Effect of Urine Compounds on the Electrochemical Oxidation of Urea

Publication Date (Web): June 14, 2018. Copyright © 2018 American ... Outcomes of this study contribute to the development of electrolytic systems for...
0 downloads 0 Views 917KB Size
Subscriber access provided by OXFORD BROOKES UNIVERSITY LIBRARY

Remediation and Control Technologies

Effect of urine compounds on the electrochemical oxidation of urea using a nickel cobaltite catalyst: An electroanalytical and spectroscopic investigation Andrew F. Schranck, Randal Marks, Elon Yates, and Kyle Doudrick Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b01743 • Publication Date (Web): 14 Jun 2018 Downloaded from http://pubs.acs.org on June 15, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27

Environmental Science & Technology

Effect of urine compounds on the electrochemical oxidation of urea using a nickel cobaltite catalyst: An electroanalytical and spectroscopic investigation Authors: Andrew Schrancka, Randal Marksa, Elon Yatesb, Kyle Doudricka* Affiliations: a Department of Civil and Environmental Engineering and Earth Sciences, University of Notre Dame, Notre Dame, Indiana, USA, 46556 b Department of Civil and Environmental Engineering, Florida A&M, Tallahassee, FL, USA, 32310 *

Corresponding Author Address: Kyle Doudrick, Department of Civil and Environmental Engineering and Earth Sciences, University of Notre Dame, 156 Fitzpatrick Hall, Notre Dame, Indiana, USA, 46556 Phone: 5746310305 e-mail: [email protected] JOURNAL: Environmental Science & Technology Date: June 12, 2018

Keywords: Urine, Urea, Nickel, Electrolysis, ATR-FTIR, Wastewater

1

ACS Paragon Plus Environment

Environmental Science & Technology

28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Page 2 of 22

Abstract Cyclic voltammetry (CV) and in-situ attenuated total reflectance-Fourier transform infrared (ATR-FTIR) spectroscopy were used to investigate the effect of major urine compounds on the electrooxidation activity of urea using a nickel cobaltite (NiCo2O4) catalyst. As a substrate, carbon paper exhibited better benchmark potential and current values compared to stainless steel and fluorine-doped tin oxide glass, which was attributed to its greater active surface area per electrode geometric area. CV analysis of a synthetic urine showed that phosphate, creatinine, and gelatin (i.e., proteins) had the most negative effect on the electrooxidation activity of urea with decreases in peak current up to 80% compared to a ureaonly solution. Further investigation on the binding mechanisms of the deleterious compounds using in-situ ATR-FTIR spectroscopy revealed that urea and phosphate weakly bind to NiCo2O4 through hydrogen bonding or long-range forces, whereas creatinine interacts strongly, forming deactivating inner-sphere complexes. Phosphate is presumed to disrupt the interaction between urea and NiCo2O4 by serving as a hydrogen bond acceptor in place of catalyst sites. Weak binding of urea supports the hypothesis that it is oxidized through an indirect electron transfer. Outcomes of this study contribute to the development of electrolytic systems for treating source-separated urine.

2

ACS Paragon Plus Environment

Page 3 of 22

47

Environmental Science & Technology

TOC/ABSTRACT ART

48

3

ACS Paragon Plus Environment

Environmental Science & Technology

49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92

Page 4 of 22

1. INTRODUCTION In municipal wastewater treatment, nitrogenous compounds require additional treatment steps that increase operational complexity and costs. Urine is a significant contributor to this issue as it makes up only 1% of municipal wastewater volume, yet accounts for approximately 80% of wastewater nitrogen.1, 2 This positions decentralized, urine source-separation as a practical solution for reducing nitrogen removal costs.3, 4 Urea (CO2(NH2)2) is the main nitrogen compound in urine, and development of sustainable technologies for its targeted removal will be key to achieving source-separation. Electrochemical systems have emerged as a practical treatment technology for urea, presenting a potentially lower cost alternative to current denitrification methods that are limited by biological process kinetics.5 Other source-separation treatment methods for recovery of N (e.g., precipitation) require increased treatment complexity and infrastructure.1 Microbialbased electrochemical systems show promise, but they involve complex bacterial cultivation, long start-up times, and stringent working conditions.6 With the advent of nanotechnology, inorganic catalysts have emerged as favorable electrode candidates for urea electrooxidation.729 Noble metals (e.g., Pt, Rh, Ru, Ir) were the focus of many early reports,9, 22, 24-28, 30-34 but fullscale application is limited by their high cost and low-abundance. Recent advances in earthabundant catalyst development have shown that nickel, which is also the active metal in the urease enzyme, is a good alternative to noble-metals.7-16, 18, 20, 26, 29, 35-37 While promising, nickel still suffers from limitations including a high overpotential requirement (~1 V)29, 35, 38, 39 and catalyst deactivation by urea oxidation by-products.40-42 For nickel electrodes in alkaline solutions, the consensus has been that urea is oxidized through an indirect electron transfer (i.e., chemical-electrochemical). In this reaction, nickel oxyhydroxide (NiOOH) oxidizes urea while being reduced to nickel hydroxide (Ni(OH)2). Ni(OH)2 is then oxidized back to the active NiOOH via an electron transfer from the electrochemical system. Computational43 and in-situ X-ray diffraction8 evidence has been presented to support this reaction mechanism, and similar behavior has been observed for other organic compounds.44-52 Electrochemical impedance spectroscopy analysis provided contradictory evidence for the reaction mechanism, suggesting the presence of both indirect and direct mechanisms.41 A direct electron transfer indicates that urea forms inner-sphere complexes with the nickel catalyst, and this pathway would be affected by other compounds competing for reactive sites. In support of this, a decrease in current density with progressing treatment time for the electrooxidation of urea using a nickel electrode was observed, and this was attributed to catalyst poisoning or deactivation from urea oxidation by-products.5, 8 Computational analysis of the by-products suggested the adsorption/desorption of cyanate (OCN-) and carbon dioxide (CO2) products were the rate-limiting steps causing catalyst poisoning.43 Less is known about how other compounds present in more realistic matrices (e.g., urine) will affect the urea electrooxidation activity. Though, presumably, select organic compounds and ions will impact reactivity as this has been observed for platinum electrocatalysts.53-58 In this study, we investigated the effect of major urine compounds on the urea electrooxidation activity for nanostructured nickel-based catalysts. The catalysts were synthesized on various electrode substrates, and the electrochemical performance (i.e., potential and current) was investigated using cyclic voltammetry for urea-only and synthetic 4

