Mechanistic Study of the Validity of Using Hydroxyl ... - ACS Publications

Jan 10, 2017 - Department of Chemical Engineering, University of Illinois at Chicago, 810 South Clinton Street, Chicago, Illinois 60607, United. State...
0 downloads 0 Views 3MB Size
Subscriber access provided by the Henry Madden Library | California State University, Fresno

Article

A Mechanistic Study of the Validity of Using Hydroxyl Radical Probes to Characterize Electrochemical Advanced Oxidation Processes Yin Jing, and Brian P. Chaplin Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b05513 • Publication Date (Web): 10 Jan 2017 Downloaded from http://pubs.acs.org on January 12, 2017

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 free 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 accessible to all readers and 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.

Environmental Science & Technology 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 41

Environmental Science & Technology

2 3 4 5

A Mechanistic Study of the Validity of Using Hydroxyl Radical

6

Probes to Characterize Electrochemical Advanced Oxidation

7

Processes

8 9

Yin Jing, Brian P. Chaplin*

10

Department of Chemical Engineering, University of Illinois at Chicago, 810 South Clinton Street,

11

Chicago, Illinois 60607, United States

12



ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 41

13

ABSTRACT

14

The detection of hydroxyl radicals (OH•) is typically accomplished by using reactive probe

15

molecules, but prior studies have not thoroughly investigated the suitability of these probes for

16

use in electrochemical advanced oxidation processes (EAOPs), due to the neglect of alternative

17

reaction mechanisms. In this study, we investigated the suitability of four OH• probes (coumarin,

18

p-chlorobenzoic acid, terephthalic acid, and p-benzoquinone) for use in EAOPs. Experimental

19

results indicated that both coumarin and p-chlorobenzoic acid are oxidized via direct electron

20

transfer reactions, while p-benzoquinone and terephthalic acid are not. Coumarin oxidation to

21

form the OH• adduct product 7-hydroxycoumarin was found at anodic potentials lower than that

22

necessary for OH• formation. Density functional theory (DFT) simulations found a

23

thermodynamically favorable and non-OH• mediated pathway for 7-hydroxycoumarin formation,

24

which is activationless at anodic potentials > 2.10 V/SHE. DFT simulations also provided

25

estimates of Eo values for a series of OH• probe compounds, which agreed with voltammetry

26

results. Results from this study indicate that terephthalic acid is the most appropriate OH• probe

27

compound for the characterization of electrochemical and catalytic systems.

28

KEYWORDS: hydroxyl radical, probe molecule, electrochemical systems, density functional

29

theory, Marcus theory, reaction mechanism



ACS Paragon Plus Environment

2

Page 3 of 41

30

Environmental Science & Technology

Graphical Abstract

31



ACS Paragon Plus Environment

3

Environmental Science & Technology

32

Page 4 of 41

1. Introduction

33

The hydroxyl radical (OH•) is a highly reactive oxidant species, which is produced via

34

autoxidation in biological systems, advanced oxidation processes (AOPs), and electrochemical

35

advanced oxidation processes (EAOPs).1–20 Its ability to react with organic compounds, often

36

with diffusion limited rate constants, and its rapid decay to an innocuous end product of water,

37

has led to an increased interest in using technologies that generate OH• for toxic pollutant

38

degradation.17,21–25 The formation of OH• at the solid-electrolyte interface is key to the operation

39

of photocatalytic and EAOP technologies, and it is an unwanted side product in several energy

40

technologies (e.g., polymer electrolyte membrane fuel cells; secondary batteries; redox flow

41

batteries).26–35 Therefore, quantifying OH• production and understanding formation mechanisms

42

are goals across numerous scientific fields. The short lifetime of OH• (~10-9 s in biological cells,22,24,25,36,37 and 10-6 to 10-3 s in water38–40)

43 44

and high reactivity complicate its analytical determination. Therefore, two indirect methods are

45

primarily employed for their detection. The first method is using electron spin resonance (ESR)

46

spectroscopy in the presence of a spin trap, which is a compound that selectively reacts with a

47

radical to form an adduct species that has a sufficient lifetime to allow ESR detection. The

48

second method uses probe compounds that selectively react with OH• and often form adduct

49

species that are detected analytically by conventional chromatographic techniques.

50

The most common spin traps for OH• determination are substituted nitrones, where 5,5-

51

Dimethyl-1-pyrroline N-oxide (DMPO) is commonly used due to its high selectivity for OH•.41–50

52

However, spin traps can result in the false positive detection of OH•, which is documented in the

53

biochemical and toxicology fields.51–56 False positive detection of OH• has been attributed to

54

inverted spin trapping52–54,57 or the Forrester-Hepburn mechanism.51,55,56 Inverted spin trapping

ACS Paragon Plus Environment

4

Page 5 of 41

Environmental Science & Technology

55

involves a direct electron transfer oxidation reaction of the spin trap (R) followed by nucleophilic

56

attack by water (Reaction 1). !!!

R•!

!!"!

57

R

58

The Forrester-Hepburn mechanism is the reverse of this process, which involves nucleophilic

59

R − OH

(1)

attack of the spin trap by water followed by direct electron transfer oxidation (Reaction 2). !!"!

R − OH !

!!!

