couples was determined using the Nicholson and Shain method. [58, 59] The rate constants on the MAC for Fe(CN)63-/4- and Ru(NH3)62+/3+ were 1.74 +/- 0.10 x 10-3 cm/s and 1.17 +/- 0.09 x 10-2 cm/s, respectively. Both redox couples indicated semi-linear diusion due to the peak currents on the forward scans showing a linear relationship with the square root of the scan rate. The CVs for the MEA was dierent from the MAC’s in that they displayed a sigmoidal shape for both the oxidation and reduction of the redox couples. This shows that the reaction is limited by diusion of the reactants to the electrodes surface due to the lack of overlap in the molecular diusion towards each individual microelectrode causing the overall output current to be the sum of all the elements in the array. [23, 24] Due to the small size of the microelectrodes, the reactant transport is dominated by hemispherical diusion, improving the signal-to-noise ratio. [18, 21, 23] This diusion profile can shift back to semi-infinite diusion behavior if the spacing between the microelectrodes is too small. [22, 23] The CV’s for the MEA are shown in Figure 2.4 (B). The rate constants for this sensor con- 35 figuration were determined using fast-scan CV (FSCV) measurements and the Klinger and Kochi method. [60] This was due to the fact that the peak separations were above the maximum limit for the Nicholson and Shain method preciously used. The Ep value at 100 V/s were 560 mV and 127 mV for Fe(CN)63-/4- and Ru(NH3)62+/3+, respectively. The rate constants for Fe(CN)63-/4- and Ru(NH3)62+/3+ were 5.21 +/- 1.03 x 10-3 cm/s and 1.40 +/- 0.13 -1 cm/s, respectively. Both redox couples displayed a linear relationship between the forward peak current and the square root of the scan rate. Surface morphology and cleanliness is a determining factor for Fe(CN)63-/4- reaction kinetics. This is due to the fact that the reaction proceeds through surface sites which can become blocked by adsorbed oxygen. [61, 47] Hydrogen terminated films typically show rate heterogeneous rate constants between 0.01 and 0.1 cm/s compared to oxygen terminated films that can be 2 orders of magnitude lower. [61, 47, 48] The resulting rate constants for both the MAC and MEA were in the range typically displayed for oxygen terminated films, being consistent with the studies into surface termination after the ET pretreatment. For Ru(NH3)62+/3+, the rate constant is influenced by the density of electronic states near its formal potential. [57] Diamond films have shown rate constants between 0.01 and 0.1 cm/s for Ru(NH3)62+/3+.[47] Both the MAC and MEA fell within this range. 2.3.2 BDD as a Quasi-Reference Electrode Since the three electrodes making up the two sensor configurations are made completely of BDD, two questions needed to be addressed: (1) Does this BDD film have a stable potential in solutions of diering pH values? (2) Is the potential stable over an extended period of measurement time? These two questions were investigates using the MAC sensor by measuring its reference elec- trode’s potential in solutions of various pH values compared to the potential of a Ag/AgCl electrode. Figure 2.5 shows the potential-time curves at a pH of 4.0-7.4 over a 12 hour period of time. The BDD’s steady-state potential shifts in a negative direction with increasing pH. This phenomenon was reported in literature previously for diamond. [30] The potential stabilizes quickly in all the 36 tested electrolyte solutions regardless of the pH. Figure 2.5: The BDD’s reference potential in 0.1M phosphate buered saline at various pH values measured against a Ag/AgCl reference electrode for 12 hours. Ponnuswamy attributed the change in BDD’s potential in dierent pH solutions to the reactions between protons in the electrolyte solution and the surface. [38] The results found here are applicable to their suggestion that the change in proton concentration and interaction with the diamond surface causes the shift in potential. At each pH value, the potential stabilized after the first hour and remained relatively unchanged for the full 12 hour period. These results indicate that the quasi-reference material of BDD is suitable for a sensing application for a 3-in-1 device. To further evaluate the quasi-reference material in this device, a 28 hour stability experiment was performed. The average potential was -0.197 V with a standard deviation of 0.015 V as shown in Figure 2.6. Although the response fluctuated in the first 4 hours, it reached a stable potential where it remained for the remaining time. This shows that the BDD quasi reference can have a stable response for an extended measurement time. 2.3.3 Isatin Detection Due to the stability of the 0.1 M phosphate buered saline solution at pH 7.4, and its biological relevance, it was chosen to be used for the isatin voltammetric studies. Before each measurement, oxygen was removed from the solution by nitrogen (N2) bubbling. To further prevent oxygen from re-entering the environment, a N2 blanket was used on top of the solution during a measurement. 37 Figure 2.6: The BDD’s reference potential for the extended 28 hours in 0.1M phosphate buered saline at a pH of 7.4 measured against a Ag/AgCl electrode. Figure 2.7 shows the CVs for 350 µM of isatin on both the (A) MAC and (B) MEA sensor configurations at a scan rate of 0.1 V/s. Preconditioning of the sensors was done using the ET pretreatment process as previously mentioned. The CV for the MAC electrode, Figure 2.7 (A), shows the independent anodic and cathodic peaks that have been reported previously. [11] The anodic peak is located at +0.832 V; -0.2 V from the peak location reported in literature against a Ag/AgCl reference electrode. [11] This shift seen on this electrode is due to the fact that the BDD quasi-reference displays a potential that is -0.2 V when referenced against a Ag/AgCl reference electrode. This same phenomenon occurs with the cathodic peak for isatin. The formal potential and half wave potential were determined as -0.034 V and -0.686 V, respectively. The MEA CV, Figure 2.7 (B), shows a sigmoidal response for the reactions occurring for isatin. This sensor configuration also displays a lower faradaic current and background current. Since both sensor configurations clearly display the cathodic process and the cathodic process avoids interferences with other compounds such as ascorbic acid and glucose, it was chosen for the electrochemical detection comparison. [11] Before performing the detection experiments the possible adsorption of isatin and its reduction products were studied on the BDD surface using CV and the MAC sensor configuration. The CV 38 Figure 2.7: Current versus voltage cyclic voltammogram curves for 350 µM isatin in 0.1 M PBS at pH 7.4 on the 3-in-1 (A) MAC sensor and (B) MEA sensor. Scan rate of 0.1 V/s. was measured for 350 µM of isatin on the sensor before removing it from solution and rinsing it with DI water and immediately placing the sensor into a clean solution of 0.1 M PBS. N2 was bubbled before measuring the original isatin CV and the following CV in the clean PBS buer for 5 minutes. Figure 2.8 shows the results from this study. It was determined that there was no adsorption on the BDD’s surface from isatin or its reduction products even at these elevated levels. This result coincides with that described by Diculescu et al. where they found the lack of adsorption of isatin on a glassy carbon electrode. Square wave voltammetry was used to study the reduction of isatin based on the standard addition method with concentrations from 0.3 to 12 µM in 5 mL of 0.1 M PBS at a pH of 7.4. N2 bubbling was performed and blanketed on top the electrolyte solution. Both sensor configurations 39 Figure 2.8: Current versus voltage cyclic voltammogram curves of 350 µM isatin in 0.1 M PBS at pH 7.4 (red) and after rinsing the 3-in-1 MAC electrode and rescanning the same electrode in clean electrolyte solution (black). Scan rate of 0.1 V/s. showed a linear range from 0 to 12 µM of isatin and a second linear range from 20 to 100 µM. In order to determine the sensitivity, limit of detection (LOD), and limit of quantification (LOQ) of each sensor, three tests were performed. The LOD and LOQ were calculated using 3/ m and 10 /m, respectively where represents the standard deviation of the current for the lowest concentration measured and m is the slope of the obtained calibration curve. values for the MAC and MEA were 7.3 nA and 0.01 nA, respectively. After each standard addition of isatin, the solution was stirred for 10 seconds. Figure 2.9 shows the voltammograms for 0.3 to 12 µM of isatin on the MAC (A) and MEA (B) sensors. The peak heights at each concentration were measured from drawing a line at the background current level and determining the distance from the background to the peak maximum. [62, 50, 51] From Figure 2.9 two dierent peak shapes can be seen; one for the MEA and one for the MAC electrode. The shape of the peak produced by the MEA sensor is much wider with a steep rise on the leading edge of the peak and a drawn out decay peak. The MAC had a full width half max (FWHM) of 0.23 V at 12 µM whereas the MEA had a FWHM of 0.26 V. The irreversibility of the reduction of isatin is presumed to be the cause of these broad peak shapes for both sensor configurations. SWV has been shown to result in wider peaks when analyzing irreversible systems. [62] This phenomenon occurs due to the reverse pulse peak being more broad than the forward 40 Figure 2.9: Squarewave voltammograms for various concentrations of isatin on the 3-in-1 (A) MAC and (B) MEA sensor configurations in 0.1 M phosphate buered saline with pH of 7.4. pulse peak whereas in a reversible system they would be nearly mirror images of one another. [63, 64, 52] The broader peak shape for these sensors can also be a result of the charge transfer coecient (↵) being below 0.2. ↵ values this low have been shown to distort peak shapes as seen with those for the MAC configuration. [50, 51] The dierence in peak shapes could also be a result of the dierent diusive properties between the two electrode configurations. On a literature reported MAC made of glassy carbon, isatin reduction occurred in a two step reaction where the products reduced to isatide. [11] This reduction appeared as a shoulder in their detection scans. In the study presented here, the MAC sensors have a tail at the end of the stripping peak which could be related to the reduction of the products to isatide on this BDD surface. The MEA does not display this same characteristic tail suggesting the reduction products diuse away faster on this sensor configuration and do not become further 41 reduced. It has been shown that analytes diuse away from MEA’s faster than they do when using an MAC. [21–29] Figure 2.10: Calibration curves for (A) the MAC and (B) the MEA in 0.1 M phosphate buered saline with a pH of 7.4 Ultimately the MEA exhibited a lower sensitivity due to the standard deviation of the current measurement being 2 orders of magnitude lower at 0.3 µM compared to the MAC. The calibration curves for both sensor configurations are shown in Figure 2.10. The calculated LOD and LOQ were both lower for the MEA due to this. The LOD of the MEA was 0.04 µM and that of the MAC was 0.22 µM. Both the MEA and MAC sensors displayed linearity with an excellent R2 value. These values can be seen in Table 2.1. Also shown in Table 2.1 is the comparison of the electrochemical results to those obtained with UV-vis spectroscopy and other literature studies. The wavelength used for UV-Vis measurements 42 Table 2.1: Figures of merit for isatin detection using the 3-in-1 MAC and MEA sensors compared to other reported literature studies on dierent detection methods. Standard deviations calculated at lowest concentration detected for this study was from 200 to 450 nm and the absorbance of isatin was recorded for concentrations of 0.3-12 µM. The absorbance spectra for these concentrations is provided as Figure 2.11. The LOD was determined by generating a calibration curve from the peak absorbance at 242 nm. This value was higher than that obtained for the electrochemical detection mainly due to the large standard deviation of the signal between each test. In fact, this dierentiation is orders of magnitude higher than that seen with both electrochemical sensors. Due to this, both the MEA and MAC have proven to have lower detection limits than the UV-vis detection. Compared to the other electrochemical sensing study that did not involve a pretreatment between concentration measurements, the study performed here obtains a lower LOD and LOQ with a higher linear correlation coecient and linear range of concentrations. [11] The study here also advances this technology by removing the need of separate counter and reference electrodes. Further, the use of BDD can be more reproducible with the ET pretreatment compared to having to polish a glassy carbon electrode before each day of measurements. Polishing drastically changes the surface of the electrode and can cause large fluctuations in data and a higher degree of uncertainty in the actual concentrations measured. BDD is more suitable for understanding the concentrations in the human body and how the reactions are occurring. This more advanced sensing technology will allow for a quicker turn around on information related to the advancement of pharmaceuticals. 43 Figure 2.11: UV-vis absorbance spectra for isatin in 0.1 M of PBS pH 7.4 2.3.4 Urine Simulant Study To further advance the understanding of this compound using electrochemical methods, calibration experiments were performed in a urine simulant with no N2 bubbling prior to measurement for evaluation in a more complex sample matrix. Isatin was studied in the range of 1.8 - 12 µM due to the lower concentrations being hidden in the background of the new matrix. Due to the increased complexity of this study, the sensitivity of the MAC decreased by 10x with an increase in the LOD to 0.82 µM. The calibration curve still displayed linearity of 0.995. The MEA sensitivity only decreased by a factor of 3 and the LOD increased to 0.80 µM but this sensor also displayed linearity with an R2 value of 0.994. Figure 2.12 shows the voltammograms for both the MEA and MAC. The calibration curves for both sensors are shown in Figure 2.13. The calculated LODs are higher than those that were determined in the buered solutions, but most LC methods use pretreatment methods as well as extraction procedures before determining isatin concentrations that include diluting the complex matrix, pH adjustments, and using ethyl acetate for extraction. [8, 14, 65] Even with these pretreatment procedures, the LOD when using LC-MS was well above that obtained for the two electrochemical sensors at 2 µM. No studies of isatin had previously been performed in synthetic urine solution. This study allows further understanding of the capability to use electrochemistry in the measurement of real samples. 44 Figure 2.12: Squarewave voltammograms for the 3-in-1 MAC sensor (A) and MEA sensor (B) in Surine Negative Urine Control with no additional N2 bubbling 45 Figure 2.13: Calibration curves for the (A) MAC and (B) MEA for isatin in Surine Negative Urine Control with no N2 bubbling 46 2.3.5 Repeatability Study A 5-day repeatability study was performed to analyze the repeatability of the BDD sensors. The measured data is shown in Figure 2.14. The ET pretreatment procedure was used on the first day of the study for both sensors. Square wave voltammetry was then performed in a 3 µM solution of isatin in PBS buer of pH 7.4. Over the next 3 days, each sensor was used to rest a fresh solution for 7 total trials without repeating the pretreatment method. Both sensors had a decrease in signal; 63% for the MEA and 73% for the MAC. The ET process was the repeated and the sensor were retested. The MEA regained 64% of the original signal it had and the MAC regained 76%. Over the course of the 7 trails, the MEA remained stable and regained full signal strength in trial 6. The MAC showed a similar trend for those without the ET method applies but the signal did not decrease as significantly with the ET procedure compared to without. This study suggests that there are substantial changes with the signal response over the course of a few days and additional research is required to better understand how the surface regeneration works. Figure 2.14: Repeatability data for the 3-in-1 (A) MAC and (B) MEA sensors with the discussed electrochemical pretreatment and without any addition of a pretreatment 2.3.6 Key Advancements Obtained from Study The advancements obtained from this study include: 47 • Simplified technical equipment for the electrochemical detection of isatin by removing the need for three separate electrodes with a 3-in-1 sensor • Time savings in the analysis of isatin samples by implementing one sample pretreatment process compared to the previous complicated and multistep processes used on other electrode materials • The ability to directly detect isatin in a complex sample, such as a urine simulant, without the necessary extraction methods typically used for analysis with gas chromatography. • Using the MEA sensor configuration obtained a sensitivity that was two orders of magnitude lower than the MAC resulting in a lower LOD in simulated solution and as overall more repeatable with the electrochemical pretreatment method compared to the MAC. 2.4 Summary Two sensors were fabricated to have a reference, auxiliary, and working electrode, all comprised of BDD, and were fully characterized. The reference electrode had a potential variation with pH, but stabilized quickly and remained stable over the course of 30 hours. Both sensors, one comprising of a MEA and one comprising of a macro electrode, were characterized with a surface sensitive and surface insensitive redox couples and displayed comparable rate constant values. Adsorption of isatin and its reduction products were evaluated and no adsorption was found for the BDD surface. Comparison of the MEA and MAC lead to the MEA having a LOD of 0.04 µM whereas the MAC had a LOD 0.22 µM; both below the LOD of UV-vis spectroscopy measurements. Both sensors were further evaluated in a urine simulant study as well as repeatability tests. This work concludes that the fully BDD devices can be feasible for the measurement of isatin. Future applications of these sensors include the detection following a flow injection analysis system. 