ACS Paragon Plus Environment

Page 5 of 22

93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135

Environmental Science & Technology

urine solutions. The molecular-scale sorption behavior of select deleterious urine compounds on the catalyst was investigated using in-situ attenuated total reflectance Fourier transform infrared spectroscopy. The outcomes provide support for using electrolysis to remove nitrogen from urine to improve wastewater treatment processes. 2. EXPERIMENTAL 2.1. Electrode Synthesis and Material Characterization. Complete details on the hydrothermal synthesis of nickel oxide (NiO), cobalt oxide (CoxOy), and nickel cobaltite powders (NiCo2O4) and working electrodes can be found in Section S1. Electron micrographs of NiCo2O4 and Ni powders were taken using transmission electron microscopy (TEM, FEI, Titan 80-300). Samples were suspended in isopropanol by sonication and dropcast onto copper TEM grids for analysis. The crystallinity of the powders was determined using selected area electron diffraction (SAED). The morphology of NiCo2O4 and NiO nanostructures grown on electrode supports was determined using a field emission scanning electron microscope (FESEM, FEI, Magellan 400). The oxidation states of both the synthesized powders and the supported catalysts were determined using X-ray photo-electron spectroscopy (XPS, PHI, VersaProbe II). All recorded XPS spectra were normalized to the C 1s emission (284.6 eV) and fit using PHI MultiPak software. 2.2 Electrochemical Measurements. Electrochemical experiments were conducted in a three-electrode reactor (Figure S1). The reactor consisted of a 50-mL beaker (Sigma Aldrich, Z740596-12EA), working electrode, counter electrode, reference electrode, and machined Teflon™ cap with conductive rods (McMaster-Carr, 9100K29) and set screws (McMaster-Carr, 92991A103) to secure the electrodes. The hydrothermally-prepared electrodes served as the working electrodes, Pt on fluorine-doped tin oxide served as the counter electrode, and the reference electrode was either a mercury oxide electrode (Hg/HgO in 20% KOH, Koslow Scientific Company, 5088) for alkaline pH experiments or a saturated calomel electrode (SCE, VWR, 89501-040) for pH 8.7 experiments. The potential was controlled using a Biologic SP200 potentiostat (BioLogic USA), and data were recorded and analyzed using Electrochemical (EC) Lab software. Electrochemical experiments were conducted using various electrolyte conditions. 1 M KOH (Alfa Aesar, A18854) was used for alkaline experiments and 1 M sodium perchlorate (Fisher Scientific, S360) was used for pH 8.7 experiments. The electrolyte reduces solution resistance, minimizes migration effects, and eliminates low conductivity concerns. The reported concentration of urea in human urine varies widely from 11–25 g/L (0.18– 0.42 M),59-63, while 19.82 g/L (0.33 M) is the most commonly used concentration in recent urea electrooxidation studies.9, 12, 26, 35, 42, 64 To align our outcomes with previous studies, we chose to use a urea concentration of 0.33 M. Synthetic urine was prepared as described previously65 but with 19.82 g/L urea instead of 25.0 g/L. Chemicals were used as received and included urea (Amresco, 0568), disodium phosphate (Fisher Scientific, S374-500), monopotassium phosphate (ACS grade Amresco, 0781), sodium chloride (Fisher Scientific, S271-3), ammonium chloride (BDH, BDH9208), creatinine (TCI, C0398), sodium sulfite (Spectrum, 51475), gelatin, (Amresco, 9764), and Difco nutrient broth (BD, 234000). The synthetic urine compounds and corresponding concentrations are listed in Table 1.

5

ACS Paragon Plus Environment

Environmental Science & Technology

136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179

Page 6 of 22

Cyclic voltammograms (CVs) were used to evaluate potential (E (V)) and current density (J (μA/cm2)). Scans were performed at 10 mV/s ranging from -0.1 to 1.242 V vs. the specified reference electrode. Unless otherwise specified, all potentials in this paper are versus the Hg/HgO reference electrode (i.e., normal hydrogen electrode (NHE) + 0.1 V). Ten CV cycles were used to activate the electrode surface and develop steady-state, reproducible scans. The tenth cycle of each experiment is reported herein. CVs can vary slightly from one electrode or solution to the next based on errors in measured area and solution preparation. Experiments were conducted to show the repeatability of our cyclic voltammetry technique (Figure S2). All presented data is for the working electrodes. The onset potential (Eonset) was chosen as the potential at which significant oxidation began (i.e., when the current density curve experienced a positive inflection, as indicated by a positive change in slope of the I-V curve). Because a method for determining Eonset is not well defined in the literature, and often subjective, two other benchmarks were recorded when the previous method was not sufficient for comparing CV data. Specifically, potential values were recorded at 500 µA/cm2, denoted E500, which is similar to an approach used in the electrochemical watersplitting literature.66 E500 was used to serve as an analog to Eonset. Current density values were recorded at 0.65 V, denoted J0.65, which is based on the average potential that the peak anodic current density (Jpa) for urea oxidation occurred before diffusion limitations materialized. 2.3. In-situ Vibrational Spectroscopy. The adsorption behavior of urea and select urine compounds onto NiCo2O4 under no applied potential was analyzed using attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy. Spectra were collected on an infrared spectrophotometer (Bruker, Vertex V70) using a multi-reflection, horizontal ATR cell (HATR, PIKE) equipped with a germanium internal reflectance element (IRE; surface area 6.16 cm2). A liquid nitrogen cooled mercury cadmium telluride (MCT) detector was used for all measurements. The reactor cell was monitored in real-time using Bruker OPUS software with the CHROM package. To make the catalyst film on the IRE, the catalyst was first suspended in water (1 g/L) and then bath sonicated for 30 min. 1 mL of this solution was pipetted evenly onto the surface of the IRE and allowed to dry in air overnight. Prior to analysis, unbound catalyst was removed with a gentle water rinse. Batch adsorption experiments were conducted by first filling the ATR cell with 2.5 mL ultrapure water using a syringe. All subsequent ATR-FTIR experimental solutions contained 1 M potassium hydroxide (KOH, 99.9%, Sigma-Aldrich, 306568) coinciding with electrochemical experiment conditions. After 15 min to allow for equilibration, a background scan was collected (64 scans, 4 cm-1 resolution). The solution was then drained and experimental solutions (i.e., 40 mM urea, 40 mM disodium phosphate, 40 mM creatinine) were subsequently injected into the cell. A total volume of 2.5 mL remained in the cell after injection. The IR spectrum was collected continuously throughout the injection of experimental solutions (64 scans, 4 cm-1), and it was monitored until equilibrium was achieved (~15 min). All experiments conducted in the presence of the catalyst film were replicated in the absence of a film to obtain the solution-phase spectra. Aqueous spectra were collected by the same method over the bare Ge crystal, rather than the NiCo2O4coated crystal. Peak shifts for creatinine were confirmed through replicate experiments for urea and creatinine on NiCo2O4 (Figure S3). 3. RESULTS AND DISCUSSION 6

ACS Paragon Plus Environment

Page 7 of 22

180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196

Environmental Science & Technology

3.1. Materials Characterization of Electrodes. The morphology of NiO (Figure 1A) and NiCo2O4 (Figure 1B-D) consisted of flower-like nanostructures on all substrates, with the presence of small and large clusters (e.g., Figure 1C), similar to that reported previously.37 TEM micrographs (Figure S4A and B) of the catalysts synthesized in the absence of a substrate confirmed the nanosheet morphology of NiO and NiCo2O4, with planar widths of a few hundred nanometers. The SAED pattern of the NiO powder (Figure S4C) matched the (111), (220), and (311) facets of the cubic crystal structure of NiO (JCPDS 04-007-8202). The SAED pattern of the NiCo2O4 powder (Figure S4D) matched the (311) and (440) planes of the cubic crystal structure of NiCo2O4 (JCPDS 00-002-1074). These SAED patterns confirmed the nanosheets are monocrystalline. The chemical composition and oxidation state of the catalysts were determined using XPS analysis (Figure S5). For NiO, a single peak at 855.2 eV (860.8 eV satellite) in the Ni 2p3/2 and at 529.6 eV in the O 1s region matched ranges reported for NiO.67-69 For NiCo2O4, two peaks were observed for both Ni and Co in their respective 2p3/2 regions. For Co, the peak at 780.6 eV and 781.4 eV is attributed to Co(III) and 781.4 Co(II), respectively.70 For Ni, the peak at 854.6 eV and 856.0 eV is attributed to Ni(II) and Ni(III), respectively.71 This mixture of 2+ and 3+ oxidation states on the surface is in agreement with AB2O4 spinel structures.72

197 198 199 200 201 202 203 204 205 206 207

Figure 1. SEM micrographs of (A) NiO on carbon paper, (B) NiCo2O4 on carbon paper, (C) NiCo2O4 on stainless steel, and (D) NiCo2O4 on fluorine doped tin oxide. 3.2. The Effect of Electrode Design and pH on Urea Electrooxidation. The addition of cobalt to nickel electrodes has been shown to improve the electrochemical activity for urea electrooxidation,37, 39 but there is no information on the electrooxidation of urea using cobalt with nickel-based catalysts on carbon paper. Figures 2A and B show the CVs for CoxOy, NiO, and NiCo2O4 catalysts on carbon paper in the absence and presence of urea and a KOH supporting electrolyte. In the absence of urea (Figure 2A), each electrode exhibited a characteristic bump in current density (e.g., NiO Eonset ~ 0.5 V), which was attributed to the 7