60

R

61

Both methods form a radical adduct (R-OH) which is identical to that formed by OH• attack

62

on the original spin trap. Cyclic voltammetry experiments have shown that spin trap compounds

63

used for OH• detection have peak potentials (Ep) of 1.47 to 2.17 V/SHE on a Pt electrode in

64

nonaqueous solvents (e.g., DMPO, Ep = 1.87 V/SHE),15,58 and thus they are not suitable for OH•

65

detection during photocatalytic and EAOP studies. Studies have also shown that metal ions (e.g.,

66

Cu(II), Fe(III), Cr(V)) can catalyze the nucleophilic addition step of the Forrester-Hepburn

67

mechanism,46,59 which greatly decreases the redox potential necessary for the subsequent direct

68

electron transfer oxidation step.53 Eberson showed that the DMPO-OH- product of nucleophilic

69

addition was oxidized to the DMPO-OH adduct by compounds with low redox potentials (-0.26

70

to 0.74 V/SHE).53 While these mechanisms are well reported in the biochemistry literature,51–55

71

this knowledge has not been adequately transferred to catalytic and electrochemical fields, and

72

thus

73

photocatalytic,41,49,60–62

74

studies.46,50,61,64–68

spin

traps

R − OH

have

(2)

been

used

electrochemical

to

quantify

OH•

oxidation,6,11,42,44,63

production and

during

numerous

Fenton-based

oxidation

75

The alternative approach of using probe molecules to detect OH• production has also been

76

widely used in photocatalytic,8,24,37,62,69–76 electrochemical,7,11–16,20,63,77–83 Fenton,5,23,84–86 and

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 41

77

electrochemical Fenton studies.10,87,88 Several compounds have been used, including coumarin

78

(COU),13,14,16,24,37,69,72,79–81 salicylic acid,4,5,11,20 benzoic acid (BA),84,85,89,90 terephthalic acid

79

(TA),8,10,13,23,69–71,73–76,87,91 1,4-benzoquinone (p-BQ),12,14,15,92–94 p-chlorobenzoic acid (p-

80

CBA),7,95–103 p-nitrosodimethylaniline (RNO),15,63,82,83 luminol,73 3-carboxyproxyl,62 1,4-

81

dioxane,88 1-propanol,85 and methylene blue dye.86 A summary of commonly used OH• probes

82

used in various processes are included in the Supporting Information (SI Section S-1), and

83

oxidation potentials of select molecules are summarized in the SI (Table S-5). The most

84

commonly used OH• probes are COU, TA, BA, p-CBA, and p-BQ. The use of probe molecules

85

relies on either the detection of the adduct product, or the disappearance of the probe molecule.

86

Both direct electron transfer and Forrester-Hepburn mechanisms are also possible with probe

87

molecules, and thus could lead to the false positive detection of OH•. These mechanisms are

88

largely missing from the catalytic and electrochemical literature.

89

In principle an appropriate OH• probe should 1) react with OH• at diffusion-limited rates; 2)

90

be resistant to direct electron transfer oxidation reactions at solid/electrolyte interfaces and with

91

soluble redox species; 3) have sufficient solubility in aqueous solutions; and 4) yield a stable

92

product that is detectable by conventional analytical techniques. In this study, we experimentally

93

investigated the suitability of four common probes (i.e., COU, p-CBA, TA, and p-BQ) for OH•

94

detection at a boron-doped diamond (BDD) electrode using electrochemical methods. Density

95

functional theory simulations were conducted to interpret the experimental data and determine

96

mechanisms for probe compound oxidation. The primary objectives of this study were to

97

determine a stable, selective, and sensitive probe for the detection of OH• produced during

98

EAOPs, and determine the oxidation mechanisms for their removal. Results are discussed in a

99

broader context related to catalytic and photocatalytic systems.



ACS Paragon Plus Environment

6

Page 7 of 41

100

Environmental Science & Technology

2. Materials and Method

101

2.1 Reagents. Chemicals were reagent-grade and used as received (Sigma-Aldrich). Solution

102

pH was adjusted with H3PO4 or NaOH. All solutions were made with deionized water obtained

103

from a NANOPure water purification system (Barnstead) with resistivity greater than 18.2

104

MΩ·cm (25o C).

105

2.2 Electrochemical Characterization. Linear sweep voltammetry (LSV) and cyclic

106

voltammetry (CV) experiments were performed using a Gamry Reference 600 potentiostat

107

(Warminster, PA). A three-electrode setup was used, with BDD as the working electrode (0.35

108

cm2), a 10 cm long, 0.3 mm diameter Pt wire as counter, and a single-junction saturated Ag/AgCl

109

electrode as reference (Pine Research Instruments, Grove City, Pennsylvania). Experiments used

110

a 50 mL, 1.0 M KH2PO4 background electrolyte (pH = 5.8) at 25 °C. Potentials were reported

111

versus the standard hydrogen electrode (SHE) by subtracting 0.197 V from the potential

112

measured versus the reference. Before each experiment, the BDD electrode was preconditioned

113

in a blank electrolyte solution (100 mM KH2PO4) at an anodic current density of 20 mA cm-2 for

114

10 min, followed by a cathodic current density of 20 mA cm-2 for another 10 min. The

115

preconditioning step was used to remove adsorbed organic contaminants from the electrode

116

surface. LSV simulations were performed in Matlab 2015b, and details can be found in the SI

117

Section S-3. 2.3 Bulk Oxidation Experiments. Experiments were performed using a rotating disk

118 119

electrode (RDE) setup at 5000 rpm using a Pine Research Instruments (Grove City, Pennsylvania)

120

rotator assembly (Model: AFMSRCE). Solution temperatures (15, 25, 35 and 45° C) were

121

maintained in a divided and jacketed glass reactor (H-cell) using a re-circulating water bath

122

(Thermo Scientific Neslab RTE 7). A Nafion N115 membrane separator (Ion Power Inc., New

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 41

123

Castle, Delaware) was used in the H-cell to isolate anodic/cathodic reactions. A schematic of the

124

reactor setup is shown in the SI Figure S-1.

125

2.4 Mass Transfer Rates Constants. The mass transfer rate constant (km) was determined

126

by methods described by Donaghue and Chaplin.104 Briefly, the limiting current technique was

127

used to estimate the limiting current density and km to the electrode surface at a given rotation

128

speed. The km values at 25 °C are listed in the SI (Table S-6). The diffusion coefficient was

129

calculated by the Stokes-Einstein equation, and the temperature dependent viscosity was

130

calculated using the equation proposed by Al-Shemmeri.105

131

2.5 Reaction Rate Constants and Curve Fitting. The reaction rate constants were

132

determined by by simultaneous regression of duplicate experimental data sets, and the fitting

133

results are reported with 95% confidence intervals. The specific, surface area normalized

134

reaction rate constant (ks (m s-1)) was obtained by dividing the first order rate constant by the

135

specific surface area (0.7 m-1).