48 BIBLIOGRAPHY 49 BIBLIOGRAPHY [1] Haihong Xu, Dan Wang, Wen Zhang, Wei Zhu, Katsunobu Yamamoto, and Litong Jin. Deter- mination of isatin and monoamine neurotransmitters in rat brain with liquid chromatography using palladium hexacyanoferrate modified electrode. Analytica Chimica Acta, 577:207–213, 2006. [2] Lin Hou, Chuanxia Ju, Jinyu Zhang, Jinlian Song, Yinlin Ge, and Wang Yue. Antitumor eects of isatin on human neuroblastoma cell line (sh-sy5y) and the related mechanism. European journal of pharmacology, 589(1):27–31, 2008. [3] Manjusha Verma, Surendra Nath Pandeya, Krishna Nand Singh, and James P Stables. Anti- convulsant activity of schi bases of isatin derivatives. ACTA PHARMACEUTICA-ZAGREB-, 54(1):49–56, 2004. [4] Vivette Glover, JM Halket, PJ Watkins, Angela Clow, BL Goodwin, and M Sandier. Isatin: identity with the purified endogenous monoamine oxidase inhibitor tribulin. Journal of neurochemistry, 51(2):656–659, 1988. [5] P Watkins, Angela Clow, Vivette Glover, J Halket, A Przyborowska, and M Sandler. Isatin, regional distribution in rat brain and tissues. Neurochemistry international, 17(2):321–323, 1990. [6] Alexei Medvedev, Natalia Igosheva, Michele Crumeyrolle-Arias, and Vivette Glover. Isatin: Role in stress and anxiety: Review. Stress, 8(3):175–183, 2005. [7] Severino Carlos B Oliveira, Isabel PG Fernandes, Bárbara V Silva, Angelo C Pinto, and Ana Maria Oliveira-Brett. Isatin nitro-derivatives redox behaviour. Journal of Electroanalyt- ical Chemistry, 689:207–215, 2013. [8] Shigeo Manabe, Qiang Gao, Juan Yuan, Toshie Takahashi, and Akira Ueki. Determina- tion of isatin in urine and plasma by high-performance liquid chromatography. Journal of Chromatography B: Biomedical Sciences and Applications, 691(1):197–202, 1997. [9] Alexei E Medvedev, Angela Clow, Merton Sandler, and Vivette Glover. Isatin: a link between natriuretic peptides and monoamines? Biochemical pharmacology, 52(3):385–391, 1996. [10] Surendra Nath Pandeya, Sivakumar Smitha, Mayank Jyoti, and SESHAIAH KRISHNAN Sridhar. Biological activities of isatin and its derivatives. Acta Pharm, 55(1):27–46, 2005. [11] Victor C Diculescu, Sunita Kumbhat, and Ana Maria Oliveira-Brett. Electrochemical be- haviour of isatin at a glassy carbon electrode. Analytica chimica acta, 575(2):190–197, 2006. [12] Alpana K Gupta and Rajendra S Sindal. A comparative study of electrochemical reduction of isatin and its synthesized schi bases at hmde. Journal of chemical sciences, 121(3):347–351, 2009. 50 [13] Cigdem Karaaslan and Sibel Suzen. Electrochemical behavior of biologically important indole derivatives. International Journal of Electrochemistry, 2011, 2011. [14] Naoya Hamaue, Noriko Yamazaki, Masaru Minami, Toru Endo, Masahiko Hirahuji, Yoshio Monma, and Hiroko Togashi. Determination of isatin, an endogenous monoamine oxidase inhibitor, in urine and tissues of rats by hplc. General Pharmacology: The Vascular System, 30(3):387–391, 1998. [15] Yanli Zhou and Jinfang Zhi. The application of boron-doped diamond electrodes in ampero- metric biosensors. Talanta, 79(5):1189–1196, 2009. [16] Nicolae Spataru, Bulusu V Sarada, Donald A Tryk, and Akira Fujishima. Anodic voltammetry of xanthine, theophylline, theobromine and caeine at conductive diamond electrodes and its analytical application. Electroanalysis, 14(11):721, 2002. [17] Zhao Guo-Hua, Li Ming-Fang, and Li Ming-Li. Dierential pulse voltammetric determination of dopamine with the coexistence of ascorbic acid on boron-doped diamond surface. Central European Journal of Chemistry, 5(4):1114–1123, 2007. [18] Christophe Provent, Werner Haenni, Eduardo Santoli, and Philippe Rychen. Boron-doped diamond electrodes and microelectrode-arrays for the measurement of sulfate and perox- odisulfate. Electrochimica acta, 49(22):3737–3744, 2004. [19] Karolina Peckova, Jana Musilova, and Jií Barek. Boron-doped diamond film electrodes—new tool for voltammetric determination of organic substances. Critical Reviews in Analytical Chemistry, 39(3):148–172, 2009. [20] Romana Jaroöová, Paula M De Sousa Bezerra, Catherine Munson, and Greg M Swain. Assessment of heterogeneous electron-transfer rate constants for soluble redox analytes at tetrahedral amorphous carbon, boron-doped diamond, and glassy carbon electrodes. physica status solidi (a), 213(8):2087–2098, 2016. [21] Ka L Soh, Weng P Kang, Jim L Davidson, S Basu, Yong M Wong, David E Cliel, AB Bonds, and GM Swain. Diamond-derived microelectrodes array for electrochemical analysis. Dia- mond and Related Materials, 13(11):2009–2015, 2004. [22] Cory A Rusinek, Adam Bange, Mercedes Warren, Wenjing Kang, Keaton Nahan, Ian Pa- pautsky, and William R Heineman. Bare and polymer-coated indium tin oxide as working electrodes for manganese cathodic stripping voltammetry. Analytical chemistry, 88(8):4221– 4228, 2016. [23] Kentaro Tsunozaki, Yasuaki Einaga, Tata N Rao, and Akira Fujishima. Fabrication and elec- trochemical characterization of boron-doped diamond microdisc array electrodes. Chemistry letters, 31(5):502–503, 2002. [24] Nathan S Lawrence, Markus Pagels, Andrew Meredith, Timothy GJ Jones, Clive E Hall, CS Jim Pickles, Herman P Godfried, Craig E Banks, Richard G Compton, and Li Jiang. Elec- troanalytical applications of boron-doped diamond microelectrode arrays. Talanta, 69(4):829– 834, 2006. 51 [25] JL Davidson, DR Wur, WP Kang, DL Kinser, and DV Kerns. Polycrystalline diamond pressure microsensor. Diamond and related Materials, 5(1):86–92, 1996. [26] P Bergonzo, A Bongrain, E Scorsone, A Bendali, L Rousseau, G Lissorgues, P Mailley, Y Li, T Kaumann, and F Goy. 3d shaped mechanically flexible diamond microelectrode arrays for eye implant applications: The medinas project. Irbm, 32(2):91–94, 2011. [27] M Bonnauron, S Saada, L Rousseau, G Lissorgues, C Mer, and P Bergonzo. High aspect ratio diamond microelectrode array for neuronal activity measurements. Diamond and Related Materials, 17(7):1399–1404, 2008. [28] Gaëlle Piret, Clément Hébert, Jean-Paul Mazellier, Lionel Rousseau, Emmanuel Scorsone, Myline Cottance, Gaelle Lissorgues, Marc O Heuschkel, Serge Picaud, and Philippe Bergonzo. 3d-nanostructured boron-doped diamond for microelectrode array neural interfacing. Bioma- terials, 53:173–183, 2015. [29] Agnieszka Kapa≥ka, György Fóti, and Christos Comninellis. The importance of electrode material in environmental electrochemistry: formation and reactivity of free hydroxyl radicals on boron-doped diamond electrodes. Electrochimica Acta, 54(7):2018–2023, 2009. [30] Waldemar Smirnov, Nianjun Yang, René Homann, Jakob Hees, Harald Obloh, Wolfgang Muller-Sebert, and Christoph E Nebel. Integrated all-diamond ultramicroelectrode arrays: optimization of faradaic and capacitive currents. Analytical chemistry, 83(19):7438–7443, 2011. [31] Wenjing Kang, Xing Pei, Adam Bange, Erin N Haynes, William R Heineman, and Ian Pa- pautsky. Copper-based electrochemical sensor with palladium electrode for cathodic stripping voltammetry of manganese. Analytical chemistry, 86(24):12070–12077, 2014. [32] Wenjing Kang, Xing Pei, Cory A Rusinek, Adam Bange, Erin N Haynes, William R Heine- man, and Ian Papautsky. Determination of lead with a copper-based electrochemical sensor. Analytical Chemistry, 89(6):3345–3352, 2017. [33] Wenjing Kang, Cory Rusinek, Adam Bange, Erin Haynes, William R Heineman, and Ian Pa- pautsky. Determination of manganese by cathodic stripping voltammetry on a microfabricated platinum thin–film electrode. Electroanalysis, 29(3):686–695, 2017. [34] Gustav Liljegren, Jean Pettersson, Karin E Markides, and Leif Nyholm. Electrochemical solid-phase microextraction of anions and cations using polypyrrole coatings and an integrated three-electrode device. Analyst, 127(5):591–597, 2002. [35] Markus Pagels, Clive E Hall, Nathan S Lawrence, Andrew Meredith, Timothy GJ Jones, Herman P Godfried, CS James Pickles, Jonathan Wilman, Craig E Banks, and Richard G Compton. All-diamond microelectrode array device. Analytical chemistry, 77(11):3705– 3708, 2005. [36] Xing Pei, Wenjing Kang, Wei Yue, Adam Bange, William R Heineman, and Ian Papautsky. Disposable copper-based electrochemical sensor for anodic stripping voltammetry. Analytical chemistry, 86(10):4893–4900, 2014. 52 [37] Miroslav Pohanka and Petr Skládal. Electrochemical biosensors–principles and applications. J. Appl. Biomed, 6(2):57–64, 2008. [38] Thomas Ponnuswamy, Jin-Jian Chen, Fei Xu, and Oliver Chyan. Monitoring metal ion contamination onset in hydrofluoric acid using silicon–diamond and dual silicon sensing electrode assembly. Analyst, 126(6):877–880, 2001. [39] Xinhao Wang, Manas Ranjan Gartia, Jing Jiang, Te-Wei Chang, Junle Qian, Yong Liu, Xian- grong Liu, and Gang Logan Liu. Audio jack based miniaturized mobile phone electrochemical sensing platform. Sensors and Actuators B: Chemical, 209:677–685, 2015. [40] Youngsam Yoon, Gil S Lee, Koangki Yoo, and Jeong-Bong Lee. Fabrication of a micronee- dle/cnt hierarchical micro/nano surface electrochemical sensor and its in-vitro glucose sensing characterization. Sensors, 13(12):16672–16681, 2013. [41] R Ramamurti, M Becker, T Schuelke, T Grotjohn, D Reinhard, G Swain, and J Asmussen. Boron doped diamond deposited by microwave plasma-assisted cvd at low and high pressures. Diamond and Related Materials, 17(4):481–485, 2008. [42] E Bustarret, E Gheeraert, and K Watanabe. Optical and electronic properties of heavily boron-doped homo-epitaxial diamond. physica status solidi (a), 199(1):9–18, 2003. [43] HJ Xiang, Zhenyu Li, Jinlong Yang, JG Hou, and Qingshi Zhu. Electron-phonon coupling in a boron-doped diamond superconductor. Physical Review B, 70(21):212504, 2004. [44] T Teraji, H Wada, M Yamamoto, K Arima, and T Ito. Highly ecient doping of boron into high-quality homoepitaxial diamond films. Diamond and Related Materials, 15(4):602–606, 2006. [45] Dario Narducci, C Richard Guarnieri, and Jerome J Cuomo. Defect clustering and boron elec- trical deactivation in p-doped polycrystalline diamond films. Journal of The Electrochemical Society, 138(8):2446–2451, 1991. [46] P Gonon, E Gheeraert, A Deneuville, F Fontaine, L Abello, and G Lucazeau. Characterization of heavily b-doped polycrystalline diamond films using raman spectroscopy and electron spin resonance. Journal of applied physics, 78(12):7059–7062, 1995. [47] Joel W Ager III, W Walukiewicz, Matthew McCluskey, Mary Anne Plano, and Maurice I Landstrass. Fano interference of the raman phonon in heavily boron-doped diamond films grown by chemical vapor deposition. Applied Physics Letters, 66(5):616–618, 1995. [48] E Gheeraert, P Gonon, A Deneuville, L Abello, and G Lucazeau. Eect of boron incorporation on the “quality” of mpcvd diamond films. Diamond and Related Materials, 2(5-7):742–745, 1993. [49] NG Ferreira, AF Azevedo, AF Beloto, M Amaral, FA Almeida, FJ Oliveira, and RF Silva. Nanodiamond films growth on porous silicon substrates for electrochemical applications. Diamond and related materials, 14(3):441–445, 2005. 53 [50] Bard Allan J. and Faulkner Larry R. Electrochemical Methods Fundamentals and Applica- tions. John Wiley and sons, Inc., United States of America, second edition, 2011. [51] Strobel Howard A. and Heineman William R. Chemical Instrumentation: A Systematic Approach. John Wiley and Sons, Inc., United States of America, third edition, 1989. [52] Julie V Macpherson. A practical guide to using boron doped diamond in electrochemical research. Physical Chemistry Chemical Physics, 17(5):2935–2949, 2015. [53] Michael C Granger, Malgorzata Witek, Jishou Xu, Jian Wang, Mateusz Hupert, Amy Hanks, Miles D Koppang, James E Butler, Guy Lucazeau, and Michel Mermoux. Standard electro- chemical behavior of high-quality, boron-doped polycrystalline diamond thin-film electrodes. Analytical Chemistry, 72(16):3793–3804, 2000. [54] Anne E Fischer, Yoshiyuki Show, and Greg M Swain. Electrochemical performance of diamond thin-film electrodes from dierent commercial sources. Analytical chemistry, 76(9):2553–2560, 2004. [55] Yoshiyuki Show, Ma≥gorzata A Witek, Prerna Sonthalia, and Greg M Swain. Characteriza- tion and electrochemical responsiveness of boron-doped nanocrystalline diamond thin-film electrodes. Chemistry of materials, 15(4):879–888, 2003. [56] Matt Hupert, Alexander Muck, Jian Wang, Jason Stotter, Zuzana Cvackova, Shannon Hay- mond, Yoshiyuki Show, and Greg M Swain. Conductive diamond thin-films in electrochem- istry. Diamond and Related Materials, 12(10):1940–1949, 2003. [57] Michael C Granger and Greg M Swain. The influence of surface interactions on the re- versibility of ferri/ferrocyanide at boron-doped diamond thin-film electrodes. Journal of the Electrochemical Society, 146(12):4551–4558, 1999. [58] Richard S Nicholson. Theory and application of cyclic voltammetry for measurement of electrode reaction kinetics. Analytical Chemistry, 37(11):1351–1355, 1965. [59] Joseph Wang. Analytical Electrochemistry. John Wiley and Sons, Inc, Online, 3 rd edition, 2006. [60] RJ Klingler and JK Kochi. Electron-transfer kinetics from cyclic voltammetry. quantitative description of electrochemical reversibility. The Journal of Physical Chemistry, 85(12):1731– 1741, 1981. [61] Matthias Werner, Otto Dorsch, H-U Baerwind, Emst Obermeier, Colin Johnston, Paul R Chalker, and Simon Romani. The eect of metallization on the ohmic contact resistivity to heavily b-doped polycrystalline diamond films. IEEE Transactions on Electron Devices, 42(7):1344–1351, 1995. [62] Kissinger Peter T. and Heineman William R. Laboratory Techniques in Electroanalytical Chemistry. Marcel Dekker, Inc., New York, second, revised and expanded edition, 1996. 54 [63] Michael J Nuwer, John J O’Dea, and Janet Osteryoung. Analytical and kinetic investigations of totally irreversible electron transfer reactions by square-wave voltammetry. Analytica chimica acta, 251(1-2):13–25, 1991. [64] John J O’Dea, Janet Osteryoung, and Robert A Osteryoung. Theory of square wave voltam- metry for kinetic systems. Analytical Chemistry, 53(4):695–701, 1981. [65] JM Halket, PJ Watkins, A Przyborowska, BL Goodwin, Angela Clow, Vivette Glover, and M Sandler. Isatin (indole-2, 3-dione) in urine and tissues: Detection and determination by gas chromatography—mass spectrometry. Journal of Chromatography B: Biomedical Sciences and Applications, 562(1-2):279–287, 1991. 55 CHAPTER 3 INDIUM TIN OXIDE FILM CHARACTERISTICS FOR CATHODIC STRIPPING VOLTAMMETRY 56 3.1 Introduction This work has been published in ACS Applied Materials and Interfaces1. Indium tin oxide, or ITO, is a common conductive and optically transparent film that is used in applications such as transparent electrodes for solar cells, surface heaters for windows, liquid-crystal displays, and spectroelectrochemical applications. [1–10] Although sputtering is the most common deposition method for ITO films, it can also be performed using chemical vapor deposition, ion beam evapo- ration, and spray pyrolysis. [1, 2, 4–7, 11–17] ITO films can be tailored to having dierent optical and electrical properties and can be deposited as amorphous or crystalline films. The wide range of characteristics that each film can possess comes from the changes in the ordered structure of the film due to the generation of oxygen vacancies. [7] The tunable process parameters that allow these changes include gas pressures, substrate temperatures, deposition time, film thickness, and post processing annealing steps in dierent gases and at dierent temperatures. [2, 4, 5, 7, 11, 18, 19] Between the temperatures of 100 and 150 C, there is an amorphous to crystalline transition for ITO. [19] This shift causes films to dier in their presence of oxygen and conducting carrier concentration. [1, 7, 19, 20] When the substrate temperature is above 100 C the film’s resistivity drops leading to an increase in carrier concentration and mobility. [6, 19] A film thickness of greater than 200 nm, deposited at room temperature, are typically crystalline, but those below 100 nm are typically amorphous. [4] All of these process parameters can be tuned to create unique films for specific applications. [2, 4, 7, 11, 19, 21] Post deposition process parameters can also have an eect on the ITO film’s characteristics. dc-sputtered post deposition annealed films show an increase in grain size, decrease in electrical resistivity, and a decrease in intrinsic strain. [4, 5, 11, 20] The temperature needed for crystallization during annealing is higher for films that were deposited on non-heated substates compared to those [4, 7] The optical properties of the film are also aected by deposited on heated substrates. annealing. [4] Various It can lead to an increased band gap and a steeper cuto absorption. 