ACS Paragon Plus Environment

Environmental Science & Technology

208 209 210 211 212 213 214 215 216 217 218 219 220

221 222 223 224 225 226 227 228 229

Page 8 of 22

oxidation of Ni(II) or Co(II) species as described previously.26, 37, 41-43, 51, 73 The sharp increase in current observed at higher potentials is due to water oxidation. In the presence of urea (Figure 2B), a new peak emerges for NiO and NiCo2O4 and is attributed to urea oxidation.26, 73 Between 0.65 V and 0.70 V, the curve reaches a maximum current density that is twentyfold higher than in the system without urea. Beyond this max, the current density decreases due to diffusion limitations in this unmixed system. Cobalt did not show appreciable activity (i.e., < 50 μA/cm2) for urea oxidation under the conditions tested (Figure 2B) and thus it was excluded from further testing. Regardless of the presence of urea, the Eonset for NiO and NiCo2O4 was between approximately 0.42–0.46 V and 0.36–0.40 V, respectively. Although cobalt was not active for urea oxidation, its addition to NiO (i.e., NiCo2O4) decreased the Eonset and increased the Jpa in all scenarios tested. This has been attributed to shifting electron densities and increased conductivity.74, 75 This moved Eonset to less positive potentials and promoted greater activity, illustrating how cobalt can improve activity while not directly acting as a urea catalyst.

Figure 2. CVs for NiCo2O4, NiO, and CoxOy in 1 M KOH and the (A) absence or (B) presence of 0.33 M urea; (C) FTO, SS, and CP substrates loaded with NiCo2O4 in 1 M KOH with 0.33 M urea; and (D) NiCo2O4 on CP in 1 M KOH (pH = 13.8) and 1 M NaClO4 (pH = 8.7) and 0.33 M urea. The inset in (A) highlights the Co region between 0.45–0.65 V. The choice of catalyst substrate, which varies in surface area and composition, affects the performance of electrochemical systems.6, 38, 76 We investigated the electrochemical activity of 8

ACS Paragon Plus Environment

Page 9 of 22

230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273

Environmental Science & Technology

fluorine-doped tin oxide (FTO), stainless steel (SS), and carbon paper (CP) substrates for urea oxidation. Figure 2C shows CVs for FTO, SS, and CP loaded with NiCo2O4 in the presence of urea. Eonset and Jpa were not useful descriptors for evaluating these data due to the subjective nature of Eonset and the lack of a visible Jpa for urea due to the current increase beyond 0.5 V caused by water oxidation. Thus, we used the two benchmarks described in Section 2.2, E500 and J0.65, and these are presented in Table S3. With no catalyst present, FTO and CP electrodes performed poorly with J0.65 values below 50 µA/cm2 (Figure S6). On the other hand, a J0.65 of 5,540 μA/cm2 was observed for bare SS, which was approximately twice the J0.65 without urea present (Figure S6). This suggests SS can readily oxidize urea without additional catalysts, which is presumably due to the activity of its metals content (Fe:Cr:Ni:Mo = 67.5:17:13:2.5 wt%). The addition of NiCo2O4 onto FTO and CP increased J0.65 over 600 times (Table S3). However, NiCo2O4 addition to SS only showed a 34% increase in J0.65. (Figure 2C). The increase in electrode activity for NiCo2O4 is attributed to urea-specific reactive sites, a reduction in the overpotential for urea oxidation, and an increased available surface area due to nanostructuring. The range in E500 values was only 0.061 V for substrates loaded with NiCo2O4, meaning the substrates had little effect on Eonset. The potential at which chemical reactions begin to occur (i.e., E500), ideally should not change based on surface area, but the current density or reactivity will be affected. The superior J0.65 of the CP electrode (19,120 μA/cm2) was attributed to the greater total catalyst surface area per geometric surface area. In this study, we report current as a function of the electrode geometric surface area, i.e., the product of the length and width of the submerged electrode. This assumes the substrates have flat surfaces when in reality they have a three-dimensional structure. The FTO has a relatively flat, rough surface, the SS is a mesh, and the CP is a complex fiber network (Figure S8). The fiber structure of the CP gives it a higher total catalyst surface area per geometric area, thus it has a larger number of sites compared to equally-sized FTO and SS electrodes. While mechanistically normalizing current to the total catalyst surface area would seem ideal, reporting current as a function of the geometric area is more appropriate for realistic purposes that seek to minimize cost and footprint. The pH of an electrochemical system plays a significant role in determining electrode stability and activity, and its deviation from circumneutral affects the cost of water treatment. However, alkaline conditions are desirable for the system tested because they inhibit the production of unwanted ammonia formation by urease while also promoting urea oxidation using nickel-based electrodes. Here we compare pH 8.7 and alkaline (i.e., 13.8) pH conditions to reinforce the requirement that OH- is needed to activate the Ni(II)/Ni(III) redox mechanism for urea electrooxidation26 (Figure 2D). The J0.65 for pH 8.7 conditions in the absence and presence of urea were the same (~8 μA/cm2, Figure S9), indicating urea was not oxidized. For pH 13.8 conditions, the addition of urea resulted in a substantially higher J0.65 (19,120 μA/cm2). This inactivity under pH 8.7 conditions across the range of applied potential reinforces the need for alkaline electrolyte conditions for nickel-based electrodes. The primary reason for the performance disparity between pH conditions was attributed to the need of hydroxide to form urea-active NiOOH.51, 52 A similar pH-catalytic requirement has been observed for water oxidation.77 A secondary and related explanation for the reduced activity under pH 8.7 conditions is the increased thermodynamic barrier for oxidation reactions at lower pH as described by the Nernst equation (E = ESRP – 0.059 x pH; E = pH adjusted 9