136

2.6 Analytical Methods. Liquid chromatography (LC) (Shimadzu, LC-20AD) used a SPD-

137

M30A variable wavelength UV-Visible detector set at 254 nm and a reverse-phase C18 column

138

(Phenomenex, 4.6 × 250 mm, 5 μm). Methanol:water (40:60, v/v for COU and 50:50 for p-BQ)

139

and methanol:water with 0.1% (v/v) formic acid (60:40 v/v for TA) were used as mobile phases

140

at a total flow rate of 1.0 mL·min-1. The injection volume was programmed at 5 µL (COU or p-

141

BQ) or 10 µL (TA). LC with a fluorescent detector (TF-20A, Shimadzu) was used for 7-

142

hydroxycoumarin (7-OH COU) (λex=332 nm and λem = 471 nm) and 2-hydroxyterephthalic acid

143

(2-OH TA) (λex = 315 nm and λem = 435 nm) quantification. Analysis was performed in duplicate

144

and evaluated using LabSolutions Lite Version 5.63 software.



ACS Paragon Plus Environment

8

Page 9 of 41

Environmental Science & Technology

145

2.7 Quantum Mechanical Simulations. Density functional theory (DFT) simulations were

146

performed using Gaussian 09 software.106 Unrestricted spin, all-electron calculations were

147

performed using the 6-31G+(d) basis set for geometry optimization and frequency calculations

148

and the 6-311G+(3df, 2p) basis set for energy calculations. Scale factors of 0.9806 and 0.989

149

were used to correct for known systematic errors in frequency and thermal energies,

150

respectively.107 Frequency calculations allowed for adjustment of the Gibbs free energies of the

151

reactants and products to standard state conditions (∆! ! ). The gradient corrected Becke, three-

152

parameter, Lee−Yang−Parr (B3LYP) functional was used for exchange and correlation. Implicit

153

water solvation was incorporated using the SMD model.108 The standard reduction potential of the direct electron transfer reaction (! ! ) was calculated

154 155

according to Equation (3). !! = −

156

∆! ! ! !"

! − !!"# (SHE)

(3)

157

where ∆! ! ! is the free energy of the reduction reaction, n is the number of electrons transferred,

158

! F is the Faraday constant, and !!!" (SHE) is a reference value for the absolute standard

159

! reduction potential of the SHE (!!"# SHE = 4.28 eV).109–111 Values for ∆! ! ! were calculated

160

from the ∆! ! values determined from the geometrically optimized reactant and product

161

structures. Gibbs free energy of activation (Ea) versus electrode potential profiles for direct electron

162 163

transfer oxidation reactions were determined using Marcus theory according to Equation (4).112 !! =

164

!! !

1−

!".!(!!! ! )

!

(4)

!!

165

where E is the electrode potential (V/SHE) and !! is the total reorganization energy for the

166

forward reaction (oxidation reaction). The reorganization energy represents the energy needed to



ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 41

167

transform the geometric structure of the reactant and solvent to those of the product. Values for

168

!! were calculated by subtracting ∆! ! of the reactant from that of a compound with an identical

169

geometry of the product, but with the same charge as the reactant. Ionic effects of the supporting

170

electrolyte on !! were not calculated, as previous work has shown the effect is minimal in polar

171

solvents with high dielectric constants.113 Transition state structures were determined using the

172

synchronous transit-guided quasi-newton (STQN) method114,115 using the QST3 scheme in the

173

Gaussian 09 software. When this method did not converge, the pseudo reaction coordinate

174

method was used. Transition state structures were verified by confirming that frequency

175

calculations yielded a single negative frequency corresponding to the normal mode along the

176

reaction coordinate. Previous work indicates that errors in the calculation of the reaction energies

177

by DFT methods may be up to ~ 16 kJ mol-1.116

178

3. Results and Discussion

179

3.1 Voltammetry Results. In order to determine if COU, TA, p-CBA, and p-BQ underwent

180

direct electron transfer reactions at the BDD surface, LSV experiments were conducted with

181

various concentrations of these compounds. Results from LSV experiments indicated that only

182

COU (Figure 1a) and p-CBA (Figure 1c) exhibited an increased current with respect to the

183

background electrolyte at approximately 1.7 and 2.0 V/SHE, respectively, while TA and p-BQ

184

did not (SI Figure S-2). The absence of a current response for TA and p-BQ in this potential

185

range suggested that these compounds did not undergo direct electron transfer reactions. By

186

contrast, linear relationships between the peak currents and concentrations were observed for

187

both COU and p-CBA (Figures 1b and 1d), suggesting they were oxidized by direct electron

188

transfer reactions at the BDD surface. Additionally, LSV study of COU, p-CBA, and TA on a Pt

189

electrode showed that oxidation peaks associated with direct electron transfer were found for

ACS Paragon Plus Environment

10

Page 11 of 41

Environmental Science & Technology

190

both COU and p-CBA at 1.3 and 1.4 V/SHE, respectively (SI Figure S-3). Evidence for a direct

191

electron transfer reaction for TA was not found (SI Figure S-3). Therefore, it is concluded that

192

the findings of this study are not exclusive to BDD.

193

Figures 2a and 2b show the background subtracted LSV scans of 5 mM COU at different

194

scan rates and solution pH values. The reversal scans are not shown, since corresponding

195

cathodic peaks were not observed in the CV scans (SI Figure S-4), indicating either a fast

196

chemical reaction removed the direct electron transfer oxidation product or the product itself was

197

not electrochemically active. The anodic peaks for COU oxidation shifted to less positive

198

potentials with increasing solution pH, indicating OH- participated in the reaction.117 For

199

example, at a scan rate of 300 mV s-1 the peak potentials for COU oxidation at pH 2.16, 5.80 and

200

11.95 were 2.45, 2.38 and 2.12 V/SHE, respectively. The slope between the anodic peak

201

potential and pH at a scan rate of 20 mV s-1 was -22.5 ± 3.40 mV per pH unit. Figure 2c shows

202

that the peak current plotted against the square-root of scan rate follows a linear relationship,