1ACS Applied Materials and Interfaces 2019, DOI:10.1021/acsami.8b22157 57 studies have been carried out to study the changes in the ITO film based on annealing atmosphere, temperature, and if vacuum is used. [1, 7, 11, 20, 22] The most common use of ITO in the field of electrochemistry is as an optically transparent film for spectroelectrochemical measurements. [9, 23, 24] It has been used to detect analytes including Azure A, total alkaline phosphate in human blood, potassium ferricyanicde, and iron using complexation with 2,2’-bipyridine. [2, 23, 25, 26] This application is primarily due to this ITO film possessing a large positive potential window beyond +1.5 V while remaining transparent in the range of 390-700 nm. [23, 27, 28] Since the film is already in its oxidized form, it does not react in this positive potential range. [25] Several metals including cadmium (Cd), lead (Pb), copper (Cu), cerium (Ce), mercury (Hg), and manganese (Mn), has also been detected with ITO, by using the method of stripping voltammetry. [27–30] Cd and Cu had their signal monitored as they were plated onto the electrode surface using attenuated total internal reflectance stripping voltammetry. [29] Ce additionally has been detected using cathodic stripping voltammetry (CSV) and ITO. [27] Recently, bare ITO was used to detect Pb, Cu, and Hg simultaneously using this method. [30] Mn has been investigated on ITO also using CSV. [28] This was performed using an ITO coated with polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene-sulfonate (SSEBS) as well as compared to a bare ITO electrode. [28] This study was followed up with a paper reporting the detection of Mn in digested bovine blood samples that obtained similar detection capabilities as inductively coupled plasma mass spectrometry (ICP-MS) and graphite furnace atomic absorption spectroscopy (GF-AAS). [31] The detection of Mn is important due to it having toxic eects and a maximum contaminant level (MCL) in drinking water of 50 parts per billion (ppb) set by the United States Environmental Protection Agency (EPA). It is also significant to test human blood, urine, and serum levels when monitoring its eects. [32, 33] Palladium (Pd) on Cu, boron-doped diamond (BDD), and various other carbon materials have also been used to detect Mn with CSV. [34–39] Compared to these other materials, the positive potential window of ITO makes it more beneficial for CSV. Cu-based Pd electrodes obtained a 58 LOD at 19 ppb which is below the MCL in drinking water but higher than the typically biological concentrations. [32, 33, 35] The BDD sensor was evaluated in sea water and required the use of ultrasound, but only obtained a LOD of 5.5 ppb which is still not sensitive enough for serum analysis. [36] Additionally BDD had a higher resistivity compared to ITO and can be expensive to fabricate into sensor designs. [40] The bare carbon paste electrode previously studied obtained the same LOD as the BDD sensor. [39] The most promising electrode material for Mn detecting using CSV was edge plane pyrolytic graphite that obtained a LOD of 0.80 ppb. [37] However, the previously studied ITO films reached an even lower LOD compared to this material which may be due to the metal oxide surface of ITO being beneficial in the deposition of manganese dioxide (MnO2) on the surface. Sensors being used for metal detection ultimately do not require the use of optical transparency when utilizing the method of CSV. ITO has been deposited on opaque substrates before, but never has then been applied to CSV metal detection. [27, 41] Fabrication of such a sensor on silicon (Si) is much safer and easier than the process used for sensor fabrication on transparent substrates due to avoiding the use of hydrofluoric acid (HF). [42] Si can utilize dry etching processes for fabrication allowing for faster and more precise etching without putting fabricators lives in jeopardy. ITO is already mass produced for other applications on non-transparent substrates that would also allow for an economical mass production of sensors. In addition to the unique electrochemical properties that ITO possess, it is already widely used in mass production for solar cells which means it is available at much lower costs, and is more reproducible than other electrode materials used in CSV. This study reports on the optimization of ITO working electrodes on Si substrates for the application of metal detection using CSV. At this time, there is no report in literature using opaque substrates for this technique. Several processing parameters such as heated and non-heated deposition and the eects of post deposition annealing were investigated. Each corresponding film was thoroughly characterized and evaluated before using each to detect Mn by CSV. 59 3.2 Experimental Section 3.2.1 Chemicals and Materials A stock solution of potassium ferrocyanide (K4Fe(CN)6) was prepared using powder purchased from J.T Baker and deionized (DI) water. Manganese (Mn) was purchased from Acros Organics. Deionized (DI) water and acetic acid (Fisher Scientific) were used to prepare a 0.2 M acetate buer with 0.2M sodium acetate purchased from Sigma-Aldrich. Ferrocene (Aldrich Chemical Company Inc.) was prepared in acetonitrile (Sigma-Aldrich) using a 1mM solution with 0.1 M tetrabutylammonium hexafluorophosphate. 3.2.2 Instrumentation An AFM5100 from Hitachi was used in tapping mode to complete the atomic force microscopy (AFM) measurements on an area of 4 µm2 with a scan frequency of 0.84 Hz. The roughness of each sample was determined using the open source software of Gwyddion using three evaluation lines that were placed across the top, middle, and bottom of the scanned area. A cuto length of one-fifth of the scan length was used in these calculations. X-Ray diraction (XRD) measurements were performed with a Rigaku SmartLab 3 kW Cu source generating a parallel beam with electron current of 44 mA and an acceleration voltage of 40 kV. Out of frame diraction peaks were collected at 4 /min using symmetric (2✓ - !) scanning stepping at 0.02 over a range of 30 - 90. Diracted X-rays were collected through collimating slits at 0.5 pitch and aperture of 20 mm using a 0D SC-70 photomultiplier tube. An area of 10 x 12 mm was set to limit the X-ray beam spot on the sample. A Shimadzu UV-vis-NIR spectrophotometer with a wavelength of 190-2500 nm was used for all optical measurements. A 414513 semiconductor parameter analyzed from Hewlett-Packard was used for the four-point probe measurements with a correction factor of 4.3. [43] An Epsilon Eclipse from Bioanalytical Systems, Inc. (BASi) was used in a three electrode setup for all electrochemical measurements. Each film was used as a working electrode with a 0.5 60 mm circle diameter exposed to the electrolyte solution. Inside the cell was a silver/silver chloride (Ag/AgCl) reference electrode and either a platinum (Pt) or graphite auxiliary electrode. A silver quasi-reference was used for the chronoamperometry (CA) experiments in ferrocene. Each CV experiment utilized 0.1 mM Fe(CN)6 4- in 1.0 M KCl by scanning the potential from -0.1 to 0.6 V. Peak height measurements were performed with an extrapolated baseline method described in literature. [44–46] 3.2.3 Electrode Fabrication Pulsed dc magnetron sputtering was utilized for all fabrication runs. Each run was performed for 20 minutes using a Kurt J. Lesker PVD 775 system. One set was performed at room temperature and one utilized a set substrate temperature of 190 C. An ITO target (In90/Sn10) with a 1.6% oxygen (O2) to argon (Ar) ratio was used. The power density was 20 W/in. Two groups of microscope slides and nonconductive Si substrates were formed: those annealed and not annealed. The annealing was done at 400 C for 1 hour in ambient air. Four final samples were evaluated; heated and annealed (HA), heated and not annealed (HNA), not heated and annealed (NHA), and not heated and not annealed (NHNA). All film’s thicknesses were in the range of 230 and 250 nm. 3.3 Results and Discussion 3.3.1 Physical and Optical Characterization: Atomic Force Microscopy Atomic force microscopy (AFM) was used to evaluate each electrode film’s surface morphology. Figure 3.1 shows the two dimensional topography views for HA, HNA, NHA, and NHNA on nonconductive silicon (Si) substrates. From this Figure, visual similarities can be seen between the two heated samples and the two non-heated samples. NHA and NHNA, shown in (C) and (D) respectively, are comprised of smaller grain sizes that appear to be more inhomogeneous than the two heated samples. Also between these two films, the annealing is noted to cause a significant appearance of taller gains shown in the lighter color. HA and HNA, shown in (A) and (B) respectively, have larger grain sizes compared to the non-heated 61 Figure 3.1: Tapping mode atomic force microscopy (AFM) images for the four analyzed ITO films (A) HA, (B) HNA, (C) NHA, and (D) NHNA. samples. Annealing the heated sample (A) also brings about higher grain structures compared to the not annealed heated sample. The AFM data also provided the root-mean-square (rms) roughness values for each film. NHNA had the lowest rms value with HA having the highest value. All the film’s rms values are provided in Table 3.1. The roughness values are aected more by the heating than the annealing, but all the films are smooth overall. There is a benefit to having a greater roughness to electrode materials. A higher roughness leads to a greater microscopic surface area which can lead to a greater sensitivity and lower limit of detection (LOD) in CSV measurements.[44–46] From this physical and optical characterization, HA and HNA are potentially more advantageous as a sensing material than the other two films. 62 Table 3.1: Comparison of each of the figures of merit for each of the four ITO films that were investigated. 3.3.2 Physical and Optical Characterization: XRD All four films were confirmed as polycrystalline from the presence of multiple peaks in the dirac- tion pattern performed with symmetric scans. The peaks were identified as Ia3 ITO using the crystallography open data base. [47–50] In Figure 3.2 (A) the (222) peak decreases as a function of heat treatment, regardless if performed by deposition or annealing. The (400) peak, Figure 3.2 (B), shows an opposite trend by increasing with heat treatment. Literature states that the (400) peak decreases when there is increasing oxygen in the ITO film suggesting that the heated deposition promotes the (400) orientation of the film. [50] The center of the (222) orientation shifts during both heating during deposition, and annealing a non heated film displaying a change in the strain of this phase. Using the NHNA film as a zero point, the NHA peak center is +0.16 °and the HNA film’s is -0.2 °. HA’s center of the (222) orientation peak is in roughly the same position as NHNA’s. The ratio of the (222) and (400) peaks, I(222)/I(400), for NHNA is 23 times greater than HNA’s, indicating a transition to a preferred orientation (Figure 3.3 (left)). The full width half-maximum (fwhm) of the (222) and (400) peaks is also a function of heat treatment. In general, the fwhm decreases from both heating and annealing as shown in Figure 3.3 (right). This trend shows that either during heating or after annealing, the crystalline quality improves. Without the application of heat during processing, the mobility of the tin and indium-oxygen clusters is restricted and limits 63 Figure 3.2: XRD scans for the (A) (222) peak and (B) (400) peak for each ITO film investigated. the likelihood of miss-orientations leading to bond distortions. Figure 3.3: (left) ITO crystal quality (I(222)/I(400)) ratio obtained from XRD analysis, and (right) the full width half max (FWHM) data for each film from XRD analysis pertaining to either the (222) or (400) peak The FWHM for NHNA is caused by this process. The application of heat above 150 °C during deposition enables the mobility of the tin and indium oxide clusters. [51] An exception is the jump in FWHM from the (400) in NHNA and the (400) in NHA and the reverse FWHM positions of (222) and (400) for HNA. This may be the result of having frozen clusters when using room temperature deposition for NHA. Annealing has shown the ability to unfreeze these sites but the diusion through the film is limited and restricts their chance of occupying favorable lattice positions. However, with heated deposition HNA allows the clusters a higher mobility and increases the probability of stable bond formations and a lower FWHM. 64 Figure 3.4: ITO crystal size derived using the Scherrer equation and the (222) and (400) peak of each film HA supports all the conclusions drawn when combining heating and annealing by having the smallest FWHM, least amount of stress or strain, and the largest (400) crystallite size as seen in Figure 3.4. This leads to the conclusion that HA could be the best film for Mn detection due to its larger grain size and increased conductivity. [19] This could also suggest that HNA may be the optimum film due to its mirrored properties for the FWHM and crystal size compared to HA. Additionally, The I(222)/I(400) of HNA is lower than that of HA. 3.3.3 Physical and Optical Characterization: UV-Vis Spectroscopy To gain insight into the charge carriers of the films, microscope slides were included in each run and analyzed using UV-Vis Spectroscopy. The Transmittance (T%) spectra of a blank microscope slide, for reference, and each of the four films is shown in Figure 3.5 The T% spectrums of HNA and NHNA are similar to one another as well as HA and NHA, suggesting that the annealing step changes the optical properties of the films. Literature states that annealing of ITO films shifts the absorption edge to a lower wavelength, and this is the same trend seen with the films in this study. [4] HA and NHA are shifted to lower wavelengths and HNA is also seen shifted in this direction 65 which may be due to the heating of the substrate. Compared to NHNA, it is presumed that these three films will have a lower defect density and higher crystallinity due to this shift in their T% spectra. [4] Figure 3.5: Transmittance spectra measured with UV/Vis for each of the four films deposited on a transparent microscope slide In the wavelength range of 1500-2500 nm, NHNA has the highest T% value with HA displaying the lowest. Due to the fact that literature has shown that a higher carrier concentration displays a lower T% in this wavelength, it is assumed that HA has the highest carrier concentration. [4, 52] The direct optical band gap of each film was also determined from the T% data using the reflectance spectrum as well. The direct optical band gap provides information about the crystallinity of each film due to the relation between an increase in band gap with a decrease in localized state density. [53, 54] In order to calculate the band gap, the energy gap (Eg) was first determined using Equation 2.1. [55, 56] ↵ ⇤ h ⇤ v = (h ⇤ v Eg)r (3.1) Where hv is the photon energy, ↵ is the absorption coecient, and r is the transition coecient with a value of either 0.5 or 2 for a direct band gap or indirect band gap, respectively. ↵ is determined using Equation 2.2. 66 T = (1 R)e(-↵d) (3.2) Where d is the film’s thickness. By extrapolating the linear region of the Tauc Plot, Figure 3.6, the direct optical band gap was determined for each film. All of the fitting parameters are displayed in Table 3.2. Figure 3.6: Tauc graph used to determine the direct optical band gap for the four ITO films studied Table 3.2: Fitting parameters for the band gap determination from the Tauc plot of each film. The lowest and highest optical band gap belonged to NHNA and HA, respectively. The band gap was increased by both heating and annealing. According to the density of states model by Mott and Davis, the process of annealing or heating removes the defects that were present in the amorphous film and forms saturated bonds that increase the band gap. [56] This trend agrees with the FWHM trend from the XRD analysis. 67 Since HA has the highest band gap followed by NHA and HNA, these three films are highly likely to be applicable for a sensing application. All the optical band gaps determined for these films compare well with literature values for other reported ITO films. [5, 7, 21] 3.3.4 Electrical Properties In terms of having a good electrode material for electrochemical sensing applications, low resistance is necessary. For each film, the resistance and resistivity values are summarized in Table 3.1. NHNA had the highest resistivity and sheet resistance and HA had the lowest. NHA and HNA had almost the same sheet resistance and resistivity as one another. According to literature the resistivity of an ITO film decreases when it is heated or annealed to temperatures above 100 C. [4, 7, 11, 19] Compared to other electrodes used for Mn detection, the ITO films have a higher resistivity than carbon, but lower than boron-doped diamond. [40, 38, 57] The results for the electrical properties suggest that HA should be expected to perform as the best electrochemical sensor; agreeing with the optical property analysis performed earlier. 