ACS Paragon Plus Environment

Environmental Science & Technology

274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317

Page 10 of 22

reduction potential, ESRP = standard reduction potential).78 The addition of OH- to the wastewater stream increases cost, complexity, and potential hazards to the treatment process, but it is required for this treatment technology. However, chemical addition is a common and accepted practice in many drinking and wastewater processes (e.g., coagulants, disinfectants), and the cost of simple electrolytes is comparable. 3.3. The Effect Urine Compounds on Urea Electrooxidation. Though urea is the primary compound in urine, other urine constituents could impact the electrode surface chemistry and electrochemical performance. The current-potential behavior of the NiCo2O4 catalyst on carbon paper was investigated for a nine-compound synthetic urine (Table S1). Additionally, each urine compound was investigated separately to determine the individual effect on urea electrooxidation. The CVs of the forward potential sweeps for these experiments are shown in Figure 3A, and the J0.65 values obtained from these are summarized in Table 1. The E500 (Figure 3B) varied minimally for all samples, ranging from 0.400 V to 0.439 V. This tight range of potential variation suggested the urine compounds did not have a major impact on Eonset. Conversely, the current density, which is an indicator of the electrochemical reaction rate, varied for all samples. The J0.65 for the synthetic urine mixture (1152 μA/cm2) was approximately three times lower than urea only samples (3145 μA/cm2), indicating one or more of the compounds had a negative impact on the urea oxidation rate. Ammonium chloride had the highest J0.65 (5154 μA/cm2), which was almost twice that of urea. This phenomenon was attributed to ammonium and not chloride, as a similar effect was not observed for sodium chloride. Under alkaline conditions, the ammonium ion will be deprotonated (pKa = 9.24) as aqueous ammonia. Previous CV analyses of ammonia have consistently shown a peak at ~0.65 V (vs. RHE),79 which overlaps with urea and thus explains why the J0.65 is greater than urea only. All compounds besides ammonia negatively impacted urea electrooxidation to varying degrees. Difco broth, NaCl, and Na2SO3 had a minor impact on the overall urea electroactivity with decreased Jpa values in the urea electrooxidation window (i.e., 0.45 to 0.70 V) of less than 30%. On the other hand, Na2HPO4, KH2PO4, creatinine, and gelatin had a more significant impact on urea electrooxidation (i.e., 0.45 to 0.70 V), with observed J0.65 values of 1316, 1318, 2176, and 512 μA/cm2, respectively, and decreased Jpa values up to ~80%. The sodium and potassium phosphate salts performed similarly to one another, with overlapping CVs between 0.45 and 0.70 V, indicating PO43-, and not Na+ or K+, was impeding urea electrooxidation through interactions with the catalyst surface and/or weak force interactions in the Stern layer. Creatinine (i.e., C4H7N3O) consists of a five-side aromatic ring with an amino and a methyl group, and these groups could potentially interact with the NiCo2O4 through weak hydrogen bonding or stronger inner-sphere complexes. Most compounds had similar current-potential behavior as urea, which includes onset potentials, peak potentials, and diffusion limitations. However, creatinine did not follow this trend; instead, a steady increase in current density with increasing applied potential was observed. This is presumably due to creatinine adsorbing to NiCo2O4 active sites (i.e., non-Faradaic current) followed by oxidation of creatinine preferentially over urea. Gelatin, which is a complex animal protein mixture, behaved similarly to creatinine, indicating strong sorption interactions with the catalyst. Due to its similar CV behavior as synthetic urine, gelatin was most likely the underlying cause of synthetic urine’s decreased activity compared to urea alone. The greater J0.65 observed for synthetic urine over 10

ACS Paragon Plus Environment

Page 11 of 22

Environmental Science & Technology

318 319 320 321 322 323 324 325 326 327

gelatin was due to the additive Faradaic and non-Faradaic current obtained from active compounds (e.g., ammonia). Like gelatin, Difco broth is a heterogeneous mixture of organics (e.g., proteins), but it had only a minor impact on urea electrooxidation. This difference was attributed to the lower concentration of Difco broth (0.00016 g/L) compared to gelatin (1 g/L), which would lower its impact. To further confirm the detrimental effect of the synthetic urine components, two-hour constant voltage experiments were conducted with urea and synthetic urine at 0.65 V. Results showed a decreased current density for synthetic urine compared to the urea only experiment over the course of two hours (Figure S10). Given the likelihood of phosphate and creatinine to interfere with urea electrooxidation, further investigation was required of them to test the stated hypothesis.

328 329 330 331 332 333 334

Figure 3. CV anodic sweeps of (A) synthetic urine compounds and highlighted regions detailing (B) E500 and (C) J0.65 values for select urine compounds. All solutions contained 1 M KOH. The potassium hydroxide curve represents “no urea” control, and all other solutions contained 0.33 M urea. All experiments were conducted with NiCo2O4 loaded onto carbon paper.

11

ACS Paragon Plus Environment

Environmental Science & Technology

335 336

337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365

Page 12 of 22

Table 1. Synthetic urine compound concentrations and their J0.65 for NiCo2O4 on CP. All solutions contained 1 M KOH. All solutions contained 0.33 M urea. Concentration J0.65 Compound (g/L) (μA/cm2) Synthetic urine 1152 Urea 19.82 3145 Ammonium chloride 3 5154 Sodium sulfite 3 2926 -4 Difco broth 1.6x10 2205 Creatinine 2 2176 Sodium chloride 9 1950 Monopotassium phosphate 2.5 1318 Disodium phosphate 2.5 1316 Gelatin 1 512 3.4. In-Situ ATR-FTIR Analysis of Synthetic Urine Compound Adsorption onto NiCo2O4. In-situ ATR-FTIR was used to investigate the individual adsorption behavior of urea, phosphate, and creatinine on NiCo2O4, and the competitive adsorption behavior of urea and phosphate or creatinine on NiCo2O4. The FTIR spectra of urea in the absence of NiCo2O4 (i.e., solution-phase) agreed with previously reported major bands for urea80, 81 (Figure 4A; Table 2). Deconvolution identified three peaks between ~1700 to 1550 cm-1. The peak at 1662 cm-1 is mostly due to the in-plane symmetric NH2 bending with contributions from CO and CN stretching (i.e., δs(NH2) + ν(CO) + νs(CN)). The peak at 1632 cm-1 is due to out-of-plane asymmetric NH2 bending with contribution from symmetric CN stretching (i.e., δas(NH2) + νs(CN)). The peak at 1604 cm-1 is due to CO stretching with contributions from symmetric NH2 stretches (i.e., ν(CO) + δs(NH2) + ρ(NH2)). The peak at 1466 cm-1 is due to asymmetric CN stretching with contribution from asymmetric NH2 bending and CO bending (i.e., νas(CN) + δas(NH2) + δ(CO)). The peak at 1160 cm-1 is due to in-plane, asymmetric NH2 bending (i.e., ρ(NH2)). Urea has a resonant structure with double bond character shifting between CO and CN, and bonding of the N or O to a catalyst will shift resonance to more single bond character for CN and CO vibrations, respectively.24, 25, 34 This change in bond character upon adsorption to the catalyst would result in red- or blue- shifts for CN and CO vibrations. In the presence of NiCo2O4 (Figure 4B), only minor shifts (-3 to +3 cm-1) were observed for all peaks compared to the solution-phase peak positions. The lack of new peak development, peak splitting, or major shifts indicates that the interaction between urea and NiCo2O4 is presumably due to hydrogen bonding or long-range forces (i.e., outer-sphere complexes), and not complexation between urea N or O and the catalyst (i.e., inner-sphere complexes). However, changes in the relative absorption intensities between peaks were observed, mostly with respect to the CO stretching and NH2 bending modes. For example, the ratio between and δas(NH2) and ν(CO) increased from 0.59 to 0.85 upon interaction with NiCo2O4. This increased NH2 character is presumably due to orientation of the amide groups toward the catalyst through hydrogen bonding. This supports the hypothesized mechanism that NiOOH oxidizes urea through an indirect electrochemical-chemical pathway.41, 42, 73

12

ACS Paragon Plus Environment

Page 13 of 22

366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390

Environmental Science & Technology

Phosphate can form inner-sphere complexes with metal (hydr)oxide surfaces,82, 83 thus it was hypothesized that the decreased current we observed in the presence of phosphate was due to site blocking. However, ATR-FTIR analysis showed that phosphate did not interact strongly with NiCo2O4. At pH 13.8, the phosphate ion was fully deprotonated (PO43-), and its solutionphase spectra matched its spectra in the presence of NiCo2O4 (Figure S11), indicating weak if any interaction occurred. The only notable difference in the spectra was a minor loss of intensity for the 1007 cm-1 peak. This can be attributed to the repulsion of the negative PO43from the NiCo2O4 surface, which is also negatively charged at alkaline pH (e.g., pHzpc = 3.5).84 Similar adsorption behavior was previously observed for TiO2.85 These results suggest there is no meaningful interaction between phosphate and NiCo2O4. Though phosphate did not form strong complexes with NiCo2O4, it did affect the adsorption behavior of urea (Figure 4C). In the presence of phosphate, only minor peak shifts in the urea peaks were observed. However, significant variations in the peak intensities (e.g., 1467 and 1165 cm-1) occurred along with slight distortion of the peak at 1662 cm-1. These anomalies, all of which involve the urea amide groups, suggest there was a change in the adsorption behavior of these functional groups when phosphate was introduced. A few possible interactions can be hypothesized. Urea can act as a hydrogen donor for hydrogen bonding between oxygens on the phosphate molecule (hydrogen acceptor) and the urea amide groups.86, 87 In the presence of phosphate and catalyst, the ratio of δas(NH2) and ν(CO) increased from 0.59 to1.35, which suggests further orientation of the amide groups toward the catalyst through hydrogen bonding between urea and phosphate. This could lessen the interaction between urea and NiCo2O4 or otherwise hinder electrochemical oxidation on the electrode surface. Alternatively, phosphate could be electrostatically blocking urea from interacting with NiCo2O4 in the Stern layer.