203

indicating a diffusion controlled process and therefore interaction with the electrode surface was

204

not significantly affecting the electron transfer reaction.112

205

The kinetic parameters governing the electrochemical reaction kinetics were determined by

206

fitting a mathematical model to the experimental LSV data. The LSV model fits for COU (root-

207

mean-square error (rms) = 0.072) and p-CBA (rms = 0.058) are shown in Figures 2b and 1c. The

208

simulation results provided E0 values for COU and p-CBA oxidation of 1.95 and 2.34 V/SHE,

209

respectively. The oxidation kinetics of p-CBA were more facile than COU, as the first order

210

heterogeneous kinetic rate constant ko for p-CBA oxidation (1.46 × 10-3 m s-1) was

211

approximately 3 orders of magnitude higher than COU (6.00 × 10-6 m s-1). The values for



ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 41

212

transfer coefficient α were determined as 0.7 and 0.6 for COU and p-CBA oxidation,

213

respectively, which agree with reported values.112

214 215 216 217 218



Figure 1 LSV of (a) COU and (c) p-CBA (background subtracted) with different concentrations, and relationship between peak currents and concentrations with linear regression fit of the data for (b) COU and (d) p-CBA. In panel (c), hollow symbols and solid lines represent experimental and simulation data, respectively. Temperature: 25°C and electrolyte: 1.0 M KH2PO4 (pH 5.8).



ACS Paragon Plus Environment

12

Page 13 of 41

Environmental Science & Technology

219



220 221 222 223 224

Figure 2 LSV scans of 5 mM COU (background subtracted) under different scan rates and solution pH values, (a) pH 2.16, (b) pH 5.80 and (c) pH 11.95 (d) Relationship between peak current and square-root of scan rate in different pH solutions. In panel (b), hollow symbols and solid lines represent experimental and simulation data, respectively. Temperature: 25°C and electrolyte: 1.0 M KH2PO4 (see each panel for pH).

225

3.2 Bulk Oxidation Experiments. Results from LSV scans indicated that COU reacted via a

226

direct electron transfer reaction at potentials > 1.7 V/SHE (Figure 1a), and therefore bulk

227

oxidation of 1 mM COU was performed at anodic potentials of 1.7 and 2.0 V/SHE to determine

228

the Ea for COU oxidation. The solution was also monitored for the presence of the 7-OH COU

229

product, which is generally considered to be an adduct compound that is formed by OH• addition

230

(Figure 3). Pseudo first-order reaction rate constants were fit to the concentration versus time

231

profiles measured during COU oxidation, and results are summarized in the SI (Table S-7). The



ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 41

232

calculated km for COU was nearly one order of magnitude higher than ks values determined for

233

COU oxidation at both potentials, and therefore the oxidation of COU was assumed to be

234

kinetically limited. The Arrhenius plots yielded Ea values of 45.2 ± 2.75 (SI Figure S-5) and 16.8

235

± 3.63 kJ mol-1 (Figure 3d) at 1.7 and 2.0 V/SHE, respectively. Significant concentrations of the

236

fluorescent 7-OH COU product were detected during the bulk oxidation of COU (Figure 3). At

237

the conclusion of the oxidation experiments 2.12 × 10-5 and 4.36 × 10-4 mM of 7-OH COU was

238

observed at 1.7 and 2.0 V/SHE, respectively, corresponding to average percent yields of 0.03%

239

and 0.22%, respectively. However, OH• production should be negligible at the anodic potentials

240

of the experiments (1.7 and 2.0 V/SHE), due to the high redox potential for OH• formation (Eo =

241

2.80 V/SHE118). Negligible OH• production is also supported by the low background current (i.e.,

242

0.29 - 0.31 mA) observed in LSV scans of the blank electrolyte (Figure 1a). Therefore, it was

243

assumed that the production of 7-OH COU at anodic potentials of both 1.7 and 2.0 V/SHE was

244

by an alternative reaction mechanism.

245

Sweep voltammetry results indicate that TA and p-BQ were resistant to direct electron

246

transfer oxidation, and thus the oxidation of both compounds at a potential greater than that

247

necessary for OH• formation was attributed solely to OH• oxidation. An anodic potential of 3.0

248

V/SHE, which was sufficient to produce OH• at the BDD electrode,119 was chosen for the

249

oxidation of both compounds. Results from the bulk oxidation of TA and p-BQ at various

250

temperatures are shown in the SI (Figure S-6 and S-7). The values of ks obtained for the

251

concentration profiles of both compounds are summarized in the SI (Table S-7). A significant

252

concentration (4.63 × 10-4 mM and average percent yield of 0.11%) of the fluorescent 2-OH TA

253

product from the bulk oxidation of TA was detected at the conclusion of the experiment (SI;

254

Figure S-6). Meanwhile, there were not any products detected by either the UV-vis or fluorescent



ACS Paragon Plus Environment

14

Page 15 of 41

Environmental Science & Technology

255

detectors from p-BQ oxidation, indicating that ring cleavage of p-BQ occurred, which is

256

consistent with previous studies.93,120,121 The ks values for both TA and p-BQ oxidation were

257

comparable to the km values (SI Table S-7), indicating the bulk oxidation of both TA and p-BQ at

258

3.0 V/SHE were mass transfer controlled. Mass transfer controlled oxidation of p-BQ and TA

259

was attributed to the fast reaction with OH• (kOH•, p-BQ = 1.2×109 M-1s-1 21 and kOH•, TA = 4.4×109 M-1s-1

260

91

261

therefore the disappearance of p-BQ during anodic oxidation experiments was used to assess

262

OH• formation. However, p-BQ is easily reduced to hydroquinone (E0 = 0.28 V/SHE122), and

263

therefore p-BQ is only a useful OH• probe in a divided cell, where the anode and cathode are

264

separated by a selective membrane (e.g., Nafion membrane). Therefore, p-BQ is also not a useful

265

probe in photocatalytic systems where the catalyst possesses both oxidative and reductive active

266

sites.