3.3.5 Electrochemical Characterization Cyclic voltammograms (CVs) and Ohm’s resistance plots were generated to investigate each film’s electrochemical behavior. The CVs allowed for the double layer capacitance of each film to be determined in 1 M potassium chloride (KCl) with a scan rate range of 0.1 to 0.4 V/s. Each current value at 0.0V was recorded and was graphed against its corresponding scan rate, Figure 3.7. The slope of this graph provided the capacitance of each film that could then be divided by the exposed area to obtain the typical units of F/cm2. The inner sphere electron transfer redox couple of potassium ferrocyanide was used to evaluate each ITO film. This was chosen since it has a surface sensitive nature and can be used to compare to other electrode materials although it is believed that the reaction of this redox on the surface varies with its monolayer adsorption and does not vary with surface oxide coverage. [58] Figure 3.8(A) shows the CV for each film obtained from a scan rate of 0.1 V/s. 68 Figure 3.7: Graph showing the current vs scan rate study of each of the ITO films used to determine their corresponding capacitance values The peak separations determined at 0.1 V/s were 90 mV, 83 mV, 89 mV, and 95 mV for HA, HNA, NHA, and NHNA, respectively compared to previously published at 160 mV. [59] Potassium ferrocyanide is a common redox couple used to characterize boron doped diamond (BDD) and glassy carbon. The peak separations on these materials are 70-85 mV for BDD and 61-65 mV for glassy carbon. [59–61] Depending on pretreatments, carbon electrode materials can have drastically dierent peak separations including values of closer to 145 mV. [59, 61] The ITO films investigated here are similar to the peak separations for BDD. Figure 3.8: (A) CV curves of 0.1 mM ferrocyanide in 1.0M KCl with a scan rate of 0.1 V/s. (B) Potential windoe study performed in acetate buer of pH 5.0. (C) CSV od 25 ppb of Mn in acetate buer of pH 5.0, deposition time of 3 minutes and a deposition potential of +1.2 V. 69 The rate constant of ferrocyanide on each ITO film was determined using the Nicholson Shain method for electrodes that display low peak separation with exact psi values that were interpolated from the supplied table in the reference. [62] In terms of peak current, Figure 3.8(A), HA had the highest followed by HNA and NHA with very similar responses. NHNA displayed the lowest current value out of the four films. This trend of peak current matches the trend for the conductivity of each film. The rate constants for each film are shown in Table 3.1. HA and HNA had the fastest rate constant and NHNA has the slowest. This result suggests, again, that HA or HNA would make a great electrochemical sensor since a high current response and a fast rate constant are desired. Each ITO film’s electrochemical potential window was also studied; Figure 3.8 (B). The potential began at 0.0V and was scanned anodically to +2.5 V before being reversed and scanned to -1.75 V and then back to 0.0V. All four films display an excellent anodic potential window, but have several peaks in the cathodic scan. These peaks are believed to be from the reduction of oxygen and ITO to metallic indium (In) and tin (Sn) as studied by Liu et al. [63] The peak at -0.3 V is the largest on HA and is likely due to the reduction of oxygen in each film. This would also support the claim that the reduction of the (400) XRD peak is due to increasing oxygen in the films. According to all four scans, it appears that both heating and annealing ITO can cause an increase of oxygen incorporation into the films. The other peaks at around -1.2 V are also largest in HA and are related to the reduction of the ITO films. [63] This film with the In- and Sn- terminated phases and larger crystallite size has an important roll for the electrode performance. Plotting the change in peak potential (Ep) versus the change in peak current (ip) for each film allowed for the generation of Ohm’s resistance plots. The slope of these plots is the resistance in the electrochemical cell in Ohm. All four film’s resistance plots are shown in Figure 3.9. These plots were generated since the resistance values from the four-point-probe only provide information of the resistance through the film between the probes in air. Whereas these plots measure the resistance when exposed to an electrolyte solution as the films would be in a practical application. Each of the four film’s has a constant resistance when the temperature is constant and are ohmic resistors. HA had the lowest resistance calculated and NHNA had the highest. NHA and HNA were 70 Figure 3.9: Plots for the Ohm’s resistance of each of the four ITO films investigated very similar to each other once again. All the values obtained from these plots is also displayed in Table 3.1. The trends seen from the Ohm’s plots are expected based on the four-point-probe measurements. Chronoamperometry measurements were performed to determine the microscopic surface area [44] All calculations of each film by using an applied potential step and the Cottrell equation. were performed with a constant diusion coecient of 2.24 x 10-5 cm/s. all four films displayed microscopic surface areas that were statistically similar at roughly 0.20 cm2. The geometric surface area of 0.196 cm2 is almost identical to this value; probably due to the smoothness of each film provided in Table 3.1. From this, it was determined that the surface roughness may not be a critical parameter in discriminating between the films. 3.3.6 CSV of Manganese Manganese was used to further test the applicability of each ITO film by doing CSV with 25 ppb of Mn with a deposition potential of +1.2V for 3 min, based on previous Mn experience when using ITO. [28] The stripping voltammograms of each film were performed a total of three times and an average for each is plotted in Figure 3.8 (C). NHNA is the only film that did not have a response peak at this Mn concentration. HNA and HA displayed a comparable peak to one another 71 and NHA was visually comparable to them as well, but showed a smaller magnitude of current. Table 3.1 displays the peak current responses for each film. These results, and previous, support the hypothesis that the heated deposition run makes the most appropriate ITO film for Mn analysis, but further investigation into HA and HNA still needs to be performed to decide between them. 3.3.7 Calibration and Figures of Merit Figure 3.10: Mn stripping voltammograms for the full concentration range on HA (A1) and the smaller range (A2), for the full range on HNA (B1) and the smaller range (B2), and for the full range on NHA (C1) and smaller range (C2). Performed in pH 5.0 acetate buer with a deposition potential of 1.2 V and a deposition time of 3 min Mn concentrations of 1, 5, 10, 25, and 50 ppb were used to generate calibration curves performed in triplicate. Figure 3.10 shows the stripping voltammograms for the HA, NHA, and HNA films. Calibration curves pertaining to the full concentration range investigated are sown in Figure 3.11. The uncertainty in the slope, y-intercept, and the uncertainty in the y and correlation coecient (R2) values were determined using a Linest function in Excel. The equation for propagation of uncertainty in the best fit line was used to determine the uncertainty in the LOD for each calibration curve. [64] NHNA was excluded from both Figure 3.10 and Figure 3.11 due to its inability to perform for this application resulting from an R2 value of 0.663. 72 Figure 3.11: Calibration curves for Mn on HA (A), HNA (B), and NHA (C) The overall average response of NHA was [I(A) = (-3.3 x 10-8 +/- 3.5 x 10-9)c (A/ppb) + (1.2 x 10-7 +/- 8.9 x 10-8) (A) R2 = 0.967] with a LOD of 0.8 ppb. The downsides to this film were that the background shifted slightly throughout the measurements and the lowest concentration of 1 ppb was hard to distinguish from the background, as seen in Figure 3.10(C2). HNA had a response of [I(A) = (-3.0 x 10-8 +/- 3.4 x 10-9)c (A/ppb) +(1.4 x 10-7 +/- 8.6 x 10-8) (A) R2 = 0.964] and a LOD of 0.3 ppb. Compared to the peaks obtained in NHA, the peaks on HNA are more symmetric and the background level shifts less. The lowest concentration is still hard to distinguish from the background in Figure 3.10(B2). HNA also exhibited an improved reproducibility and lower 73 standard deviation at 1 ppb compared to NHA. HA had an overall response with the lowest LOD, highest sensitivity, and greatest correlation coecient; [I(A) = (-3.4 x 10-8 +/- 1.2 x 10-9)c (A/ppb) + (8.2 x 10-8 +/- 3.1 x 10-8) (A) R2 =0.996]. The LOD on HA was 0.1 ppb which is 3x lower than that of HNA. The R2 value for this film was the only analytically acceptable value. Previously recorded LODs for Mn on bare ITO are reported at 0.27 ppb, making the film studied here superior. [28] The lowest concentration of 1 ppb on this film displays a clear and distinct peak compared to HNA and NHA. The background also remains the most stable and appears the flattest from the other films. HA was also the most reproducible film overall. The LODs obtained on HA, HNA, and NHA are lower than the MCL of 50 ppb for Mn in drinking water, blood (4-15 ppb), urine (1-8 ppb), and serum (0.4-0.85 ppb). [32, 33] It is also worth noting that the films investigated here did not undergo a deposition time optimization which can lead to even lower LOD values if necessary. 3.3.8 Possible Metals Interference Since manganese is not the only metal that is present in environmental and biological samples, it is important to consider if other metals may have an aect on the sensors Mn detection response. These metals can include cadmium, chromium, copper, iron, lead, and nickel. Table 3.3 shows each of these metals reaction potentials. By using cathodic stripping voltammetry and the application of a positive deposition potential, the interference from most metals plating onto the working electrode’s surface is avoided due to them having a negative reduction potential. In other literature it has been shows that cadmium chromium and zinc do not interfere with the manganese detection when using a carbon film electrode. [38] They also found that at elevated concentrations nickel, copper and lead had an influence on the manganese signal. Since lead does have a positive redox potential to form an oxide, this could also plate onto the electrode’s surface during the deposition of manganese. Additionally iron has been shown to influence the ability to detect manganese on both the carbon film electrode and an ITO electrode. [38, 28] This interference is possibly due to the fact that iron chemically strips the manganese oxide back o the electrode before the potential 74 sweep can be performed. [28] However, a previous study showed that typically Fe2+ concentrations in drinking water do not cause a concern for the ITO electrodes. Additionally anion interferences can be reduced with a negatively charged polymer placed onto the ITO’s surface. [28] A literature report also shows that using sodium citrate can help to eliminate this potential interference in waters that have high iron concentrations. [65] Table 3.3: Redox potentials for other metals that are potentially in environmental and biological samples 3.3.9 Key Advancements from Study Key advancements from this study include: • Development of an economically viable and consistent film material for metal sensing appli- cations that is already mass produced for other applications • Lower detection limits for manganese that allows for improved accuracy in human serum measurements • Understanding that heating and annealing of ITO film’s create the best film structure to be used as an electrode for electrochemical sensing applications 75 3.4 Summary The performance of four dierent indium tin dioxide films were investigated to determine which film would be the optimum electrode material for sensing manganese (Mn) with cathodic stripping voltammetry. Two samples were heated at 190 C during deposition and two samples were annealed in air for 1 hour at 400 C; generating one heated and annealed (HA), one heated and not annealed (HNA), one not heated and annealed (NHA), and one not heated and not annealed (NHNA). All films were thoroughly characterized. The samples that were heated showed comparable trends using AFM and XRD analysis. Both of these samples (HA and HNA) also showed higher roughness, crystallite sizes, and (400) XRD peaks with lower fwhm values compared to the non-heated films. The annealed samples had higher optical band gap values and carrier concentrations than those not annealed. 4-point-probe measurements displayed NHNA having the highest resistance and resistivity with HA having the lowest. HA, HNA, and NHA all exhibited good electrochemical detection ability for Mn. Only the HA film displayed a calibration curve that was analytically in acceptable range for the linear correlation coecient of 0.996. It can be concluded that heating during the deposition of ITO and post deposition annealing creates an optimum film for the CSV detection of Mn. The key parameters of the film may be its roughness, crystallite size, resistivity, and I(222)/I(400) obtained from XRD analysis. This study expands previous understanding of ITO and lays the ground work for further analyte detection on this material using CSV for an economical way to mass produce sensors. 76 BIBLIOGRAPHY 77 BIBLIOGRAPHY [1] Radhouane Bel Hadj Tahar, Takayuki Ban, Yutaka Ohya, and Yasutaka Takahashi. Tin doped indium oxide thin films: Electrical properties. Journal of Applied Physics, 83(5):2631–2645, 1998. [2] Giampiero de Cesare, Domenico Caputo, and Mario Tucci. Electrical properties of IEEE Electron Device ito/crystalline-silicon contact at dierent deposition temperatures. Letters, 33(3):327–329, 2012. [3] F Kurdesau, G Khripunov, AF Da Cunha, Md Kaelin, and AN Tiwari. Comparative study of ito layers deposited by dc and rf magnetron sputtering at room temperature. Journal of non-crystalline solids, 352(9):1466–1470, 2006. [4] C Guillén and J Herrero. Structure, optical, and electrical properties of indium tin oxide thin films prepared by sputtering at room temperature and annealed in air or nitrogen. Journal of applied physics, 101(7):073514, 2007. JH Park, C Buurma, S Sivananthan, R Kodama, W Gao, and TA Gessert. The eect of post- annealing on indium tin oxide thin films by magnetron sputtering method. Applied Surface Science, 307:388–392, 2014. [5] [6] Eunkyoung Nam, Young-Hun Kang, Dong-Jin Son, Donggeun Jung, Sung-Jei Hong, and Young Sung Kim. Electrical and surface properties of indium tin oxide (ito) films by pulsed dc magnetron sputtering for organic light emitting diode as anode material. Surface and Coatings Technology, 205:S129–S132, 2010. [7] M Gulen, G Yildirim, S Bal, A Varilci, I Belenli, and M Oz. Role of annealing temperature on microstructural and electro-optical properties of ito films produced by sputtering. Journal of Materials Science: Materials in Electronics, 24(2):467–474, 2013. [8] Necati Kaval, Carl J Seliskar, and William R Heineman. Spectroelectrochemical sensing based on multimode selectivity simultaneously achievable in a single device. 16. sensing by fluorescence. Analytical chemistry, 75(22):6334–6340, 2003. [9] Yining Shi, Andrew F Slaterbeck, Carl J Seliskar, and William R Heineman. Spectroelectro- chemical sensing based on multimode selectivity simultaneously achievable in a single device. 1. demonstration of concept with ferricyanide. Analytical chemistry, 69(18):3679–3686, 1997. [10] Jennifer M DiVirgilio-Thomas, William R Heineman, and Carl J Seliskar. Spectroelectro- chemical sensing based on multimode selectivity simultaneously achievable in a single device. 6. sensing with a mediator. Analytical chemistry, 72(15):3461–3467, 2000. [11] L Kerkache, A Layadi, E Dogheche, and D Rémiens. Annealing eect in dc and rf sputtered ito thin films. The European Physical Journal-Applied Physics, 39(1):1–5, 2007. 78 [12] Mika Yamaguchi, Ari Ide-Ektessabi, Hiroshi Nomura, and Nobuto Yasui. Characteristics of indium tin oxide thin films prepared using electron beam evaporation. Thin solid films, 447:115–118, 2004. [13] M Ait Aouaj, Rosy Diaz, Azzam Belayachi, F Rueda, and M Abd-Lefdil. Comparative study of ito and fto thin films grown by spray pyrolysis. Materials Research Bulletin, 44(7):1458–1461, 2009. [14] Kunisuke Maki, Nobuo Komiya, and Asako Suzuki. Fabrication of thin films of ito by aerosol cvd. Thin Solid Films, 445(2):224–228, 2003. [15] T Karasawa and Y Miyata. Electrical and optical properties of indium tin oxide thin films deposited on unheated substrates by dc reactive sputtering. Thin Solid Films, 223(1):135–139, 1993. [16] C May and J Strümpfel. Ito coating by reactive magnetron sputtering–comparison of properties from dc and mf processing. Thin solid films, 351(1):48–52, 1999. [17] M Rottmann and K-H Heckner. Electrical and structural properties of indium tin oxide films deposited by reactive dc sputtering. Journal of Physics D: Applied Physics, 28(7):1448, 1995. [18] I Baia, B Fernandes, P Nunes, M Quintela, and R Martins. Influence of the process parameters on structural and electrical properties of rf magnetron sputtering ito films. Thin Solid Films, 383(1-2):244–247, 2001. [19] Byung Hak Lee, Iee Gon Kim, Sung Woo Cho, and Si-Hyun Lee. Eect of process parameters on the characteristics of indium tin oxide thin film for flat panel display application. Thin Solid Films, 302(1-2):25–30, 1997. [20] Eugenia Matveeva. Electrochemistry of the indium-tin oxide electrode in 1 m naoh electrolyte. Journal of the electrochemical society, 152(9):H138–H145, 2005. [21] Hamid Reza Fallah, Mohsen Ghasemi, Ali Hassanzadeh, and Hadi Steki. The eect of annealing on structural, electrical and optical properties of nanostructured ito films prepared by e-beam evaporation. Materials Research Bulletin, 42(3):487–496, 2007. [22] K Sreenivas, T Sudersena Rao, Abhai Mansingh, and Subhash Chandra. Preparation and characterization of rf sputtered indium tin oxide films. Journal of Applied Physics, 57(2):384– 392, 1985. [23] John N Richardson, Aubrey L Dyer, Michael L Stegemiller, Imants Zudans, Carl J Seliskar, and William R Heineman. Spectroelectrochemical sensing based on multimode selectiv- ity simultaneously achievable in a single device. 13. detection of aqueous iron by in situ complexation with 2, 2 ‘-bipyridine. Analytical chemistry, 74(14):3330–3335, 2002. [24] Shirmir D Branch, Amanda M Lines, John Lynch, Job M Bello, William R Heineman, and Samuel A Bryan. Optically transparent thin-film electrode chip for spectroelectrochemical sensing. Analytical chemistry, 89(14):7324–7332, 2017. 79 [25] Hong-Jeong Kim and Juhyoun Kwak. Electrochemical determination of total alkaline phos- phatase in human blood with a micropatterned ito film. Journal of Electroanalytical Chem- istry, 577(2):243–248, 2005. [26] J Agrisuelas, D Giménez-Romero, JJ García-Jareño, and F Vicente. Vis/nir spectroelec- trochemical analysis of poly-(azure a) on ito electrode. Electrochemistry communications, 8(4):549–553, 2006. [27] Kolade Ojo, Daoli Zhao, Cory A Rusinek, Sarah K Pixley, and William R Heineman. Cathodic stripping voltammetric determination of cerium using indium tin oxide (ito). Electroanalysis, 29(4):1124–1130, 2017. [28] Cory A Rusinek, Adam Bange, Mercedes Warren, Wenjing Kang, Keaton Nahan, Ian Pa- pautsky, and William R Heineman. Bare and polymer-coated indium tin oxide as working electrodes for manganese cathodic stripping voltammetry. Analytical chemistry, 88(8):4221– 4228, 2016. [29] Tanya Shtoyko, Sean Conklin, Anne T Maghasi, John N Richardson, Aigars Piruska, Carl J Seliskar, and William R Heineman. Spectroelectrochemical sensing based on attenuated total internal reflectance stripping voltammetry. 