13

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 22

391 392 393 394 395 396 397 398

Figure 4. In-situ ATR-FTIR spectra of 40 mM urea (A) solution-phase, (B) in the presence of a NiCo2O4 particulate film, and (C) with NiCo2O4 and 40 mM of phosphate. All experiments were conducted with 1 M KOH to mimic electrochemical pH conditions. Spectra are offset for clarity. Table 2. Urea peak locations (in cm-1) observed in this study, and a comparison to those observed or calculated previously. The bold assignments represent the major vibration. Solution

On

On NiCo2O4 and

Reported

Phase

NiCo2O4

with Phosphate

Values80, 81

δs(NH2) + ν(CO) + νs(CN)

1662

1660

1664

1663–1668

δas(NH2) + νs(CN)

1632

1629

1631

1629–1636

Peak Assignment81

399 400 401 402 403 404 405 406 407 408 409 410 411

ν(CO) + δs(NH2) + ρ(NH2)

1604

1601

1599

1597–1602

νas(CN) + δas(NH2) + δ(CO)

1466

1464

1467

1463–1473

ρ(NH2)

1160

1160

1165

1155–1160

Contrasting to urea and phosphate, creatinine, which had a complex solution-phase spectrum (Figure 5A), showed strong and complex adsorption behavior with NiCo2O4 (Figure 5B). In the presence of urea, there were numerous overlapping peaks that occurred (Figure 5C; Table S5). To date, the peak-structure relationships of the aqueous creatinine spectra have not been fully evaluated. Hydrogen bonding in crystalline creatinine creates an extremely rich spectrum that does not match the solvated creatinine species found in aqueous conditions.90 Further, relevant ATR-FTIR studies have explored the spectra only at biologically relevant pH88, 89 and not alkaline conditions (i.e., pH = 13.8) used in our study. At pH 13.8, aqueous creatinine will predominantly exist in deprotonated form (pKas = ~3.8, 12.7), the structure of which has not yet been fully determined.90, 91 For our observed spectra, peak deconvolution was not reliable due to the complexity and the numerous possible fitting outcomes. Therefore, the peaks observed in this study have not been positively identified, and a full evaluation is 14

ACS Paragon Plus Environment

Page 15 of 22

412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429

430 431 432 433 434 435 436 437 438 439 440 441 442

Environmental Science & Technology

beyond the scope of the study. However, we did observe significant and obvious shifts in major peak locations and intensity along with the formation of new peaks when creatinine was analyzed in the presence of NiCo2O4 (Figure 5B). The large, complex peak cluster between 1700–1500 cm-1 is difficult to interpret, but new peaks at 1421 cm-1 and 1309 cm-1 were unique to creatinine in the presence of NiCo2O4, indicating the formation of strong, inner-sphere complexes. Upon adsorption to NiCo2O4 the solution-phase peaks at 1485 cm-1, 1462 cm-1, and 1421 cm-1 all disappear, while the peak at 1632 cm-1 is shifted to lower wavelengths (Figure 5A). These vibrations are associated with either CN or NH2, suggesting the binding occurs at the amino group and/or the heterocyclic N. Therefore, when creatinine and urea are both present, creatinine will interact preferentially with the catalyst due to its strong bonding, inhibiting or blocking urea due to its weaker interaction with the catalyst. This behavior is consistent with our electrochemical observations, which showed the presence of creatinine decreased the current generated in the urea oxidation potential window. Unlike phosphate, whose removal is not probable due to unfavorable oxidation thermodynamics, creatinine will presumably be oxidized at higher potentials, and thus removed simultaneously with urea. Increased current density at potentials greater than 0.65 V for creatinine-urea solutions compared to urea-only solutions (i.e., Figure 3A) supports this hypothesis.

Figure 5. In-situ ATR-FTIR spectra of 40 mM creatinine (A) solution-phase, (B) on NiCo2O4, and (C) on NiCo2O4 and with 40 mM urea. All experiments conducted in 1 M KOH to mimic electrochemical pH conditions. Spectra are offset for clarity. See Table S5 for peak identification. 3.5. Implications for Source-Separated Urine Treatment. These electrochemical and insitu vibrational spectroscopy analyses using a synthetic urine have revealed new information about the interaction between nickel-based electrodes and urine compounds. Phosphate, creatinine, and gelatin were major inhibitors of urea electrooxidation, hypothesized to disrupt urea oxidation through different mechanisms in the Stern Layer (Figure 6). This insight into the solid-liquid interface will be useful for furthering catalyst development. While sorption knowledge presented herein is useful, future studies should investigate the impact of applied 15

ACS Paragon Plus Environment

Environmental Science & Technology

443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459

460 461 462 463 464

Page 16 of 22

potential as changes in electron density at the catalyst surface will shift sorption behavior, as has been observed for platinum and rhodium electrodes.22, 24, 34 The high activity of carbon paper as a substrate makes it conducive for development of flow cell technologies that use membrane electrode assemblies.15, 16, 38, 92 Solution pH after urea electrooxidation remains a challenge for nickel-based electrode systems. Effluents sent to municipal treatment would not drastically impact the pH of these systems due to the low volume of urine compared to total wastewater volume. However, effluents should be neutralized to prevent transportation hazards and infrastructure damage. Depending on the goal of an electrochemical system, non-urea compounds may or may not need to be treated. The initial application of these electrolysis systems will be for N removal (and H2 generation) to reduce N burden on municipal wastewater treatment plants. Similarly, recovering direct energy from urine using a urea fuel cells is a possible early application.26, 38 In either case, downstream treatment would be required for other compounds, though a single system to achieve complete treatment is a future goal. With continued development and research focused on improving current generation while minimizing negative effects from urine compounds, electrolytic cells represent an attractive option for sustainably treating sourceseparated urine.

Figure 6. Conceptual schematic of the solid-solution interface, including interactions between NiCo2O4, urea, phosphate, and creatinine. The 0, 1, and 2 indicate the Stern plane numbers.

16

ACS Paragon Plus Environment

Page 17 of 22

465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480

Environmental Science & Technology

ACKNOWLEDGEMENTS This work was made possible through financial support from Dr. Doudrick’s startup funds provided by the University of Notre Dame, and Andrew Schranck’s scholarship funding from the Patrick and Jana Eilers Graduate Student Fellowship for Energy Related Research (Center for Sustainable Energy at Notre Dame (ND Energy)), the Rothblatt Memorial Scholarship (Lake Michigan States Section of the Air and Waste Management Association (A&WMA)), and the Post Graduate Scholarship (National Collegiate Athletic Association (NCAA)). The authors thank Dr. Ian Lightcap of the Materials Characterization Facility (MCF) for aiding with the PHI VersaProbe II system. The MCF is funded by the Sustainable Energy Initiative (SEI), which is part of ND Energy. The authors thank Drs. Tatyana Orlova and Sergei Rouvimov of the Notre Dame Integrated Imaging Facility for assistance with SEM, TEM, and SAED analysis. The authors thank Rob Roberts of BioLogic for technical support with electrochemical experiments, Brent Bray and Leon Hluchota for help machinging electrochemical apparatus parts, and undergraduate students David Clark and Juan Velazquez for experimental assistance.

17

ACS Paragon Plus Environment

Environmental Science & Technology

481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524

Page 18 of 22

REFERENCES 1.

Maurer, M.; Pronk, W.; Larsen, T. A., Treatment processes for source-separated urine. Water Research 2006,

40, (17), 3151-3166. 2.