). There is not a signature product from p-BQ oxidation that is indicative of OH formation, and

267 268 269



Figure 3. Bulk oxidation of 1 mM COU at 2.0 V/SHE and a temperature of (a) 15 ºC, (b) 25 ºC, and (c) 35 ºC, respectively. (d) Arrhenius plot for the oxidation of COU at 2.0 V/SHE. and

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 41

270 271 272

represent reactant (COU) and product (7-OH COU), respectively. Experiments were performed in duplicate, and the 1st order reaction rate model was fitted to both data sets simultaneously (dashed line). Electrolyte: 100 mM KH2PO4 (pH 4.5).

273

3.3 DFT Modeling. DFT modeling was used to determine Eo and λf values for the direct

274

electron transfer reaction (reaction 1) of each of the OH• probe compounds investigated

275

experimentally (COU, p-BQ, TA, p-CBA). These values were used in Equation (4) to determine

276

Ea profiles as a function of electrode potential. Some additional common OH• probe compounds

277

were also investigated theoretically, including BA and RNO. Results are shown in Figure 4, and

278

Eo and λf values are reported in the SI (Table S-8). The Eo values of 0.84, 2.04, 2.48, 2.55, 2.76,

279

and 3.21 V/SHE were determined for RNO, COU, p-CBA, BA, TA, and p-BQ, respectively.

280

These results were consistent with values determined by model fits of the LSV scans for COU

281

and p-CBA, which yielded Eo values of 2.34 V/SHE (Figure 1c) and 1.95 V/SHE (Figure 2b),

282

respectively. A comparison between experimental and DFT determined Ea values for COU also

283

showed close agreement (Figure 4). Experimentally measured Ea values for COU were 45.2 ±

284

2.75 kJ mol-1 at 1.7 V/SHE and 16.8 ± 3.63 kJ mol-1 at 2.0 V/SHE, which were within -1.2 kJ

285

mol-1 and 12.1 kJ mol-1 of DFT determined values, respectively (Figure 4). The close agreement

286

between experimental and theoretical Ea values at 1.7 V/SHE shows the robustness of the

287

theoretical method. The difference between experimental and theoretical Ea values at 2.0 V/SHE

288

is largely a result of unactivated processes influencing the experimental Ea measurement. For

289

example, temperature effects on molecular diffusion, adsorption, or the structure of the double

290

layer can cause errors in accurately measuring Ea values in this range.123

291

Figure 4 also contains the Ea versus potential profile for OH• formation at an oxidized BDD

292

surface, taken from Chaplin et al.124 Results indicated that the direct electron transfer reaction for

293

RNO and COU proceeded at potentials much less than that for OH• formation, and thus neither



ACS Paragon Plus Environment

16

Page 17 of 41

Environmental Science & Technology

294

RNO nor COU are useful OH• probes. The Ea versus potential profiles for p-CBA and BA in

295

Figure 4 are intersected by the profile for OH•, indicating direct electron transfer reactions were

296

likely dominant at potentials < 2.3 V/SHE, and that both direct electron transfer and OH•

297

oxidation likely occurred at potentials > 2.3 V/SHE for these compounds. The results from LSV

298

experiments showed the direct electron transfer oxidation of p-CBA occurred at potentials > 2.0

299

V/SHE (Figure 1c). By contrast, direct electron transfer reactions for TA and p-BQ were

300

predicted to occur after OH• formation. At anodic potentials > 2.3 V/SHE the water oxidation

301

reaction dominated the measured current and therefore LSV scans were not able to provide

302

evidence for direct electron transfer reactions involving p-BQ and TA. In addition, these

303

compounds have fast reaction rate constants with OH• (i.e., kOH•, BQ = 1.2×109 M-1 s-1; kOH•, COU =

304

2.0×109 M-1 s-1),21,91 and at high anodic potentials reactions became mass transfer limited (as

305

noted in SI Table S-7). Past studies have concluded that under mass transfer limited conditions

306

the indirect oxidation with OH• was the dominant oxidation mechanism, because the substrates

307

were depleted via reaction with OH• in the diffusion boundary layer and did not reach the

308

electrode surface.104 The results above suggest that p-BQ and TA are resistant to direct electron

309

transfer reactions and therefore should be appropriate OH• probes, while other probe compounds

310

investigated are not, and their use in electrochemical and heterogeneous catalytic systems should

311

be revisited.

312

An additional requirement for the OH• probes is that they are nonreactive by the Forrester-

313

Hepburn mechanism. Reactivity by this mechanism was assessed by determining Ea and ∆! ! !

314

values for the nucleophilic addition of OH- to each of the compounds, followed by calculation of

315

Eo and λf values for the subsequent direct electron transfer reaction (Table 1). Transition state

316

structures for the nucleophilic reactions are provided in the SI (Figures S-8). Simulations were



ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 41

317

conducted for OH- attack on the C7 and C2 atoms for COU and TA, respectively, as these were

318

the fluorescent adduct probes used to assess OH• production in this and other

319

studies.8,10,13,14,16,23,24,37,69–76,79,81,87,91,125 Simulations determined that nucleophilic addition of OH-

320

to COU yielded Ea = 130 kJ mol-1 and ∆! ! ! = 118 kJ mol-1, and addition of OH- to TA yielded

321

Ea = 123 kJ mol-1 and ∆! ! ! = 111 kJ mol-1 (Table 1). The relatively high Ea values and positive

322

∆! ! ! values suggested that these reactions were unlikely to be significant at room temperature.

323

Experimental results supported these theoretical calculations, as removal of COU and TA were

324

not observed under alkaline conditions (pH = 12.4, T = 25 oC). However, catalytic nucleophilic

325

addition has been reported to occur by metal ions in solution (e.g., Fe, Cu, Cr).46,59 Therefore,

326

this reaction may be catalyzed by Fenton type reactions and therefore warrants further study. The

327

Eo values for the corresponding direct electron transfer reactions are -0.23 and -0.05 V/SHE for

328

COU and TA (Table 1), respectively; indicating that these reactions would proceed at anodic

329

potentials provided the nucleophilic addition reactions occurred.