3. determination of cadmium and copper. Analytical chemistry, 76(5):1466–1473, 2004. [30] Shruti Peshoria and Anudeep Kumar Narula. Bare indium tin oxide electrode for electro- chemical sensing of toxic metal ion. Journal of Materials Science: Materials in Electronics, pages 1–6, 2018. [31] Cory A Rusinek, Wenjing Kang, Keaton Nahan, Megan Hawkins, Cooper Quartermaine, Angela Stastny, Adam Bange, Ian Papautsky, and William R Heineman. Determination of manganese in whole blood by cathodic stripping voltammetry with indium tin oxide. Electroanalysis, 29(8):1850–1853, 2017. [32] US EPA, 2004. [33] Agency for Toxic Substances Registry and Disease. Toxicological profile for manganese. [34] John Stanislaus Roitz and KW Bruland. Determination of dissolved manganese (ii) in coastal and estuarine waters by dierential pulse cathodic stripping voltammetry. Analytica Chimica Acta, 344(3):175–180, 1997. [35] Wenjing Kang, Wei Yue, Xing Pei, Adam Bange, Erin Haynes, William R Heineman, and Ian Papautsky. Palladium-based sensor for electrochemical detection of manganese in the environment. pages 1146–1149. IEEE, 2013. [36] Alexander Goodwin, Andrew L Lawrence, Craig E Banks, Frédéric Wantz, Dario OmanoviÊ, äebojka Komorsky-LovriÊ, and Richard G Compton. On-site monitoring of trace levels of free manganese in sea water via sonoelectroanalysis using a boron-doped diamond electrode. Analytica chimica acta, 533(2):141–145, 2005. 80 [37] M Christine Welch, E Craig Banks, äebojka Komorsky-LovriÊ, and G Richard Compton. Electroanalysis of trace manganese via cathodic stripping voltammetry: Exploration of edge plane pyrolytic graphite electrodes for environmental analysis. 2006. [38] Olga MS Filipe and Christopher MA Brett. Cathodic stripping voltammetry of trace mn (ii) at carbon film electrodes. Talanta, 61(5):643–650, 2003. [39] Zuzana Janosikova Dusan Bustin Richard G. Compton Miroslav Rievaj, Peter Tomcik. Deter- mination of trace mn(ii) in pharmaceutical diet supplements by cathodic stripping voltammetry on bare carbon paste electrode. Chem. Anal. (Warsaw), 53(153), 2008. [40] Jason Stotter, Yoshiyuki Show, Shihua Wang, and Greg Swain. Comparison of the electrical, optical, and electrochemical properties of diamond and indium tin oxide thin-film electrodes. Chemistry of Materials, 17(19):4880–4888, 2005. [41] Asma El Kasmi, Michael C Leopold, Ryan Galligan, Rebecca T Robertson, S Scott Saavedra, Kacem El Kacemi, and Edmond F Bowden. Adsorptive immobilization of cytochrome c on indium/tin oxide (ito): electrochemical evidence for electron transfer-induced conformational changes. Electrochemistry Communications, 4(2):177–181, 2002. [42] Ciprian Iliescu, Hayden Taylor, Marioara Avram, Jianmin Miao, and Sami Franssila. A practi- cal guide for the fabrication of microfluidic devices using glass and silicon. Biomicrofluidics, 6(1):016505, 2012. [43] SM Sze. Physics of semiconductor devices. New York, Wiley-Interscience, 1981. 878 p., 1981. [44] Bard Allan J. and Faulkner Larry R. Electrochemical Methods Fundamentals and Applica- tions. John Wiley and sons, Inc., United States of America, second edition, 2011. [45] Strobel Howard A. and Heineman William R. Chemical Instrumentation: A Systematic Approach. John Wiley and Sons, Inc., United States of America, third edition, 1989. [46] Kissinger Peter T. and Heineman William R. Laboratory Techniques in Electroanalytical Chemistry. Marcel Dekker, Inc., New York, second, revised and expanded edition, 1996. [47] Saulius Graûulis, Daniel Chateigner, Robert T Downs, AFT Yokochi, Miguel Quirós, Luca Lutterotti, Elena Manakova, Justas Butkus, Peter Moeck, and Armel Le Bail. Crystallog- raphy open database–an open-access collection of crystal structures. Journal of Applied Crystallography, 42(4):726–729, 2009. [48] Saulius Graûulis, Adriana Daökevi, Andrius Merkys, Daniel Chateigner, Luca Lutterotti, Miguel Quiros, Nadezhda R Serebryanaya, Peter Moeck, Robert T Downs, and Armel Le Bail. Crystallography open database (cod): an open-access collection of crystal structures and platform for world-wide collaboration. Nucleic acids research, 40(D1):D420–D427, 2011. [49] Andrius Merkys, Antanas Vaitkus, Justas Butkus, Mykolas Okuli-Kazarinas, Visvaldas Kairys, and Saulius Graûulis. Cod:: Cif:: Parser: an error-correcting cif parser for the perl language. Journal of applied crystallography, 49(1):292–301, 2016. 81 [50] Robert T Downs and Michelle Hall-Wallace. The american mineralogist crystal structure database. American Mineralogist, 88(1):247–250, 2003. [51] David C Paine, T Whitson, D Janiac, R Beresford, Cleva Ow Yang, and Brian Lewis. A study of low temperature crystallization of amorphous thin film indium–tin–oxide. Journal of Applied Physics, 85(12):8445–8450, 1999. [52] Ho-Chul Lee and O Ok Park. Electron scattering mechanisms in indium-tin-oxide thin films: grain boundary and ionized impurity scattering. Vacuum, 75(3):275–282, 2004. [53] EA Davis and NFf Mott. Conduction in non-crystalline systems v. conductivity, optical absorption and photoconductivity in amorphous semiconductors. Philosophical Magazine, 22(179):0903–0922, 1970. [54] Nevill Francis Mott and Edward A Davis. Electronic processes in non-crystalline materials. OUP Oxford, 2012. [55] H Kim, CM Gilmore, A Pique, JS Horwitz, H Mattoussi, H Murata, ZH Kafafi, and DB Chrisey. Electrical, optical, and structural properties of indium–tin–oxide thin films for organic light-emitting devices. Journal of Applied Physics, 86(11):6451–6461, 1999. [56] A El-Denglawey, M Dongol, and MM El-Nahass. Photoinduced absorption edge shift of as 20 se 60 tl 20 films. Journal of Luminescence, 130(5):801–804, 2010. [57] J Barjon, N Habka, C Mer, F Jomard, J Chevallier, and P Bergonzo. Resistivity of boron doped diamond. physica status solidi (RRL)–Rapid Research Letters, 3(6):202–204, 2009. [58] Peihong Chen and Richard L McCreery. Control of electron transfer kinetics at glassy carbon electrodes by specific surface modification. Analytical Chemistry, 68(22):3958–3965, 1996. [59] Imants Zudans, Jean R Paddock, Hideki Kuramitz, Anne T Maghasi, Chamika M Wansapura, Sean D Conklin, Necati Kaval, Tanya Shtoyko, David J Monk, and Samuel A Bryan. Electro- chemical and optical evaluation of noble metal–and carbon–ito hybrid optically transparent electrodes. Journal of Electroanalytical Chemistry, 565(2):311–320, 2004. [60] Romana Jaroöová, Paula M De Sousa Bezerra, Catherine Munson, and Greg M Swain. Assessment of heterogeneous electron-transfer rate constants for soluble redox analytes at tetrahedral amorphous carbon, boron-doped diamond, and glassy carbon electrodes. physica status solidi (a), 213(8):2087–2098, 2016. [61] Melanie Poon, Richard L McCreery, and Royce Engstrom. Laser activation of carbon elec- trodes. relationship between laser-induced surface eects and electron transfer activation. Analytical Chemistry, 60(17):1725–1730, 1988. [62] Richard S Nicholson. Theory and application of cyclic voltammetry for measurement of electrode reaction kinetics. Analytical Chemistry, 37(11):1351–1355, 1965. [63] Liang Liu, Shai Yellinek, Ido Valdinger, Ariela Donval, and Daniel Mandler. Important implications of the electrochemical reduction of ito. Electrochimica Acta, 176:1374–1381, 2015. 82 [64] Daniel C. Harris. Quantitative Chemical Analysis. Clancy Marshall, New York, 8th edition, 2010. [65] Kathleen E Berg, Jaclyn A Adkins, Sarah E Boyle, and Charles S Henry. Manganese detection using stencil-printed carbon ink electrodes on transparency film. Electroanalysis, 28(4):679– 684, 2016. 83 CHAPTER 4 ELECTROCHEMICAL OXIDATION FOR THE MINERALIZATION OF PER- AND POLYFLUOROALKYL SUBSTANCES 84 4.1 Introduction A portion of this work has been published in Water and Environment Journal1. Per- and polyfluoroalkyl substances (PFAS) are man-made chemicals that began production in the early 1950’s. Since their discovery, they were used heavily in customer and industrial applications that took advantage of their oil and water repelling properties. [1–9] The main property of this class of chemicals that sets them apart from others are their fully fluorinated carbon chains or "tails". Under the broad term of PFAS, which now encompasses more than 4,000 compounds, there is a main group of perfluoroalkyl acids (PFAA) which encompasses the perfluoroalkylcarboxylic Acids (PFCA) and Perfluoroalkylsulfonates (PFSA) with each group comprising of compounds with four to 18 carbons in their background. The dierences between the PFCA and PFSA are the head groups that are attached to the carbon chain tails. The two most popular chains that are being studied in recent literature are perfluorooctanoic acid (PFOA) and perfluorooctane sulfonate (PFOS) since they are widely distributed and have potentially negative health eects. The molecular structures of PFOA and PFOS in shown in Figure 4.1. Due to their fully fluorinated carbon chain, they are stable in the environment and bioaccumulate rapidly. Both compounds have been phased out by industry but wide spread contamination still remains an issue for human exposure. Few standard water treatment methods have been successful at remediating these compounds. Those that have been can be divided into two classes, that of adsorption and filtration and that of destruction. Methods for adsorption and filtration include ion exchange columns, nanofiltration, ultrafiltration, and reverse osmosis. Ion exchange (IX) columns have the added ability of being regenerated for reuse which saves on cost. Some of the resins researched for PFAS include Purolite A600E, A5202E, A532E, PAD500, PAD428, MN102, Sorbix A3F, PFA300, IRA67, IRA958, and IRA67 [10–15]. However, although they can remove PFAS from a certain source, they do not destroy the compounds. Ultimately, the PFAS is just concentrated into a secondary waste product. With improper disposal of this waste, the contamination cycle can continue. Due to this being the 1 Water and Environment Journal 2020, DOI:10.1111/wej.12616 85 Figure 4.1: Molecular structure of (top) PFOA and (bottom) PFOS case will all filtration and adsorption techniques, there is a need for a destructive treatment method to destroy the PFAS once and for all. Destructive methods include thermal, plasma based water treatment, and electrochemical oxi- dation. Thermal destruction utilizes temperatures above 1000 C and may not completely combust of all the PFAS chains. [4, 6, 3] In addition, the use of thermal techniques lead to the generation of various undesirable green house gasses. [3] Plasma treatment is new for PFAS, but has obtained promising results for PFCA utilizing a laminar jet bubbling reactor and argon gas. [16, 17] Elec- trochemical oxidation (EO) does not require the use of additional gasses or chemicals. It operates using applied current to perform direct and indirect oxidation of contaminant species. [18] Elec- trode materials are limited for water based remediation of PFAS due to the need of high current and voltage. Boron-doped diamond (BDD) has been a widely explored material for these applications. [5, 19–31] It can withstand the high current and potential while promoting hydroxyl radicals which can be of benefit in indirect oxidation of co-contaminants. [18] 86 EO has been applied in literature studies to address PFAS contamination using current densities from 0.04 - 50 mA/cm2. [5, 28, 22, 23, 27, 21, 25, 24, 26, 19, 20, 29, 31] Time frames for these treatment experiments have ranged from minutes up to 120 hours.[21, 29, 26]. One of the daunting concerns with EO can be the high cost of the electrode material and high energy consumption, but it is important to remember that EO is one of the only remediation techniques that has proven to provide complete mineralization of PFAS. [27] All possibilities to reduce operating costs should be considered. The following sections discuss using EO independently and in conjunction with a regenerable IX column. [3, 4, 6, 1] There is a focus on combining current densities to evaluate if the overall energy associated with the process can be lowered while still obtaining the desired level of compound destruction. 4.2 Electrode Stack Characterization Before performing electrochemical oxidation experiments for the destruction of PFAS, the electrode stack being used needed to be characterized. This stack was comprised of two double side coated BDD on niobium (Nb) anode plates and three double side coated BDD on Nb cathode plates. The two anodes were sandwiched in between the cathodes and an exposed anodic surface area of 105 cm2 was used with 3 mm distance in between each plate. Characterization was done using scanning electron microscopy (SEM), Raman spectroscopy, and electrochemical methods to determine the capacitance and rate constants for two redox couples. The SEM image in Figure 4.2 shows a high surface area film with no cracks on the surface. Raman spectroscopy was used to determine the boron content of the film. The spectrum is shown in Figure 4.3 and the total boron content was 4.88 x 1020 cm-3 +/- 5.7 x 1019 cm-3. This classifies this film as highly doped. In order to find the electrochemical capacitance of the film, cyclic voltammetry (CV) was used. This experiment was performed by exposing a small area of the electrode to a 1 M potassium chloride (KCl) solution and measuring the current as the potential was scanned using a laboratory potentiostat. After each scan, the scan rate was increased and the measurement was performed again. The current was then recorded at 0 V and plotted against the scan rate at which is was 87 Figure 4.2: Scanning electron microscopy (SEM) image of the BDD surface Figure 4.3: Raman spectroscopy of BDD electrodes obtained, Figure 4.4. The scan rates of 0.1 - 0.5 V/s were used and three sets of each were performed. The capacitance of the film was 2.66 x10-04 F/cm2 +/- 4.88 x10-07 F/cm2, which was obtained from the slope of the graph relating the current and scan rate. The electrochemical characterization was completed by determining the rate constants of the BDD for an inner and outer sphere electron redox couple. The inner sphere redox couple that was chosen was potassium ferrocyanide (FeCN) and the outer sphere redox couple that was chosen was ruthenium hexaammine (RuHX). Each redox couple was used in a 1 M KCl solution and the current 88 Figure 4.4: Current versus scan rate plots to determine the capacitance of BDD was measured as the potential was scanned for scan rates of 0.05 to 1.0 V/s. The Nicholson and Shain method was used to determine each rate constant. All together, each scan rate was performed a total of three times so an average could be taken. Figure 4.5 shows a CV for both redox couples. The rate constant for FeCN was 3.96 x10-04 +/- 4.47 x10-06 and the rate constant for RuHx was 9.00 x10-04 +/- 3.28 x10-05. Figure 4.5: Cyclic voltammogram of both ferrocyanide and ruthenium hexamine on the BDD surface 89 4.3 Constant Current 4.3.1 Methods All electrochemical experiments were performed with a 180W multi range 60V/8A DC power supply from BK Precision Instruments. All solution matrices were made with analytical grade sodium sulfate (Na2SO4) from J.T. Baker Inc. at a concentration of 50 mM. Perfluorooctanoic acid (PFOA) and perfluorooctane sulfonate (PFOS) were purchased from Sigma Aldrich. Nalgene bottles were used for all solutions and testing volumes were 750 ml. Solutions were constantly stirred during testing using a Hanna Instruments magnetic stirrer HI 190M on max. A pH probe, temperature probe, and fluoride ion selective electrode (ISE) were purchased from Thermofisher Scientific and a fluoride standard that was used to calibrate the ISE was purchased from Sigma Aldrich. 2ml of sample was removed and stored in the refrigerator in polyethylene vials at 4 C until analysis was performed with the fluoride ISE. A total ionic strength buer (TISAB) was made using glacial acetic acid (Fisher Chemical), reagent grade sodium chloride (NaCl) (Sigma Aldrich), sodium hydroxide (NaOH) (Spectrum Quality Products Inc.) and DI water and used to dilute each sample 1:1. The Michigan State University College of Natural Science Research Technology Support Facility performed the liquid chromatography mass spectroscopy (LC/MS) measurements for the samples spiked with only PFOA. 4.3.2 Results and Discussion 4.3.2.1 PFOS PFOS studies were performed with the previously characterized electrode stack. Constant current densities of 5, 10, 25, and 50 mA/cm2 were used. Each experiment was performed in triplicate. The current densities of 5 and 10 mA/cm2 were performed for 1.5 hours but the current densities of 25 and 50 mA/cm2 were only performed for 1 hour. The starting concentration of 90 PFOS was 10 parts per million (ppm) as calculated for the amount of powder that was added. With increasing current density the initial temperature, pH, and potential of each system also increased. Figure 4.6: PFOS Temperature, pH, and voltage trends for the constant current density studies These figures are shown in Figure 4.6. pH for all tests began at 6 and increased over time. The pH for 10, 25, and 50 mA/cm2 increased to 10.5 whereas the pH of the 5 mA/cm2 solution has an upwards trend tending towards 10.5 but does not reach there in 1.5 hour. BDD in a known ecient producer of hydroxyl radicals and previous studies using BDD have shown that the pH could increase by 5 units depending on the applied current density. [32, 33] The temperature started at room temperature for all tests and increased with time and potential for the low current densities of 5 and 10 mA/cm2 remained constant, whereas, the potential for both 25 and 50 mA/cm2 both slightly decreased over time. PFOS degradation was monitored with the F- ISE and the concentration of free F- was used to calculate the percentage of defluorination. Figure 4.7 shows the percent of 91 defluorination for the current densities studied. Figure 4.7: PFOS reduction overtime with each current density tested The three tests at each current density are taken into account and an average is graphed with its corresponding error. It is clear that the PFOS degradation is independent of current density for these studies. Each current density also had a standard deviation of 4 % at each time point. The 1st order rate constants for each current density were also determined and graphed as an average value. The PFOS concentrations were calculated based on the percentage of fluoride released under the assumption that no shorter chains were being formed due to lack of analysis equipment. The rate constants for each were 0.39 +/- 0.02 h-1, 0.39 +/- 0.07 h-1, 0.45 +/- 0.10 h-1, and 0.45 +/- 0.10 h-1, for 5 at 1.5 hours, 10 at 1.5 hours, 25 at 1 hour, and 50 mA/cm2 at 1 hour respectively. Each rate constant had a linear correlation coecient (R2) above 0.85. The average linear correlations were 0.9821, 0.958, 0.951, and 0.8628 for 5, 10, 25, and 50 mA/cm2, respectively. 