Larsen, T. A.; Alder, A. C.; Eggen, R. I. L.; Maurer, M.; Lienert, J., Source Separation: Will We See a

Paradigm Shift in Wastewater Handling? Environmental Science & Technology 2009, 43, (16), 6121-6125. 3.

Lienert, J.; Larsen, T. A., High Acceptance of Urine Source Separation in Seven European Countries: A

Review. Environ. Sci. Technol. 2010, 44, (2), 556-566. 4.

Ledezma, P.; Kuntke, P.; Buisman, C. J. N.; Keller, J.; Freguia, S., Source-separated urine opens golden

opportunities for microbial electrochemical technologies. Trends in Biotechnology 2015, 33, (4), 214-220. 5.

Yan, W.; Wang, D.; Botte, G. G., Electrochemical decomposition of urea with Ni-based catalysts. Appl. Catal.

B-Environ. 2012, 127, 221-226. 6.

Logan, B. E.; Hamelers, B.; Rozendal, R. A.; Schrorder, U.; Keller, J.; Freguia, S.; Aelterman, P.; Verstraete,

W.; Rabaey, K., Microbial fuel cells: Methodology and technology. Environmental Science & Technology 2006, 40, (17), 5181-5192. 7.

Guo, F.; Ye, K.; Cheng, K.; Wang, G. L.; Cao, D. X., Preparation of nickel nanowire arrays electrode for urea

electrooxidation in alkaline medium. J. Power Sources 2015, 278, 562-568. 8.

Wang, D.; Botte, G. G., In Situ X-Ray Diffraction Study of Urea Electrolysis on Nickel Catalysts. Ecs

Electrochemistry Letters 2014, 3, (9), H29-H32. 9.

Miller, A. T.; Hassler, B. L.; Botte, G. G., Rhodium electrodeposition on nickel electrodes used for urea

electrolysis. Journal of Applied Electrochemistry 2012, 42, (11), 925-934. 10. Wu, M. S.; Ji, R. Y.; Zheng, Y. R., Nickel hydroxide electrode with a monolayer of nanocup arrays as an effective electrocatalyst for enhanced electrolysis of urea. Electrochim. Acta 2014, 144, 194-199. 11. Wu, M. S.; Lin, G. W.; Yang, R. S., Hydrothermal growth of vertically-aligned ordered mesoporous nickel oxide nanosheets on three-dimensional nickel framework for electrocatalytic oxidation of urea in alkaline medium. J. Power Sources 2014, 272, 711-718. 12. Wang, D.; Yan, W.; Vijapur, S. H.; Botte, G. G., Enhanced electrocatalytic oxidation of urea based on nickel hydroxide nanoribbons. J. Power Sources 2012, 217, 498-502. 13. Ye, K.; Zhang, D. M.; Guo, F.; Cheng, K.; Wang, G. L.; Cao, D. X., Highly porous nickel@carbon sponge as a novel type of three-dimensional anode with low cost for high catalytic performance of urea electro-oxidation in alkaline medium. Journal of Power Sources 2015, 283, 408-415. 14. Wang, D.; Yan, W.; Vijapur, S. H.; Botte, G. G., Electrochemically reduced graphene oxide-nickel nanocomposites for urea electrolysis. Electrochim. Acta 2013, 89, 732-736. 15. Lan, R.; Tao, S. W.; Irvine, J. T. S., A direct urea fuel cell - power from fertiliser and waste. Energy & Environmental Science 2010, 3, (4), 438-441. 16. Lan, R.; Tao, S., Preparation of nano-sized nickel as anode catalyst for direct urea and urine fuel cells. Journal of Power Sources 2011, 196, (11), 5021-5026. 17. Xu, W.; Zhang, H.; Li, G.; Wu, Z., Nickel-cobalt bimetallic anode catalysts for direct urea fuel cell. Scientific Reports 2014, 4, 1-6. 18. Xu, W.; Zhang, H. M.; Li, G.; Wu, Z. C., A urine/Cr(VI) fuel cell - Electrical power from processing heavy metal and human urine. Journal of Electroanalytical Chemistry 2016, 764, 38-44. 19. Yu, B.; Zhang, H.; Xu, W.; Li, G.; Wu, Z., Remediation of chromium-slag leakage with electricity cogeneration via a urea-Cr(VI) cell. Scientific Reports 2014, 4, 1-6. 20. Wang, G.; Ling, Y.; Lu, X.; Wang, H.; Qian, F.; Tong, Y.; Li, Y., Solar driven hydrogen releasing from urea and human urine. Energy & Environmental Science 2012, 5, (8), 8215-8219. 18

ACS Paragon Plus Environment

Page 19 of 22

525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568

Environmental Science & Technology

21. Urbanczyk, E.; Jaron, A.; Simka, W., Electrocatalytic oxidation of urea on a sintered Ni-Pt electrode. Journal of Applied Electrochemistry 2017, 47, (1), 133-138. 22. Bezerra, A. C. S.; deSa, E. L.; Nart, F. C., In situ vibrational study of the initial steps during urea electrochemical oxidation. J. Phys. Chem. B 1997, 101, (33), 6443-6449. 23. Cho, K.; Hoffmann, M. R., Molecular hydrogen production from wastewater electrolysis cell with multijunction BiOx/TiO2 anode and stainless steel cathode: Current and energy efficiency. Applied Catalysis BEnvironmental 2017, 202, 671-682. 24. Climent, V.; Rodes, A.; Perez, J. M.; Feliu, J. M.; Aldaz, A., Urea adsorption at rhodium single-crystal electrodes. Langmuir 2000, 16, (26), 10376-10384. 25. Climent, V.; Rodes, A.; Orts, J. M.; Aldaz, A.; Feliu, J. M., Urea adsorption on Pt(111) electrodes. J. Electroanal. Chem. 1999, 461, (1-2), 65-75. 26. Boggs, B. K.; King, R. L.; Botte, G. G., Urea electrolysis: direct hydrogen production from urine. Chemical Communications 2009, (32), 4859-4861. 27. King, R. L.; Botte, G. G., Investigation of multi-metal catalysts for stable hydrogen production via urea electrolysis. J. Power Sources 2011, 196, (22), 9579-9584. 28. Chen, S.; Duan, J.; Vasileff, A.; Qiao, S. Z., Size Fractionation of Two-Dimensional Sub-Nanometer Thin Manganese Dioxide Crystals towards Superior Urea Electrocatalytic Conversion. Angewandte ChemieInternational Edition 2016, 55, (11), 3804-3808. 29. Ding, R.; Li, X.; Shi, W.; Xu, Q.; Wang, L.; Jiang, H.; Yang, Z.; Liu, E., Mesoporous Ni-P nanocatalysts for alkaline urea electrooxidation. Electrochimica Acta 2016, 222, 455-462. 30. Yao, S. J.; Wolfson, S. K.; Ahn, B. K.; Liu, C. C., ANODIC-OXIDATION OF UREA AND AN ELECTROCHEMICAL APPROACH TO DE-UREATION. Nature 1973, 241, (5390), 471-472. 31. Yao, S. J.; Wolfson, S. K. J.; Tokarsky, J. M.; Ahn, B. K., DEUREATION BY ELECTROCHEMICAL OXIDATION. Bioelectrochemistry and Bioenergetics 1974, 1, (1/2), 180-186. 32. Wright, J. C.; Michaels, A. S.; Appleby, A. J., ELECTROOXIDATION OF UREA AT THE RUTHENIUM TITANIUM-OXIDE ELECTRODE. Aiche Journal 1986, 32, (9), 1450-1458. 33. Amstutz, V.; Katsaounis, A.; Kapalka, A.; Comninellis, C.; Udert, K. M., Effects of carbonate on the electrolytic removal of ammonia and urea from urine with thermally prepared IrO2 electrodes. J. Appl. Electrochem. 2012, 42, (9), 787-795. 34. Climent, V.; Rodes, A.; Orts, J. M.; Feliu, J. M.; Perez, J. M.; Aldaz, A., On the electrochemical and in-situ Fourier transform infrared spectroscopy characterization of urea adlayers at Pt(100) electrodes. Langmuir 1997, 13, (8), 2380-2389. 35. Xu, W.; Zhang, H.; Li, G.; Wu, Z., Nickel-cobalt bimetallic anode catalysts for direct urea fuel cell. Scientific Reports 2014, 4. 36. Yu, B.; Zhang, H.; Xu, W.; Li, G.; Wu, Z., Remediation of chromium-slag leakage with electricity cogeneration via a urea-Cr(VI) cell. Scientific Reports 2014, 4. 37. Wang, D.; Vijapur, S. H.; Wang, Y.; Botte, G. G., NiCo2O4 nanosheets grown on current collectors as binderfree electrodes for hydrogen production via urea electrolysis. International Journal of Hydrogen Energy 2017, 42, (7), 3987-3993. 38. Xu, W.; Wu, Z. C.; Tao, S. W., Urea-Based Fuel Cells and Electrocatalysts for Urea Oxidation. Energy Technology 2016, 4, (11), 1329-1337. 39. Yan, W.; Wang, D.; Botte, G. G., Nickel and cobalt bimetallic hydroxide catalysts for urea electro-oxidation. Electrochimica Acta 2012, 61, 25-30. 40. Forslund, R. P.; Mefford, J. T.; Hardin, W. G.; Alexander, C. T.; Johnston, K. P.; Stevenson, K. J., 19