330



ACS Paragon Plus Environment

18

Page 19 of 41

Environmental Science & Technology

331 332 333 334

Figure 4. DFT determined Ea versus electrode potential profiles. Profile for OH• taken from Chaplin et al., 2013.124 Squares represent experimental values determined for COU oxidation in RDE experiments.

335 336

Table 1. Theoretical parameters determined for the reaction of probe molecules by the ForresterHepburn mechanism. Reaction COU + OH- à COU-OHCOU-OH- à COU-OH + eTA + OH- à TA-OH-

-

TA-OH à TA-OH + e

p-BQ + OH- à p-BQ-OHp-BQ-OH- à p-BQ-OH + ep-CBA + OH- à p-HBA + Cl-

∆! !!

Ea

Eo

λf

(kJ mol-1) 118

(kJ mol-1) 130

(V/SHE) --

(kJ mol-1) --

--

--

-0.23

13.4

111

123

--

--

--

--

-0.05

11.9

14.9

42.1

--

--

--

--

1.22

14.9

-146

125

--

--

337 338

The nucleophilic addition of OH- to p-BQ was investigated at both the C1 and C2 positions.

339

Due to the symmetry of the molecule, these were the only two unique positions for OH- attack.

340

The ∆! ! ! for nucleophilic addition of OH- at the C1 and C2 positions of p-BQ were 14.9 and 289

341

kJ mol-1, respectively (Table 1). These results indicated that attack at the C1 position was

342

thermodynamically favorable and the calculated Ea of 42.1 kJ mol-1 indicated that the reaction

343

was feasible at room temperature (Table 1).126 The corresponding direct electron transfer

344

reaction has a calculated Eo = 1.22 V/SHE, and thus this reaction would readily occur at anodic

345

potentials used in EAOPs and during photocatalysis.

346

The nucleophilic addition of OH- to p-CBA was studied at the C4 position, and resulted in a

347

substitution reaction that formed p-hydroxybenzoic acid (p-HBA), with Cl- as the leaving group

348

(Table 1). The calculated Ea and ∆! ! ! values for the nucleophilic substitution reaction were 125



ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 41

349

and -146 kJ mol-1, respectively (Table 1), indicating the reaction was thermodynamically

350

favorable but likely not significant at room temperature. Since p-HBA is a stable product and p-

351

CBA is usually used as a OH• probe based on its disappearance, the direct electron transfer

352

reaction was not studied. These results indicated that p-BQ has the potential to react via the

353

Forrester-Hepburn mechanism, and therefore reactions should not be conducted in alkaline

354

conditions. Results for other compounds are consistent with studies of nucleophilic attack on

355

aromatic derivatives, which indicated that high temperatures are needed for significant reaction

356

rates to proceed.127–129

357

Experimental results showed that the anodic oxidation of COU produced the 7-OH COU

358

product at anodic potentials lower than that necessary to produce OH• (i.e., 1.7 and 2.0 V/SHE),

359

which suggested that this product was not exclusively formed by reaction with OH•. Therefore,

360

DFT simulations were conducted to determine a thermodynamically favorable mechanism for its

361

formation. The overall anodic reaction of COU to form 7-OH COU is given by Reaction (5). 362

+ OH- + H2O



+ H3O+ + 2e-

363

(5)

364

This reaction involves 2e- oxidation, which may occur by either direct electron transfer or

365

indirect oxidation with OH•. Results in Figure 4 show that the direct electron transfer oxidation

366

was activationless at potentials > 2.1 V/SHE, which was likely the initial step responsible for

367

conversion to 7-OH COU via a non-OH• radical mechanism. A possible reaction mechanism that

368

does not involve OH• oxidation is given in Figure 5, where ∆! ! ! values were calculated at an

369

electrode potential of E = 2.044 V/SHE (Ea = 2.44 kJ mol-1 (Figure 4)). The mechanism involves

370

three steps, one single electron transfer electrochemical (E) step, followed by one chemical (C)

371

step and another single electron transfer (E) step, and the overall ∆! ! ! = -390 kJ mol-1 at E =

ACS Paragon Plus Environment

20

Page 21 of 41

Environmental Science & Technology

372

2.044 V/SHE. The mechanism was defined as an ECE mechanism. At an electrode potential of

373

2.044 V/SHE, the first electrochemical step (E1) had ∆! ! ! = 0 kJ mol-1, and values for Ea versus

374

electrode potential are given in Figure 4. All subsequent steps were exothermic. The first

375

chemical step (C1) involved nucleophilic attack by OH-, which was activationless and ∆! ! ! = -

376

152 kJ mol-1. The second electrochemical step (E2) involved a coupled H+ and e- transfer to form

377

7-OH COU. This reaction had a calculated Eo = -0.43 V/SHE, and thus was activationless at an

378

electrode potential of 2.044 V/SHE (∆! ! ! = -239 kJ mol-1). The participation of OH- in the

379

reaction mechanism was consistent with LSV experimental results that showed OH- dependence

380

on the measured peak potential for COU (Figure 2). The stepwise e- and H+ transfer reactions

381

were also investigated, which yielded Eo = 0.81 V/SHE for the e- transfer step and the H+ transfer

382

to H2O was activationless.

383

The results above provided an activationless and exothermic pathway for the formation of 7-

384

OH COU from COU oxidation at an applied anodic potential of E > 2.1 V/SHE, indicating that

385

COU was not a useful OH• probe. Since the initial direct electron transfer reaction was rate

386

limiting, the Ea versus E profile given in Figure 4 should accurately describe the energy barrier

387

for 7-OH COU formation as a function of potential for the non-OH• pathway.



ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 41

+

150 100 50 ΔG (kJ mol-1)

0 -50 -100

+ OH- + H2O

+ OH- + H2O + e-

ΔGE1 = 0 (E = 2.04 V/SHE) ΔGC1 = -102

+ H2O + e- ΔGr = -341

-150 -200

ΔGE2 = -239 (E = 2.04 V/SHE)

-250 -300 -350 -400

E1: COU-H à COU-H+ + e- (Eo = 2.04 V/SHE)

+ H3O+ + 2e-

C1: COU-H+ + OH- à COU-H-OH E2: COU-H-OH + H2O à COU-OH + H3O+ + e- (Eo = -0.43 V/SHE)

388 389

Figure 5. Free energy diagram of COU oxidation by ECE mechanism.