5 and 10 mA/cm2 share the same rate constant as do 25 and 50 mA/cm2. To ensure the validity of the PFOS results, a quality assurance check was performed with the ISE. To do this, each of the 1 hour time points from the three 25 mA/cm2 runs were retested with the ISE at the end of the complete study. The repeatability in the results obtained was 105 +/- 5.6 %, validating that the ISE measurements were accurate. To advance our understanding of how 92 Figure 4.8: PFOS 1st order rate constants PFAS behave in electrochemical experiments, perfluorooctanoic acid (PFOA) was studied with the addition of LC/MS analysis. 93 4.3.2.2 PFOA The degradation of PFOA was studied at dierent applied current densities using the previously characterized electrode stack. Current densities included 1, 5, 10, 25, and 50 mA/cm2. All tests were performed for 1 hour. The stack was lowered into 750 ml of 50 mM sodium sulfate that the PFOA was spiked into using a concentrate made from the dissolved powder. Cell voltages spanned rom +4 V to +9 V at the lowest and highest current density, respectively. All solutions were started at room temperature and experiences an increase to 40 C at 50 mA/cm2. All testing solutions began at a pH of 5, one pH unit lower than the PFOS solutions, and increased over time to 10.5 except for the current density of 1 mA/cm2, where the pH remained constant. The current densities chosen here suggest that the reactions aecting the pH may be dierent between 1 mA/cm2 and the rest of the current densities. Graphs showing the change in temperature, pH, and potential for each current density are shown in Figure 4.9. The temperature increased over time for all current densities studied with the highest having the largest increase. The potentials were stable across the full testing time for all current densities except for 50 mA/cm2 where a slight increase was observed. The starting concentration of PFOA was 6.5 parts per million (ppm) as determined by liquid chromatography mass spectroscopy (LC/MS) measurements. A graph representing the PFOA degradation over time, in terms of percent defluorination obtained by using the F- ISE, is shown in Figure 4.10. The release of fluoride increases over time with increasing current density. The five tests also show good repeatability between the three sets that were performed. Each current density study was also fit to first order kinetics using the PFOA degradation over time from the LC/MS analysis. This way the direct rate constant relating to PFOA can be determined instead of inferring from the ISE analysis as was done for PFOS. These plots are shown in Figure 4.11. All current densities has a linear correlation coecient of greater than 0.90 with both 5 and 10 mA/cm2 being above 0.99. The rate is increasing as current density is increased. the values for k1 are as follows: 0.34, 1.35, 2.04, 3.77, and 6.27 h-1, for 1, 5, 10, 25, and 50 mA/cm2, respectively. 94 Figure 4.9: PFOA Temperature, pH, and voltage trends for the constant current density studies Figure 4.10: PFOA degradation over time at each current density value Another interesting takeaway from this constant current density study comes from taking a closer look at Figure 4.10. This being that the slopes of the PFOA defluorination lines only seem 95 Figure 4.11: 1st order rate constants for PFOA based on LC/MS measurements for each current density study. to dier in the first 0.25 h of each test. For the remaining time of each, the slopes are relatively the same. For instance, for the first 0.25 hours of the test, 1 mA/cm2 had a slope of 0.28 %/h that increased to 2.0 %/h at 50 mA/cm2. However, for 0.25h untill the end of 1 hour, the slope for 5 through 50 mA/cm2 were 0.54, 0.58, 0.62, and 0.49 %/h for 5, 10, 25, and 50 mA/cm2, respectively. This raises the question as too if the higher current density is truly only beneficial in the first 0.25 h of each study and could then be lowered to save on energy consumption. Fluoride quality analysis was also performed for PFOA using the 1 hour time sample from each 25 mA/cm2 test. The obtained concentration of fluoride was 99.7 +/- 1.3 % of what the original tested value was. This verified the F- ISE measurements were accurate over time for this study. 4.3.2.3 Comparison of ISE Results to LC/MS Results At the time of analysis, the Michigan State University mass spectroscopy facility could only analyze for the single compound of PFOA. Due to this limitation, shorter chain compounds could not be directly identified, but by determining if there was a dierence in the concentration of PFOA that remained from ISE analysis and what the LC/MS provided, it could be assumed that any dierence was from the generation of shorter chain fluorinated carboxylic acids. All five current 96 densities that were used were tested with LC/MS and the PFOA concentrations were graphed against those calculated from the ISE. Figure 4.12: Comparison of the PFOA concentration obtained by the F- ISE and LC/MS Figure 4.12 shows the comparison of the concentrations obtained using the two dierent analysis methods. The concentrations for both methods overlap for the lowest current density of 1 mA/cm2 at 4.3 +/- 0.2 ppm from the ISE and 4.4 from LC/MS after 1 hour, but as the current density is increased, the two concentrations begin to dier. The dierences between the two methods are shown in Table 4.1. Since there is a smaller amount of PFOA detected with LC/MS than there is free fluoride that would pertain to that amount of degradation, the remaining PFOA is assumed to have converted to shorter chains through an unzipping mechanism. 97 Table 4.1: Comparison of remaining concentration of PFOA obtained from the ISE analysis and LC/MS analysis 4.3.3 Fluorination of BDD It has been shown in literature that perfluoroalkyl compounds can be used to fluorinate a BDD surface. [34, 35] As such, it is important to consider that some of the fluoride released during the decomposition of PFAS could adsorb onto the BDD surface. This was investigated in three ways: (1) determining the contact angles between a unused and used BDD electrode for PFAS remediation (2) measuring a potassium ferrocyanide (FeCN) CV on a used and unused electrode, and (3) investigating the potential window changes between a used and unused electrode. The contact angle measurements were done using a 0.5 µL drop of water on the surface of each anode. The part of each electrode that was submerged into solution was labeled as "used" and the part that was not submerged and remained above the solution volume was labeled as "unused". For the anodes, the used electrodes displayed an average contact angle of 84 +/- 5.6 and the unused portion displayed 75 +/- 4.9. The FeCN study was used due to its behavior as an inner sphere redox couple. If the fluoride was deposited onto the BDD surface, the surface sensitive reaction of ferrocyanide could be eected. Cyclic voltammetry was performed in 1 mM FeCN in 1M KCl at 0.1 V/s on the used and unused portion of the BDD anode used in the previous PFOA degradation studies. Figure 4.13 shows the voltammogram for both electrode portions. It is immediately apparent that the response from the two diers. The unused portion has a characteristic shape showing the oxidation and reduction of ferrocyanide whereas the used portion has no visible peaks. This is indicative that there was a 98 change to the surface of the BDD anode during the remediation experiments with PFOS and PFOA. To investigate this further, the potential windows of the used and unused portions were compared. Figure 4.13: 1mM ferrocyanide in 1 M KCl measured on both the used and unused portion of the anode that was used for the constant current density PFOA degradation studies The potential window of each anode surface was investigated in 1 M sulfuric acid at a scan rate of 0.1 V/s to determine if there was a shift in the hydrogen or oxygen evolution potentials. Four scans were performed with the last two being graphed for each in Figure 4.14. It can be seen that the used anode has a slightly larger oxygen evolution potential and a further negative hydrogen evolution potential. According to literature, fluorinated BDD electrodes have a larger window compared to diamond electrodes and can show a potential window of 5 V [36–38]. There is a larger dierence in the hydrogen evolution between a fluorinated BDD and non fluorinated BDD compared to the oxygen evolution due to the fact that the fluorinated surface is more hydrophobic, seen with contact angles, which decreases the adsorption strength of water and atomic hydrogen shifting the reaction more negative [37]. Between the contact angle measurements, investigation into the redox couple of ferrocyanide, and the comparison of the potential window for the used and unused BDD, it is confirmed that there was fluorination of the diamond surface during the electrochemical oxidation experiments. In an attempt to bring the BDD surface back to its original non-fluorinated state, it underwent 99 Figure 4.14: CV in 1 M sulfuric acid for both the used and unused portion of the anode a hydrogen termination process and re-evaluation of the FeCN CV and potential window study. Hydrogen termination of the used surface was performed by placing it into a hydrogen atmosphere in a hot filament chemical vapor deposition reactor for one minute. The resulting ferrocyanide voltammogram and potential window scan are shown in Figures 4.15 in (A) and (B), respectively. Figure 4.15: (A) CV comparison in1 mM ferrocyanide in 1 M KCl for the used anode, unused anode, and used anode post hydrogen termination and (B) CV in 1 M sulfuric acid for the used, unused, and used electrode post hydrogen termination It can be seen that the FeCN peak showed a significant current response and clear oxidation and reduction peaks for the hydrogen terminated used sample. The current response after hydrogen termination was greater than the response from the unused sample potentially due to now being 100 cleaner and fully hydrogen terminated whereas the unused portion may have become partially oxygen terminated over time. In addition, the potential window shifted closer to the unused electrode by having the hydrogen evolution reaction shift to more positive values and the oxygen evolution reaction shift to more negative values. Both of these tests verified that the fluorination of the BDD surface could be removed by placing the electrode in a hydrogen plasma after being used for the degradation experiments. 4.4 Combined Current Density To further evaluate on the ability of electrochemical oxidation to degrade PFOA over time and to begin to investigate how to save on energy costs, a combined current density technique was investigated. This technique employed the use of a high current density for the first 15 minutes of each test before lowering it down for the remaining 75 minutes. This decision came from the fact that the most significant benefit for the high current density application seen with the constant current densities was in the first 15 min of each test. The experimental conditions used in the previous constant current density section were employed again in this study. The high current density chosen was 50 mA/cm2 as it showed the greatest initial change in PFOA concentration from Section 4.3.2.2. Three lower current densities of 1, 5, and 10 mA/cm2 were investigated for the remaining time frame of each test. Throughout each test, samples were removed at 0.25, 0.5, 0.75, and 1 hours and stored in the refrigerator until analysis was performed. Graphs showing the change in temperature, pH, and potential for each current density are shown in Figure 4.16. The temperature from all three combined tests increases in the first hour before steadily decreasing with the reduction in applied current. The pH for all three tests also mirrored one another with an increase before stabilizing out at 10.5. The potential slightly decreased during the higher current density application and then remained stable during the lower current density application for all three tests. Each PFOA sample was evaluated with the fluoride ISE as used previously. The starting concentration of PFOA was 6.5 ppm in 50 mM sodium sulfate determined from LC/MS experiments. There was 750 ml of solution placed in a Nalgene bottle and stirred continuously. The PFOA reduction in terms of defluorination percentage over time for 101 Figure 4.16: Temperature, pH, and voltage trends for the combined current density studies each combined current density compared to the constant current density of 50 mA/cm2 provided as Figure 4.17. Figure 4.17: PFOA reduction overtime with each combined current densities tested together as well at the standard 50 mA/cm2 for comparison 102 It can be seen that the 50 mA/cm2 combined with 1 mA/cm2 (50&1), has the lowest evolution of fluoride between each run, but a low standard deviation (n=3) between each run. The 50&10 run also has a low standard deviation between sample sets. For the run using 50&5, the points line up close to those of 50&10, and even overtake them at 1 h, but there is a much larger standard deviation between the tests. As such, using 50&5 could outperform 50&10 as well as be very similar to the degradation that is obtained with the constant current density run of 50 mA/cm2. In order to evaluate the rate constants of each combined current density test, two sequential rate constants were used to analyze the fluoride ISE data. k1 for the rate constant from time 0 to 0.25 hours, and k2 from 0.25 hours until the end of the study, 1.5 hours. Figure 4.18: First order rate constant fits for the first 0.25 h, k1, and the remainder of each test, k2. Figure 4.18 shows the fitted rate constants for the three combined current density studies. Table 4.2 shows the numerical values for each of the rate constants and their obtained linear fits. It is noted that k1 is only based o of time 0 and 15 min. With only two points the linear coecient will always be equal to 1. k2 is based o of 5 points allowing for a better understanding of if a linear relationship exists. k1 is the same for both 50&1 and 50&5 with 50&10 having a slightly lower value. Overall, since this time point it taken based solely on the applied high current density of 50 mA/cm2, it shows good repeatability. 50&5 had the largest standard deviation in the k1 value. For k2 all the 103 Table 4.2: Rate constants and their linear fits for the three tested combined current densities linear coecients are all above 0.98 with 50&1 and 50&5 being greater than 0.99. 50&1 and 50&10 have a lower standard deviation in their values compared to 50&5. Due to this the k2 value for 50&5 and 50&10 are similar. The k2 value for 50&1 is roughly half in comparison. 4.4.1 Comparison of ISE and LC/MS As performed for the constant current density study, a LC/MS and F- ISE comparison was done for the best combined current density study of 50&5. The dierence in PFOA concentrations over time for each analysis technique are shown in Figure 4.19. The same trend seen before is shown with this data. The concentration of PFOA obtained from the LC/MS data is lower than that determined from the ISE analysis. Again, this suggests that the remaining fluoride not detected from the ISE is bound in other short chains that were formed as part of the degradation process, or have fluorinated the anode’s surface. 1.71 +/- 0.1 ppm of PFOA was estimated to be remaining by the ISE analysis but LC/MS provided a final concentration of 0.40 ppm showing the increased ability in electrochemical oxidation to break apart PFOA. The rate constant was also determined for 50&5 based on LC/MS measurements. For time 0 to 0.25h the rate constant was 4.20 h-1 and from 0.25 to 1 h the rate constant was 2.85 h-1. These rate constants are much higher than those corresponding to the ISE analysis. This again suggests that PFOA may be degrading quickly into shorter chain PFAS. 104 Figure 4.19: Comparison of the PFOA concentration obtained by the F- ISE and LC/MS 4.4.2 Remediation of Simulated Ion Exchange Regenerate Solutions To advance the application of the combined current density method, a treatment train approach that can enhance an existing technology would be beneficial. One of the most promising remediation techniques is adsorption using ion exchange (IX). This has been shown successful in literature for removing a wide array of PFAS. Some of these columns that remove PFAS in particular have the ability to be regenerated and re-used. [10, 5, 13–15, 11, 12] This is a huge cost benefit, but has the downside of generating a concentrated waste product. The regenerate solution out of an IX process can comprise of PFAS in the parts per million (ppm) range in a complex matrix of a brine with an alcohol. [12] Due to diculties in disposal of this waste, and the potential to continue the contamination cycle with improper disposal, a destruction method is needed to remove the PFAS once and for all. However, due to the nature of the regenerate solution, i.e. high PFAS concentration and high salt, it is a perfect electrolyte solution to apply electrochemical oxidation (EO). In order to determine if EO is an appropriate treatment technology for these solutions, those currently existing in literature to regenerate PFAS specific columns need to be tested. The four main ion exchange (IX) regenerate solutions that were simulated in the laboratory include: 6% ammonium hydroxide (NH4OH) with 2% methanol (MeOH), 6% sodium chloride (NaCl) with 2% MeOH, 6% sodium hydroxide (NaOH) with 2% MeOH, and 0.5% NH4OH with 105 0.5% ammonium chloride (NH4Cl). The solutions with MeOH were done at 2% MeOH due to the ability to distill o the majority in a distillation process that can then be reused for future regeneration.