ACS Paragon Plus Environment

Environmental Science & Technology

569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612

Page 20 of 22

Nanostructured LaNiO3 Perovskite Electrocatalyst for Enhanced Urea Oxidation. Acs Catalysis 2016, 6, (8), 5044-5051. 41. Guo, F.; Ye, K.; Du, M. M.; Huang, X. M.; Cheng, K.; Wang, G. L.; Cao, D. X., Electrochemical impedance analysis of urea electro-oxidation mechanism on nickel catalyst in alkaline medium. Electrochim. Acta 2016, 210, 474-482. 42. Vedharathinam, V.; Botte, G. G., Direct evidence of the mechanism for the electro-oxidation of urea on Ni(OH)(2) catalyst in alkaline medium. Electrochimica Acta 2013, 108, 660-665. 43. Daramola, D. A.; Singh, D.; Botte, G. G., Dissociation Rates of Urea in the Presence of NiOOH Catalyst: A DFT Analysis. Journal of Physical Chemistry A 2010, 114, (43), 11513-11521. 44. Barbosa, A. F. B.; Oliveira, V. L.; van Drunen, J.; Tremiliosi-Filho, G., Ethanol electro-oxidation reaction using a polycrystalline nickel electrode in alkaline media: Temperature influence and reaction mechanism. J. Electroanal. Chem. 2015, 746, 31-38. 45. Motheo, A. J.; Machado, S. A. S.; Rabelo, F. J. B.; Santos, J. R., ELECTROCHEMICAL STUDY OF ETHANOL OXIDATION ON NICKEL IN ALKALINE MEDIA. Journal of the Brazilian Chemical Society 1994, 5, (3), 161-165. 46. Rahim, M. A. A.; Hameed, R. M. A.; Khalil, M. W., Nickel as a catalyst for the electro-oxidation of methanol in alkaline medium. J. Power Sources 2004, 134, (2), 160-169. 47. Danaee, I.; Jafarian, M.; Forouzandeh, F.; Gobal, F.; Mahjani, M. G., Electrocatalytic oxidation of methanol on Ni and NiCu alloy modified glassy carbon electrode. Int. J. Hydrog. Energy 2008, 33, (16), 4367-4376. 48. Kapalka, A.; Cally, A.; Neodo, S.; Comninellis, C.; Wachter, M.; Udert, K. M., Electrochemical behavior of ammonia at Ni/Ni(OH)(2) electrode. Electrochem. Commun. 2010, 12, (1), 18-21. 49. van Drunen, J.; Napporn, T. W.; Kokoh, B.; Jerkiewicz, G., Electrochemical oxidation of isopropanol using a nickel foam electrode. J. Electroanal. Chem. 2014, 716, 120-128. 50. Kapalka, A.; Joss, L.; Anglada, A.; Comninellis, C.; Udert, K. M., Direct and mediated electrochemical oxidation of ammonia on boron-doped diamond electrode. Electrochem. Commun. 2010, 12, (12), 1714-1717. 51. Miao, Y.; Ouyang, L.; Zhou, S.; Xu, L.; Yang, Z.; Xiao, M.; Ouyang, R., Electrocatalysis and electroanalysis of nickel, its oxides, hydroxides and oxyhydroxides toward small molecules. Biosensors & Bioelectronics 2014, 53, 428-439. 52. Fleischmann, M.; Korinek, K.; Fletcher, D., The Oxidation of Hydrazine at a Nickel Anode in Alkaline Solution. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1972, 34, 499-503. 53. Ordonez, S.; Diez, F. V.; Sastre, H., Characterisation of the deactivation of platinum and palladium supported on activated carbon used as hydrodechlorination catalysts. Applied Catalysis B-Environmental 2001, 31, (2), 113122. 54. Patzer, J. F.; Wolfson, S. K.; Yao, S. J., REACTOR CONTROL AND REACTION-KINETICS FOR ELECTROCHEMICAL UREA OXIDATION. Chemical Engineering Science 1990, 45, (8), 2777-2784. 55. Tokash, J. C.; Logan, B. E., Electrochemical evaluation of molybdenum disulfide as a catalyst for hydrogen evolution in microbial electrolysis cells. Int. J. Hydrog. Energy 2011, 36, (16), 9439-9445. 56. Munoz, L. D.; Bergel, A.; Feron, D.; Basseguy, R., Hydrogen production by electrolysis of a phosphate solution on a stainless steel cathode. Int. J. Hydrog. Energy 2010, 35, (16), 8561-8568. 57. Munoz, L. D.; Erable, B.; Etcheverry, L.; Riess, J.; Basseguy, R.; Bergel, A., Combining phosphate species and stainless steel cathode to enhance hydrogen evolution in microbial electrolysis cell (MEC). Electrochemistry Communications 2010, 12, (2), 183-186. 58. Elzinga, E. J.; Sparks, D. L., Phosphate adsorption onto hematite: An in situ ATR-FTIR investigation of the effects of pH and loading level on the mode of phosphate surface complexation. Journal of Colloid and Interface 20