390

3.4 Broader Significance and Environmental Impacts. Although this study was focused

391

on selecting appropriate OH• probes for EAOPs, our findings have broader impacts that extend to

392

other processes involving electron transfer, such as photocatalysis and energy storage/conversion

393

technologies. In photocatalysis, the formation of OH• occurs through a direct electron transfer

394

reaction from H2O to the UV light excited holes in the valence band of the photocatalyst. This

395

same type of direct electron transfer reaction mechanism can occur for OH• probe compounds,

396

and thus may be falsely interpreted as an indication of OH• formation. Often the direct electron

397

transfer reaction mechanism was assessed in photocatalysis by adding a OH• scavenging

398

compound and assessing its effect on a given substrate.8,24,37,69,70,72–76,98,99 This method does not

399

provide definitive proof for or against a direct electron transfer mechanism, since the OH•

400

scavenger can block direct electron transfer reaction sites and thus slow the reaction of the



ACS Paragon Plus Environment

22

Page 23 of 41

Environmental Science & Technology

401

substrate. In energy storage/conversion technologies, OH• can be formed from H2O2, which is a

402

product of the oxygen reduction reaction at the cathode surface. The formation of OH• is also

403

possible during the charging process in aqueous electrolytes. Therefore, the use of appropriate

404

probe compounds to characterize OH• formation during the charging/discharging process at the

405

electrodes or during fuel cell operation is necessary to prevent false positive detection of OH•.

406

A detailed literature review of OH• probes used in various processes is provided in the SI

407

Section S-1 (Table S-1 to S-4), and electrochemical oxidation potentials of select molecules are

408

summarized in the SI (Table S-5). Table S-5 indicates that OH• probes, such as salicylic acid,

409

luminol, 1-propanol and methylene blue, are susceptible to direct electron transfer reaction, since

410

their reported oxidation potentials are lower than 2.3 V/SHE (Figure 4), however, 1,4-dioxane,

411

with the oxidation potential at 2.4 V, may be an appropriate OH• probe in electrochemical and

412

catalytic systems, but requires further investigation.

413

Many prior studies have used EAOPs and photocatalysis for the oxidation of contaminants in

414

water, and therefore the correct identification and quantification of OH• is necessary in order to

415

understand and optimize these processes for environmental remediation. Careful attention should

416

be given when selecting the appropriate probes to detect the formation of OH• in these processes.

417

Spin trap compounds, RNO, COU, and p-CBA can react by direct electron transfer and

418

Forrester-Hepburn mechanisms, and therefore are not appropriate probes. The results of this

419

study indicated that TA is the most selective OH• probe investigated, since it has a high redox

420

potential (Eo = 2.76 V/SHE), is resistant to the Forrester-Hepburn mechanism, and is not easily

421

reducible.

422

ASSOCIATED CONTENT



ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 41

423

Supporting Information

424

Literature reviews of OH• probes used in various systems, Electrochemical oxidation potentials

425

of selective compounds, Details of LSV simulation, Schematics of System Setup, LSV of COU,

426

p-CBA and TA on Pt Electrode, Diffusion coefficients and mass transfer rates for COU, TA and

427

p-BQ, LSV of TA and p-BQ, CV of COU, Reaction rate constants for bulk oxidation of COU,

428

TA, and p-BQ, COU oxidation under various temperatures at 1.7 V vs SHE, TA and p-BQ

429

oxidation under various temperatures at 3.0 V vs SHE, Theoretical E0 and λf values determined

430

by DFT methods and used to determine Ea versus potential profiles, and Transition state

431

structure.

432

AUTHOR INFORMATION

433

Corresponding Author

434

*Phone: 312-996-0288; fax: 312-996-0808; e-mail: [email protected].

435

Acknowledgements

436

Funding for this work was provided by the National Science Foundation (CBET-1159764 and

437

CBET-1453081 (CAREER)).

438

References

439

(1)

nutrients. FASEB J. 1987, 1 (6), 441–445.

440 441

Machlin, L.; Bendich, A. Free radical tissue damage: protective role of antioxidant

(2)

Kesselman, J. M.; Weres, O.; Lewis, N. S.; Hoffmann, M. R. Electrochemical Production of Hydroxyl Radical at Polycrystalline Nb-Doped TiO2 Electrodes and Estimation of the

442



ACS Paragon Plus Environment

24

Page 25 of 41

Environmental Science & Technology

443

Partitioning between Hydroxyl Radical and Direct Hole Oxidation Pathways. J. Phys.

444

Chem. B 1997, 101 (14), 2637–2643.

445

(3)

Jurva, U.; Wikström, H. V; Bruins, A. P. Electrochemically assisted Fenton reaction:

446

reaction of hydroxyl radicals with xenobiotics followed by on-line analysis with high-

447

performance liquid chromatography/tandem mass spectrometry. Rapid Commun. Mass

448

Spectrom. 2002, 16 (20), 1934–1940.

449

(4)

Ai, S.; Wang, Q.; Li, H.; Jin, L. Study on production of free hydroxyl radical and its

450

reaction with salicylic acid at lead dioxide electrode. J. Electroanal. Chem. 2005, 578 (2),

451

223–229.

452

(5)

Jen, J.-F.; Leu, M.-F.; Yang, T. C. Determination of hydroxyl radicals in an advanced

453

oxidation process with salicylic acid trapping and liquid chromatography. J. Chromatogr.

454

A 1998, 796 (2), 283–288.

455

(6)

Cong, Y.; Wu, Z.; Li, Y. Hydroxyl radical electrochemically generated with water as the

456

complete atom source and its environmental application. Chinese Sci. Bull. 2007, 52 (10),

457

1432–1435.

458

(7)

Electrochemical Reactor. Ozone Sci. Eng. 2008, 30 (September 2014), 113–119.