[12] NH4OH with MeOH was chosen due to its ability to remove PFOA from the IX columns of PUROSORB PAD 500 and MACRONET MN102. [11] NaCl with MeOH was chosen for its ability to remove PFOS from IX columns of PFA300, IRA958, and BAC. [14, 13, 15] NaOH with methanol was chosen for its ability to regenerate PFA300 and IRA67 for PFOS removal. [14, 13] The mixture of NH4OH and NH4Cl was chosen for its ability to regenerate Purolite A600E and A520E with PFOA, PFBA, PFOS, and PFBS. [10, 11] Each solution was spiked with greater than 10 ppm of PFOA and PFOS due to both being in the tens of ppm range in known ion exchange solutions. [39] The high current density application time was increased from 15 minutes, from the earlier studies, to 1 hour due to the high chemical oxygen demand (COD) present in most of the simulated regenerate solutions. At 2% MeOH, the COD values for these solutions were around 30,000 ppm. The total testing time was also increased to 12 hours to account for the additional competing COD content and higher initial PFAS concentrations. 4.4.2.1 Comparison of 50 mA/cm2 to 50&5 mA/cm2 The first IX regenerate solution comparison performed utilized the solution of 6% NH4OH with 2% MeOH was to compare the use of 50 mA/cm2 for an entire test to using a combination of 50&5 mA/cm2. This was done by using the higher current density for an extended time of 1 hour before lowering the current density for the remainder of the test. Graphs showing the change in temperature, pH, and potential for each current density are shown in Figure 4.20. The pH for the 50&5 solution decreases much more rapidly in the first 1.5 hours although both tests employed 50 mA/cm2 for the 1st hour. 50&5 then showed an increase in pH where it ended at about 10 compared to the constant current density that had a continued gradual decrease and ended closer to 8. The temperature for moth tests mirrored on another for the higher current density application. after 50&5 was lowered to 5 mA/cm2, the temperature fell and stabled out. The 50 mA/cm2 run only displayed a slight decrease in temperature over time. The potential for other tests started 106 relatively high and fell rapidly in the first our before relatively stabling out. The 50&5 test had a lower potential for the lower applied current. Figure 4.20: Temperature, pH, and voltage trends for the comparison of 50 and 50&5 in IX simulated solution Figure 4.21: Comparison of the PFOA (left) and PFOS (right) degradation using 50 mA/cm2 and 50&5 mA/cm2 The change of PFOA and PFOS over time are shown in Figure 4.21 left and right, respectively. 107 When using the constant current density of 50 mA/cm2 for the entire test, the COD was reduced by 72% and the ammonia was reduced by 88% compared to 43% and 47% for the combined current density of 50&5, respectively. The PFOA concentration in the constant current density test decreased by greater that 99% in 12 hours. Based on the dilution factor used when sending samples for analysis, it is known that there is <66 parts per billion (ppb) remaining, but the exact number is unknown. The PFOS degradation was also greater than 99% for this constant current density study after 12 h. For the 50&5 study, the PFOA concentration degraded 90% and the PFOS degraded 59% in the same total time frame. It can be seen from the PFOS degradation graph (Figure 4.21) that PFOS reached a higher degradation percent in 8 hours then at 12 hours using the combined current density. This may be due to a number of factors including, generation of PFOS from precursors or an outlier in the measurement technique. Generation from precursors is unlikely since the solutions were spiked with purchased powders. However, these powders are not 100% pure. The human error in removing a sample and diluting x100 may have caused for the discrepancy. Overall, 50&5 obtained good degradation results for PFOA and PFOS in the same time frame as using 50 mA/cm2 for the entire time. Since this is only one potential IX regenerate solution, it is of importance to test the other solutions researched in industry with a further analysis into the degradation of PFAS. 4.4.2.2 Comparison of four dierent regenerate solutions In order to compare the ability of using 50&5 further, each of the simulated solutions were tested using 50&5 for the degradation of primarily PFOA and PFOS. Graphs showing the change in temperature, pH, and potential for each current density are shown in Figure 4.22. The highest pH belonged to the NaOH and MeOH solution at 14 compared to all of the others that fluctuated at levels below 10. The temperature increased for all solutions during the high current density application and lowered once the current density was lowered. The highest temperature of 60 C was reached with the NH4OH and MeOH solution. This was also the solution that had the highest starting potential out of all those tested. The other three solutions had similar potentials at the high 108 and low current density applications. Figure 4.22: Temperature, pH, and voltage trends for the comparison of the four IX simulated solutions Figure 4.23: Comparison of the PFOA (left) and PFOS (right) degradation using the four tested simulated IX solutions The change in PFOA and PFOS for each solution over the 12 hours is shown in Figure 4.23 with PFOA degradation on the left, and PFOS degradation on the right. PFOA degradation was obtained in three out of the four tested solutions. Its degradation was 90% in the NH4OH and 109 MeOH solution, 52% in the NaCl and MeOH solution, and >99% in the NH4OH with NH4Cl solution. It was not degraded in the NaOH based solution. This lack of degradation in NaOH may be due to the significantly higher pH that was characteristic of this solution compared to the others (Figure 4.22). It has been shown in literature that the degradation of PFOA can be inhibited in solutions with alkaline pH. [40, 29, 31] Each PFOA experiment was fit to sequential 1st order reaction kinetics, Figure 4.24. Figure 4.24: 1st order rate constant kinetics for PFOA in each simulated solution For the first hour, the PFOA rate constants were 1.93, 0.06, 0.75, and 1.5 h-1 for NH4OH and MeOH, NaOH and MeOH, NaCl and MeOH, and NH4OH with NH4Cl, respectively. After one hour, the rate constant for every solution expect for NH4OH with NH4Cl was 0 h-1. That of NH4OH with NH4Cl was 0.37 h-1. These results show that there is no additional degradation in the simulated solutions for the majority of the tests after the current density is lowered. It is interesting to note that out of the two solutions with the highest degradation for PFOA, the NH4OH solution with MeOH obtains all of that degradation at the higher current density. Whereas, the combination of NH4OH and NH4Cl does have slow continued degradation. In the first hour, the NH4OH with MeOH solution also has the lowest pH value during this time and the highest achieved temperature which both may contribute to the increased degradation of PFOA. The other ammonia based solution may have ultimately outperformed the solutions with MeOH due to the fact that 110 this solution does not have the competing COD that the other solutions contain. Table 4.3 lists the PFOA concentrations, rate constants, and linear fits obtained from the studied solutions at 1h and 12 h. This table shows that there is no added benefit, in terms of PFOA degradation, to continue to run the test past 1 h for all solutions except for that based on NH4OH and NH4Cl. It is also noticeable that in 1h the NH4OH solution with MeOH obtained the most PFOA degradation with the fastest rate constant despite having the added competition of COD although this rate constant does not have a nice linear fit. Table 4.3: Comparison of 1 h PFOA degradation and 12 h PFOA degradation For PFOS, only three out of the four solutions obtained degradation. However, the solution that PFOS did not degrade in is dierent from the one that PFOA did not degrade in. PFOS did not degrade in the NaCl with MeOH solution. In addition it obtained a much lower degradation in the NaOH with MeOH compared to the two ammonia based solutions. The degradation was 40% in the NaOH solution, 59% in the NH4OH with MeOH solution, and 91% in the NH4OH with NH4Cl solution at the end of 12 hours. Sequential first order rate constants were also determined for PFOS for the first hour followed by the remainder of the study. The two solutions with non zero rate constants for the 1st hour of the study were the two NH4OH based solutions with 0.29 h-1 and 2.73 h-1 for that with MeOH and that with NH4Cl, respectively. They had a linear correlation of 0.089 and 0.59 for NH4OH with MeOH and NH4Cl, respectively. For the remaining portion of the test, 1-12h, the rate constants were 0.05 hr-1 for both NH4OH and NaOH with MeOH and 0.02 for NH4OH with NH4Cl. The fittings for each rate constant can be shown in Figure 4.25. It is also 111 interesting to note that the NH4OH solution with NH4Cl now had the highest degradation in the first hour for PFOS compared to the NH4OH solution with MeOH; opposite of the PFOA trend. In addition the mixed ammonia solution did not have continued degradation past this point whereas the ammonia based MeOH solution did. Figure 4.25: Sequential 1st order rate constant kinetics for PFOS in each simulated solution A literature study on a dierent electrode materials also found that PFOS did not degrade in their electrolyte of sodium perchlorate (NaClO4) when starting with 20 ppm and applying roughly 7 mA/cm2. [41] It has been proposed in the literature that PFOS prefers the gas-liquid interface which may inhibit its degradation. [42] To investigate this electrochemically, the oxygen evolution reaction for each solution on the BDD surface was studied. The oxygen evolution reaction was investigated in each IX solution using a clamp cell placed on top of the used BDD anode. A counter electrode of graphite and a reference electrode of Ag/AgCl was used. The exposed surface area was 0.13 cm2. A graph of the current density compared to potential is shown in Figure 4.26 The two solutions with the lowest PFOS degradation have the earliest onset of the OER, the NaCl and NaOH based electrolytes. The two solutions with the highest PFOS degradation had the furthest potential where the OER began. This suggests that the oxygen evolution could play a role in inhibiting the PFOS from reaching the anode surface to degrade. A report comparing two dierent electrode materials in the same solution looked into the OER compared to the degradation % that 112 Figure 4.26: Oxygen evolution reaction performed for all four solutions was achieved. The authors found that the electrode with the further onset of the OER had a greater PFOS degradation. [43] Table 4.4 provides a summary of the PFOS concentrations, rate constants, and linear fits at 1h and 12h. In these complex solutions, the best electrolyte for treating both PFOA and PFOS are ammonia based solutions. It has also been discovered that there is minimal benefit from running these tests passed the 1 h time point at the higher current density except to gain a continued slow degradation of either PFOA or PFOS depending on the ammonia based solution used. Table 4.4: Comparison of 1 h PFOA degradation and 12 h PFOA degradation 113 4.4.2.3 Byproduct Formation Investigation Since the majority of the IX solutions are comprised of a salt or a base and methanol, it is important to look at the oxidation by-products from these constituents. For example, methanol can be oxidized to for formaldehyde and chloride can be oxidized to perchlorate, both of which are undesirable products. Formaldehyde is classified as a hazardous substance. The calorimetric formaldehyde detection method used was performed by mixing formaldehyde with the Nash reagent. [44] Formaldehyde reacts with the Nash reagent to form diacetyldihydrolutidine which is a yellow compound whose absorbance can be detected at 410 nm using a UV/vis spectrophotometer. The Nash reagent is a mixture of ammonium acetate, acetic acid, acetylacetone and DI water. Once the solutions are mixed they are heated and kept in the dark too cool before analysis with the UV/vis. Using this method solutions containing less than 8 µg of formaldehyde per ml can be analyzed. [44] Those of higher concentrations can be diluted prior to analysis. A standard calibration curve can be made for known concentrations of formaldehyde and then used to determine the value of an unknown sample. Methanol can be oxidized by hydroxyl radicals in the Equations 4.1-4.4 where formaldehyde is generated in Equation 4.3. CH3OH + OH· !· CH2OH + H2O CH3OH + OH· ! CH3O· + H2O ·CH2OH + OH· ! HCHO + H2O HCHO + CH3O· ! HCOOCH3 + H+ + e (4.1) (4.2) (4.3) (4.4) The constant current density study in NH4OH with MeOH and the three combined current density studies in solutions containing MeOH were all investigated for their formaldehyde content at the end of 12 hours. A calibration curve was generated for formaldehyde concentrations of 1 to 8 ppm, shown in Figure 4.27. 114 Figure 4.27: Calibration curve for formaldehyde detection The formaldehyde concentrations were 0 ppm, 134 ppm, 104 ppm, 57 ppm, and 0 ppm for NH4OH with MeOH at 50 mA/cm2, NH4OH with MeOH combined current density, NaCl with MeOH, NaOH with MeOH, and NH4OH with NH4Cl, respectively. The oxidation of chloride to perchlorate is another oxidation byproduct reaction that is un- desirable. Perchlorate generation was analyzed using ion chromatography measurements for the two solutions that contained chloride species, NaCl with MeOH and NH4OH with NH4Cl. Each solution was diluted by 200x and 500x and an average perchlorate concentration was taken as an estimated value for each 12 hour sample. Both solutions had generated 600-700 ppm of perchlorate in 12 hours. This concentration is substantial and well above the water quality standard for per- chlorate. However, there are existing remediation methods that could follow the electrochemical treatment to remove this undesirable product. [45] If the tests in each solution were only ran for 1 h based on achieving the most degradation for both PFOA and PFOS in that time, the byproduct formation may also be significantly reduced and will need to be further evaluated. 4.4.3 Energy Usage and Cost Comparison The overall energy for each PFOA constant current density study and combined current density study of 50&5 was determined using Equation (4.5). 115 Q = (4.5) A ⇤ J v π Edt Where Q is the total energy in kW*h/m3, J is the applied current density in A/cm2, A is the active anodic surface area of the anode in cm2, v is the volume of the solution in m3, and E is the potential of the system in V. The energy usage for 50&5 was calculated as the sum of two integrals. One pertaining to the first time frame with the higher current density and one pertaining to the time with the applied lower current density. The energy value for each simulated sodium sulfate solution study to reach 75% defluorination are provided in Table 4.5. In terms of comparing the constant current densities to the combined current density, we see that 50&5 has a lower energy consumption for 75% defluorination and at 1 h compared to 50 mA/cm2. In addition to consuming similar energy than the process at 25 mA/cm2, 50&5 also only required an additional 0.1 h time. The time to reach 75% defluorination was determined by extrapolating the percentage of PFOA defluorination over time graph for 1, 5, and 10 mA/cm2 on a log scale as they did not reach it in under 1 hour. These graphs are included as Figure 4.28. It is also noted that these measurements should be taken as a rough estimate. Table 4.5: Energy requirements of each PFOA study to reach 75% defluorination or at 1 hour For the 12 hour simulated IX regenerate solutions, the energy was only determined for the two ammonia based solutions as they degraded both PFOA and PFOS successfully. Since one solution obtained the most degradation of PFOA in the first hour and no additional degradation post 1h and the other obtained the most degradation of PFOS in the first hour and no additional decomposition 116 Figure 4.28: Extrapolation plots used to determine the time to 75% defluorination for (A) 1 mA/cm2, (B) 5 mA/cm2, and (C) 10 mA/cm2 after, the 1 h marked was used as the energy comparison time. Since this is the length of the high current density application, both solutions had the same current applied for the duration of this time. However, the two solutions had dierent voltage values. In fact the voltage integral over time for the NH4OH and MeOH was roughly double that of the NH4OH and NH4Cl solution leading to 117 an energy consumption of 106 kW*h/m3 compared to 53 kW*h/m3. To further compare the four IX solutions, the electric energy per order has been determined. The electric energy per order (EEO) is the total energy normalized per log removal of contaminant, Equation 4.6. [46] The initial and final contaminant concentrations are based on the total number of PFAS chains in each solution, not just PFOA or PFOS. EEO = (4.6) Q log(Ci/C f ) Where Ci is the initial total PFAS concentration present in the solution and Cf is the final total PFAS concentration present in the solution. The EEO values at the 12 hour time point for the NH4OH with MeOH, NaOH with MeOH, NaCl with MeOH, and NH4OH with NH4Cl were 362, 569, 2032, and 84 kW*h/m3, respectively. A bar graph for a visual comparison of these values is provided as Figure 4.29. This confirms that the mixed ammonia solution required the lowest energy per log removal of total PFAS compared to the other three solutions. It is also interesting to note that the NaCl and MeOH solution required the most energy per log removal of total PFAS. Figure 4.29: Bar graph comparing the conversion eciency values for the simulated IX regenerate solutions 4.5 Key Advancements from Bench Top PFAS Studies The key advancement from the bench top PFAS studies include the following: • Step down current density techniques do save on energy without inhibiting PFOA degradation 118 in simulated solutions with the combination of 50&5 showing improved energy consumption compared to 50 mA/cm2 • BDD can become fluorinated from the treatment of PFAS, but this has yet to inhibit its ability to remediate PFAS in these studies • PFOA and PFOS can be broken down in dierent complex ion exchange regenerate solutions • The optimum regeneration solutions to be treated with electrochemical oxidation are those that are ammonia based • Byproducts are generated from ion exchange regenerate solution treatments that will need to be investigated further 4.6 Summary Investigation into electrochemical oxidation of PFAS began at the bench scale in a beaker less than 1 L. Initial studies that utilized constant current densities showed that there may be a way to take advantage of the step down current density technique for these compounds even though they do not fit the traditional model. Further investigations in simulated solutions showed that one step down current density method of 50&5 did indeed provide energy savings over a constant current density method. Parameters were slightly adjusted when adapting this procedure to complex simulated ion exchange regenerate solutions, showing that in the first hour of high current density application, most solutions reached a mass transport limit for PFOA or PFOS. Although there were a few exceptions to this rule. Due to the generation of unwanted byproducts, further investigation into limiting their development should be completed. 