ACS Paragon Plus Environment

Page 21 of 22

613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656

Environmental Science & Technology

Science 2007, 308, (1), 53-70. 59. Madsen, R. F.; Thomassen, J. R.; Vial, D.; Binot, R. A., THE PROPOSAL FOR THE COMPLETELY CLOSED SYSTEM IN THE COLUMBUS SPACE STATION. Desalination 1991, 83, (1-3), 123-136. 60. Larsen, T. A.; Gujer, W., Separate management of anthropogenic nutrient solutions (human urine). Water Science and Technology 1996, 34, (3-4), 87-94. 61. Hanaeus, J.; Hellstrom, D.; Johansson, E., A study of a urine separation in an ecological village in northern Sweden. Water Science and Technology 1997, 35, (9), 153-160. 62. Wilsenach, J., Van Loosdrecht, M. Separate urine collection and treatment: Options for sustainable wastewater systems and mineral recovery; 2001-39 STOWA: 2002. 63. Udert, K. M.; Larsen, T. A.; Biebow, M.; Gujer, W., Urea hydrolysis and precipitation dynamics in a urinecollecting system. Water Research 2003, 37, (11), 2571-2582. 64. Wu, M. S.; Jao, C. Y.; Chuang, F. Y.; Chen, F. Y., Carbon-encapsulated nickel-iron nanoparticles supported on nickel foam as a catalyst electrode for urea electrolysis. Electrochimica Acta 2017, 227, 210-216. 65. Wenzler-Roettele, S.; Dettenkofer, M.; Schmidt-Eisenlohr, E.; Gregersen, A.; Schulte-Moenting, J.; Tvede, M., Comparison in a laboratory model between the performance of a urinary closed system bag with double nonreturn valve and that of a single valve system. Infection 2006, 34, (4), 214-218. 66. Zou, X.; Zhang, Y., Noble metal-free hydrogen evolution catalysts for water splitting. Chemical Society Reviews 2015, 44, (15), 5148-5180. 67. Grosvenor, A. P.; Biesinger, M. C.; Smart, R. S.; McIntyre, N. S., New interpretations of XPS spectra of nickel metal and oxides. Surface Science 2006, 600, (9), 1771-1779. 68. Lenglet, M.; Dhuysser, A.; Bonelle, J. P.; Durr, J.; Jorgensen, C. K., ANALYSIS OF X-RAY NI K-BETA EMISSION, XANES, XPS, NI2P, AND OPTICAL-SPECTRA OF NICKEL(II) SPINELS AND STRUCTURE INFERENCE. Chemical Physics Letters 1987, 136, (5), 478-482. 69. Peck, M. A.; Langell, M. A., Comparison of Nanoscaled and Bulk NiO Structural and Environmental Characteristics by XRD, XAFS, and XPS. Chemistry of Materials 2012, 24, (23), 4483-4490. 70. Gautier, J. L.; Rios, E.; Gracia, M.; Marco, J. F.; Gancedo, J. R., Characterisation by X-ray photoelectron spectroscopy of thin MnxCo3-xO4 (1 >= x >= 0) spinel films prepared by low-temperature spray pyrolysis. Thin Solid Films 1997, 311, (1-2), 51-57. 71. Marco, J. F.; Gancedo, J. R.; Gracia, M.; Gautier, J. L.; Rios, E.; Berry, F. J., Characterization of the nickel cobaltite, NiCo2O4 prepared by several methods: An XRD, XANES, EXAFS, and XPS study. Journal of Solid State Chemistry 2000, 153, (1), 74-81. 72. Sickafus, K. E.; Wills, J. M.; Grimes, N. W., Structure of spinel. Journal of the American Ceramic Society 1999, 82, (12), 3279-3292. 73. Vedharathinam, V.; Botte, G. G., Understanding the electro-catalytic oxidation mechanism of urea on nickel electrodes in alkaline medium. Electrochimica Acta 2012, 81, 292-300. 74. Wu, Z. B.; Zhu, Y. R.; Ji, X. B., NiCo2O4-based materials for electrochemical supercapacitors. Journal of Materials Chemistry A 2014, 2, (36), 14759-14772. 75. Yuan, C. Z.; Li, J. Y.; Hou, L. R.; Zhang, X. G.; Shen, L. F.; Lou, X. W., Ultrathin Mesoporous NiCo2O4 Nanosheets Supported on Ni Foam as Advanced Electrodes for Supercapacitors. Advanced Functional Materials 2012, 22, (21), 4592-4597. 76. Benck, J. D.; Pinaud, B. A.; Gorlin, Y.; Jaramillo, T. F., Substrate Selection for Fundamental Studies of Electrocatalysts and Photoelectrodes: Inert Potential Windows in Acidic, Neutral, and Basic Electrolyte. Plos One 2014, 9, (10), 13. 77. Diaz-Morales, O.; Ferrus-Suspedra, D.; Koper, M. T. M., The importance of nickel oxyhydroxide 21

ACS Paragon Plus Environment

Environmental Science & Technology

657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693

Page 22 of 22

deprotonation on its activity towards electrochemical water oxidation. Chemical Science 2016, 7, (4), 2639-2645. 78. Bard, A.; Faulkner, L., Electrochemical Methods, Fundamentals and Applications. 2nd ed.; Wiley Global Education: 2000. 79. Zhong, C.; Hu, W. B.; Cheng, Y. F., Recent advances in electrocatalysts for electro-oxidation of ammonia. Journal of Materials Chemistry A 2013, 1, (10), 3216-3238. 80. Keuleers, A.; Desseyn, H. O.; Rousseau, B.; Van Alsenoy, C., Vibrational analysis of urea. Journal of Physical Chemistry A 1999, 103, (24), 4621-4630. 81. Grdadolnik, J.; Marechal, Y., Urea and urea-water solutions-an infrared study. Journal of Molecular Structure 2002, 615, (1-3), 177-189. 82. Antelo, J.; Fiol, S.; Perez, C.; Marino, S.; Arce, F.; Gondar, D.; Lopez, R., Analysis of phosphate adsorption onto ferrihydrite using the CD-MUSIC model. Journal of Colloid and Interface Science 2010, 347, (1), 112-119. 83. Weng, L. P.; Vega, F. A.; Van Riemsdijk, W. H., Competitive and Synergistic Effects in pH Dependent Phosphate Adsorption in Soils: LCD Modeling. Environmental Science & Technology 2011, 45, (19), 8420-8428. 84. Yavuz, E.; Tokalioglu, S.; Sahan, H.; Berberoglu, A.; Patat, S., Vortexing/shaking-free solid phase extraction of lead(II) by using an urchin-like NiCo2O4 hollow microsphere adsorbent. Microchimica Acta 2017, 184, (4), 1191-1198. 85. Connor, P. A.; McQuillan, A. J., Phosphate adsorption onto TiO2 from aqueous solutions: An in situ internal reflection infrared spectroscopic study. Langmuir 1999, 15, (8), 2916-2921. 86. Rajbanshi, A.; Wan, S.; Custelcean, R., Dihydrogen Phosphate Clusters: Trapping H2PO4- Tetramers and Hexamers in Urea-Functionalized Molecular Crystals. Crystal Growth & Design 2013, 13, (5), 2233-2237. 87. Buhlmann, P.; Nishizawa, S.; Xiao, K. P.; Umezawa, Y., Strong hydrogen bond-mediated complexation of H2PO4- by neutral bis-thiourea hosts. Tetrahedron 1997, 53, (5), 1647-1654. 88. Heise, H. M.; Voigt, G.; Lampen, P.; Kupper, L.; Rudloff, S.; Werner, G., Multivariate calibration for the determination of analytes in urine using mid-infrared attenuated total reflection spectroscopy. Applied Spectroscopy 2001, 55, (4), 434-443. 89. Oliver, K. V.; Marechal, A.; Rich, P. R., Effects of the Hydration State on the Mid-Infrared Spectra of Urea and Creatinine in Relation to Urine Analyses. Applied Spectroscopy 2016, 70, (6), 983-994. 90. Gao, J.; Hu, Y. J.; Li, S. X.; Zhang, Y. J.; Chen, X., Tautomeric equilibrium of creatinine and creatininium cation in aqueous solutions explored by Raman spectroscopy and density functional theory calculations. Chemical Physics 2013, 410, 81-89. 91. Butler, A. R.; Glidewell, C., Creatinine - An Examination Of Its Structure And Some Of Its Reactions By Synergistic Use Of Mndo Calculations And Nuclear Magnetic-Resonance Spectroscopy. Journal of the Chemical Society-Perkin Transactions 2 1985, (9), 1465-1467. 92. Ye, K.; Zhang, H. Y.; Zhao, L. T.; Huang, X. M.; Cheng, K.; Wang, G. L.; Cao, D. X., Facile preparation of three-dimensional Ni(OH)(2)/Ni foam anode with low cost and its application in a direct urea fuel cell. New Journal of Chemistry 2016, 40, (10), 8673-8680.

22

ACS Paragon Plus Environment