119 BIBLIOGRAPHY 120 BIBLIOGRAPHY [1] Mohammad Feisal Rahman, Sigrid Peldszus, and William B Anderson. Behaviour and fate of perfluoroalkyl and polyfluoroalkyl substances (pfass) in drinking water treatment: a review. Water research, 50:318–340, 2014. [2] Magali Houde, Jonathan W Martin, Robert J Letcher, Keith R Solomon, and Derek CG Muir. Biological monitoring of polyfluoroalkyl substances: a review. Environmental science technology, 40(11):3463–3473, 2006. [3] Nancy Merino, Yan Qu, Rula A Deeb, Elisabeth L Hawley, Michael R Homann, and Shaily Mahendra. Degradation and removal methods for perfluoroalkyl and polyfluoroalkyl substances in water. Environmental Engineering Science, 33(9):615–649, 2016. [4] Katarzyna H Kucharzyk, Ramona Darlington, Mark Benotti, Rula Deeb, and Elisabeth Haw- ley. Novel treatment technologies for pfas compounds: A critical review. Journal of Environ- mental Management, 2017. [5] Kimberly E Carter and James Farrell. Oxidative destruction of perfluorooctane sulfonate using boron-doped diamond film electrodes. Environmental science technology, 42(16):6111–6115, 2008. [6] C Fang, M Megharaj, and R Naidu. Electrochemical advanced oxidation processes (eaop) to degrade per-and polyfluoroalkyl substances (pfass). Journal of Advanced Oxidation Tech- nologies, 20(2), 2017. [7] Laurel A Schaider, Simona A Balan, Arlene Blum, David Q Andrews, Mark J Strynar, Margaret E Dickinson, David M Lunderberg, Johnsie R Lang, and Graham F Peaslee. Flu- orinated compounds in us fast food packaging. Environmental Science Technology Letters, 4(3):105–111, 2017. [8] Victor Andres Arias Espana, Megharaj Mallavarapu, and Ravi Naidu. Treatment technologies for aqueous perfluorooctanesulfonate (pfos) and perfluorooctanoate (pfoa): A critical review with an emphasis on field testing. Environmental Technology Innovation, 4:168–181, 2015. [9] Marek Trojanowicz, Anna Bojanowska-Czajka, Iwona Bartosiewicz, and Krzysztof Kulisa. Advanced oxidation/reduction processes treatment for aqueous perfluorooctanoate (pfoa) and perfluorooctanesulfonate (pfos)–a review of recent advances. Chemical Engineering Journal, 336:170–199, 2018. [10] Alessandro Zaggia, Lino Conte, Luigi Falletti, Massimo Fant, and Andrea Chiorboli. Use of strong anion exchange resins for the removal of perfluoroalkylated substances from con- taminated drinking water in batch and continuous pilot plants. Water research, 91:137–146, 2016. 121 [11] Lino Contea, Luigi Fallettia, Alessandro Zaggiaa, and Marco Milanb. Polyfluorinated or- ganic micropollutants removal from water by ion exchange and adsorption. CHEMICAL ENGINEERING, 43, 2015. [12] Steve Woodard, John Berry, and Brandon Newman. Ion exchange resin for pfas removal and pilot test comparison to gac. Remediation Journal, 27(3):19–27, 2017. [13] Pattarawan Chularueangaksorn, Shuhei Tanaka, Shigeo Fujii, and Chinagarn Kunacheva. Re- generation and reusability of anion exchange resin used in perfluorooctane sulfonate removal by batch experiments. Journal of Applied Polymer Science, 130(2):884–890, 2013. [14] Shubo Deng, Qiang Yu, Jun Huang, and Gang Yu. Removal of perfluorooctane sulfonate from wastewater by anion exchange resins: eects of resin properties and solution chemistry. Water Research, 44(18):5188–5195, 2010. [15] Ziwen Du, Shubo Deng, Youguang Chen, Bin Wang, Jun Huang, Yujue Wang, and Gang Yu. Removal of perfluorinated carboxylates from washing wastewater of perfluorooctanesulfonyl fluoride using activated carbons and resins. Journal of hazardous materials, 286:136–143, 2015. [16] Gunnar R Stratton, Fei Dai, Christopher L Bellona, Thomas M Holsen, Eric RV Dickenson, and Selma Mededovic Thagard. Plasma-based water treatment: ecient transformation of perfluoroalkyl substances in prepared solutions and contaminated groundwater. Environmen- tal science technology, 51(3):1643–1648, 2017. [17] Raj Kamal Singh, Sujan Fernando, Sadjad Fakouri Baygi, Nicholas Multari, Selma Mede- dovic Thagard, and Thomas M Holsen. Breakdown products from perfluorinated alkyl sub- stances (pfas) degradation in a plasma-based water treatment process. Environmental science technology, 2019. [18] Ch Comninellis and C Pulgarin. Electrochemical oxidation of phenol for wastewater treatment using sno 2, anodes. Journal of Applied Electrochemistry, 23(2):108–112, 1993. [19] Beatriz Gomez-Ruiz, Sonia Gómez-Lavín, Nazely Diban, Virginie Boiteux, Adeline Colin, Xavier Dauchy, and Ane Urtiaga. Ecient electrochemical degradation of poly-and perflu- oroalkyl substances (pfass) from the euents of an industrial wastewater treatment plant. Chemical Engineering Journal, 322:196–204, 2017. [20] Beatriz Gomez-Ruiz, Nazely Diban, and Ane Urtiaga. Comparison of microcrystalline and ultrananocrystalline boron doped diamond anodes: Influence on perfluorooctanoic acid elec- trolysis. Separation and Purification Technology, 2018. [21] Zhaohui Liao and James Farrell. Electrochemical oxidation of perfluorobutane sulfonate using boron-doped diamond film electrodes. Journal of applied electrochemistry, 39(10):1993– 1999, 2009. [22] Tsuyoshi Ochiai, Yuichi Iizuka, Kazuya Nakata, Taketoshi Murakami, Donald A Tryk, Akira Fujishima, Yoshihiro Koide, and Yuko Morito. Ecient electrochemical decomposition of 122 perfluorocarboxylic acids by the use of a boron-doped diamond electrode. Diamond and Related Materials, 20(2):64–67, 2011. [23] Tsuyoshi Ochiai, Hirofumi Moriyama, Kazuya Nakata, Taketoshi Murakami, Yoshihiro Koide, and Akira Fujishima. Electrochemical and photocatalytic decomposition of perfluorooctanoic acid with a hybrid reactor using a boron-doped diamond electrode and tio2 photocatalyst. Chemistry Letters, 40(7):682–683, 2011. [24] Charles E Schaefer, Christina Andaya, Aniela Burant, Charles W Condee, Ana Urtiaga, Timo- thy J Strathmann, and Christopher P Higgins. Electrochemical treatment of perfluorooctanoic acid and perfluorooctane sulfonate: Insights into mechanisms and application to groundwater treatment. Chemical Engineering Journal, 317:424–432, 2017. [25] Álvaro Soriano, Daniel Gorri, and Ane Urtiaga. Ecient treatment of perfluorohexanoic acid by nanofiltration followed by electrochemical degradation of the nf concentrate. Water Research, 112:147–156, 2017. [26] AM Trautmann, H Schell, KR Schmidt, K-M Mangold, and A Tiehm. Electrochemical degradation of perfluoroalkyl and polyfluoroalkyl substances (pfass) in groundwater. Water Science and Technology, 71(10):1569–1575, 2015. [27] Ane Urtiaga, Carolina Fernández-González, Sonia Gómez-Lavín, and Inmaculada Ortiz. Ki- netics of the electrochemical mineralization of perfluorooctanoic acid on ultrananocrystalline boron doped conductive diamond electrodes. Chemosphere, 129:20–26, 2015. [28] Hanshuang Xiao, Baoying Lv, Guohua Zhao, Yujing Wang, Mingfang Li, and Dongming Li. Hydrothermally enhanced electrochemical oxidation of high concentration refractory perfluorooctanoic acid. The Journal of Physical Chemistry A, 115(47):13836–13841, 2011. [29] Qiongfang Zhuo, Shubo Deng, Bo Yang, Jun Huang, Bin Wang, Tingting Zhang, and Gang Yu. Degradation of perfluorinated compounds on a boron-doped diamond electrode. Elec- trochimica Acta, 77:17–22, 2012. [30] Brian P Chaplin. Critical review of electrochemical advanced oxidation processes for water treatment applications. Environmental Science: Processes Impacts, 16(6):1182–1203, 2014. [31] Junfeng Niu, Yang Li, Enxiang Shang, Zesheng Xu, and Jinzi Liu. Electrochemical oxidation of perfluorinated compounds in water. Chemosphere, 146:526–538, 2016. [32] Juyuan Jiang, Ming Chang, and Peng Pan. Simultaneous hydrogen production and elec- trochemical oxidation of organics using boron-doped diamond electrodes. Environmental science technology, 42(8):3059–3063, 2008. [33] Eleni Bitziou, Maxim B Joseph, Tania L Read, Nicola Palmer, Tim Mollart, Mark E Newton, and Julie V Macpherson. In situ optimization of ph for parts-per-billion electrochemical detection of dissolved hydrogen sulfide using boron doped diamond flow electrodes. Analytical chemistry, 86(21):10834–10840, 2014. 123 [34] Bo Guan, Jinfang Zhi, Xintong Zhang, Taketoshi Murakami, and Akira Fujishima. Electro- chemical route for fluorinated modification of boron-doped diamond surface with perfluo- rooctanoic acid. Electrochemistry Communications, 9(12):2817–2821, 2007. [35] Pralay Gayen and Brian P Chaplin. Fluorination of boron-doped diamond film electrodes for minimization of perchlorate formation. ACS Applied Materials Interfaces, 2017. [36] Sergio Ferro and Achille De Battisti. The 5-v window of polarizability of fluorinated diamond electrodes in aqueous solutions. Analytical chemistry, 75(24):7040–7042, 2003. [37] G Sine, L Ouattara, M Panizza, and Ch Comninellis. Electrochemical behavior of fluorinated boron-doped diamond. Electrochemical and solid-state letters, 6(9):D9–D11, 2003. [38] Sergio Ferro and Achille De Battisti. Physicochemical properties of fluorinated diamond electrodes. The Journal of Physical Chemistry B, 107(31):7567–7573, 2003. [39] Shangtao Liang, Randall" David" Pierce Jr, Hui Lin, Sheau-Yun Chiang, and Qingguo" Jack" Huang. Electrochemical oxidation of pfoa and pfos in concentrated waste streams. Remedia- tion Journal, 28(2):127–134, 2018. [40] Hui Lin, Junfeng Niu, Shiyuan Ding, and Lilan Zhang. Electrochemical degradation of perfluorooctanoic acid (pfoa) by ti/sno 2–sb, ti/sno 2–sb/pbo 2 and ti/sno 2–sb/mno 2 anodes. Water research, 46(7):2281–2289, 2012. [41] Qiongfang Zhuo, Xiang Li, Feng Yan, Bo Yang, Shubo Deng, Jun Huang, and Gang Yu. Electrochemical oxidation of 1h, 1h, 2h, 2h-perfluorooctane sulfonic acid (6: 2 fts) on dsa electrode: Operating parameters and mechanism. Journal of Environmental Sciences, 26(8):1733–1739, 2014. [42] Pingping Meng, Shubo Deng, Xinyu Lu, Ziwen Du, Bin Wang, Jun Huang, Yujue Wang, Gang Yu, and Baoshan Xing. Role of air bubbles overlooked in the adsorption of perfluorooctane- sulfonate on hydrophobic carbonaceous adsorbents. Environmental science & technology, 48(23):13785–13792, 2014. [43] Bo Yang, Jinbao Wang, Chaojin Jiang, Juying Li, Gang Yu, Shubo Deng, Shaoyou Lu, Peixin Zhang, Caizhen Zhu, and Qiongfang Zhuo. Electrochemical mineralization of perfluorooctane sulfonate by novel f and sb co-doped ti/sno2 electrode containing sn-sb interlayer. Chemical Engineering Journal, 316:296–304, 2017. [44] Tiany Nash. The colorimetric estimation of formaldehyde by means of the hantzsch reaction. Biochemical Journal, 55(3):416, 1953. [45] Asha Srinivasan and Thiruvenkatachari Viraraghavan. Perchlorate: health eects and tech- nologies for its removal from water resources. International Journal of Environmental Re- search and Public Health, 6(4):1418–1442, 2009. [46] Thi Xuan Huong Le, Holly Haflich, Amisha D Shah, and Brian P Chaplin. Energy-ecient electrochemical oxidation of perfluoroalkyl substances using a ti4o7 reactive electrochemical membrane anode. Environmental Science & Technology Letters, 6(8):504–510, 2019. 124 CHAPTER 5 CONCLUSION 125 5.1 Conclusion This report outlines three important studies that advance the field of electrochemistry. The first uses a boron-doped diamond (BDD) sensor that is comprised of a working, counter, and auxiliary electrode. This sensor is used to demonstrate its ability to detect isatin while eliminating pretreatment procedures that require additional time and additional equipment, simplifying the three electrode set up into one plug and play sensor, and obtaining limits of detection (LOD) applicable to biological fluids. This study also demonstrated the additional benefits of using a microelectrode array (MEA) as the working electrode in the sensor due to its lower sensitivity and more repeatable results. The second study focused on using a material that would allow for economical mass production of sensors for heavy metal detection. Indium tin oxide (ITO) is already used in industry for solar cells and does not require a transparent substrate for the detection of heavy metals, ideally manganese (Mn), using cathodic stripping voltammetry (CSV). Investigating the use of heating during deposition and post deposition annealing led to the discovery that a heated and annealed film had ideal properties to detect Mn in environmentally relevant concentrations due to its larger grain sizes and increased conductivity. This led to the concluding that ITO can be an economic sensor material for such applications. Finally, electrochemical oxidation was investigated to advance the understanding of remediation for per- and polyfluoroalkyl substances (PFAS). Current remediation techniques focus on adsorption or filtration methods that produce a waste stream that can further contribute to the contamination cycle. Using BDD and electrochemical oxidation, a combined current density technique was evaluated using both simulated salt solutions and simulated ion exchange regenerate solutions. Ultimately, the study showed that the two ideal regenerate solutions currently in literature to be further treated with electrochemical oxidation are those based on ammonia, ammonium hydroxide with methanol and the mixed solution of ammonium hydroxide and ammonium chloride. Due to the byproducts produced from these processes, further studies will need to be completed. Overall, these three studies have shown that existing electrochemical methods can be advanced 126 to address problems in today’s world to benefit the planet and humankind. 127 CHAPTER 6 FUTURE OUTLOOK 128 6.1 Future Outlook 6.1.1 Isatin Detection The next steps for the detection of isatin with the 3-in-1 BDD sensors include the detection in a real bodily fluid. Since the sensors were only tested in a urine simulant it is important to study the in a real solution to understand if there are other compounds that can influence the response of isatin. Additionally, a process for renewing the surface of the working electrode should be investigated to allow for improved reproducibilities from day to day studies. Ultimately, these sensors would be used following a liquid chromatography system so a trial with a flow through injection system should also be completed. 6.1.2 Indium Tin Oxide Characteristics Since other metals present in environmental samples have the potential to influence the response for manganese on the HA ITO sensor, the next step in this study would be to run a set of calibration curves with diering levels of other metals present. Once a better understanding is obtained for how the manganese response changes with the presence of other metals, a real environmental sample will need to be studied. In addition, if this sensor is to be used to better understand the toxicology of manganese in the human body, studies with a simulated biological fluid will need to be performed before allowing for the detection in a real biological solution. 6.1.3 Mineralization of Per- and Polyfluoroalkyl Substances The next step in the PFAS remediation studies would be to obtain and characterize an industrial ion exchange regenerate solution. These solutions could have co contaminants that can influence the degradation rates of dierent PFAS chains. They could also be producing dierent byproducts from those that are simulated. It is also important to begin to look into the byproducts that are being generated in the o gasses of the system. This will be critical to know when scaling this technology for industrial use. It can also be of interest to reverse the combined current density 129 configuration to look at the dierence in applying a lower current density before a higher one. This may be beneficial in precursor conversion in industrial solutions. 130