459 460

Kim, J.; Korshin, G. V. Generation of Hydroxyl Radicals and Ozone in a Flow-through

(8)

Xiao, Q.; Ouyang, L. Photocatalytic activity and hydroxyl radical formation of carbon-

461

doped TiO2 nanocrystalline: Effect of calcination temperature. Chem. Eng. J. 2009, 148

462

(2–3), 248–253.

463

(9)



Mishin, V. M.; Thomas, P. E. Characterization of hydroxyl radical formation by

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 41

464

microsomal enzymes using a water-soluble trap, terephthalate. Biochem. Pharmacol. 2004,

465

68 (4), 747–752.

466

(10)

Yue, Q.; Zhang, K.; Chen, X.; Wang, L.; Zhao, J.; Liu, J.; Jia, J. Generation of OH

467

radicals in oxygen reduction reaction at Pt-Co nanoparticles supported on graphene in

468

alkaline solutions. Chem. Commun. (Camb). 2010, 46 (19), 3369–3371.

469

(11)

Marselli, B.; Garcia-Gomez, J.; Michaud, P. -a.; Rodrigo, M. a.; Comninellis, C.

470

Electrogeneration of Hydroxyl Radicals on Boron-Doped Diamond Electrodes. J.

471

Electrochem. Soc. 2003, 150 (3), D79.

472

(12)

Zaky, A. M.; Chaplin, B. P. Porous substoichiometric TiO2 anodes as reactive

473

electrochemical membranes for water treatment. Environ. Sci. Technol. 2013, 47 (12),

474

6554–6563.

475

(13)

Guo, L.; Jing, Y.; Chaplin, B. P. Development and Characterization of Ultrafiltration TiO2

476

Magnéli Phase Reactive Electrochemical Membranes. Environ. Sci. Technol. 2016, 50 (3),

477

1428–1436.

478

(14)

Electrochim. Acta 2011, 56 (5), 2246–2253.

479 480

Rueffer, M.; Bejan, D.; Bunce, N. J. Graphite: An active or an inactive anode?

(15)

Bejan, D.; Malcolm, J. D.; Morrison, L.; Bunce, N. J. Mechanistic investigation of the

481

conductive ceramic Ebonex® as an anode material. Electrochim. Acta 2009, 54 (23),

482

5548–5556.

483

(16)

Bejan, D.; Guinea, E.; Bunce, N. J. On the nature of the hydroxyl radicals produced at boron-doped diamond and Ebonex® anodes. Electrochim. Acta 2012, 69, 275–281.

484



ACS Paragon Plus Environment

26

Page 27 of 41

485

Environmental Science & Technology

(17)

electrochemical wastewater treatment system. J. Hazard. Mater. 2003, 103 (1–2), 65–78.

486 487

(18)

Chaplin, B. P. Critical review of electrochemical advanced oxidation processes for water treatment applications. Environ. Sci. Process. Impacts 2014, 1 (312), 1182–1203.

488 489

Feng, C.; Sugiura, N.; Shimada, S.; Maekawa, T. Development of a high performance

(19)

Tröster, I.; Fryda, M.; Herrmann, D.; Schäfer, L.; Hänni, W.; Perret, A.; Blaschke, M.;

490

Kraft, A.; Stadelmann, M. Electrochemical advanced oxidation process for water

491

treatment using DiaChem® electrodes. Diam. Relat. Mater. 2002, 11 (3–6), 640–645.

492

(20)

Guinea, E.; Arias, C.; Cabot, P. L.; Garrido, J. A.; Rodríguez, R. M.; Centellas, F.; Brillas,

493

E. Mineralization of salicylic acid in acidic aqueous medium by electrochemical advanced

494

oxidation processes using platinum and boron-doped diamond as anode and cathodically

495

generated hydrogen peroxide. Water Res. 2008, 42 (1–2), 499–511.

496

(21)

Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B.; Tsang, W. Critical Review

497

of rate constants for reactions of hydrated electrons Chemical Kinetic Data Base for

498

Combustion Chemistry. Part 3: Propane. J. Phys. Chem. Ref. Data 1988, 17 (2), 513.

499

(22) Pryor, W. Oxy-radicals and related species: their formation, lifetimes, and reactions. Annu. Rev. Physiol. 1986, 48 (1), 657–667.

500 501

(23)

Linxiang, L.; Abe, Y.; Nagasawa, Y.; Kudo, R.; Usui, N.; Imai, K.; Mashino, T.;

502

Mochizuki, M.; Miyata, N. An HPLC assay of hydroxyl radicals by the hydroxylation

503

reaction of terephthalic acid. Biomed. Chromatogr. 2004, 18 (7), 470–474.

504

(24)

Czili, H.; Horváth, A. Applicability of coumarin for detecting and measuring hydroxyl radicals generated by photoexcitation of TiO2 nanoparticles. Appl. Catal. B Environ. 2008,

505



ACS Paragon Plus Environment

27

Environmental Science & Technology

81 (3–4), 295–302.

506 507

(25)

Louit, G.; Foley, S.; Cabillic, J.; Coffigny, H.; Taran, F.; Valleix, A.; Renault, J. P.; Pin, S.

508

The reaction of coumarin with the OH radical revisited: hydroxylation product analysis

509

determined by fluorescence and chromatography. Radiat. Phys. Chem. 2005, 72 (2–3),

510

119–124.

511

Page 28 of 41

(26)

Trogadas, P.; Parrondo, J.; Ramani, V. Degradation Mitigation in Polymer Electrolyte

512

Membranes Using Cerium Oxide as a Regenerative Free-Radical Scavenger. Electrochem.

513

Solid-State Lett. 2008, 11 (7), B113.

514

(27)

Perrot, C.; Gonon, L.; Bardet, M.; Marestin, C.; Pierre-Bayle, A.; Gebel, G. Degradation

515

of a sulfonated aryl ether ketone model compound in oxidative media (sPAEK). Polymer

516

(Guildf). 2009, 50 (7), 1671–1681.

517

(28)

Gummalla, M.; Atrazhev, V. V.; Condit, D.; Cipollini, N.; Madden, T.; Kuzminyh, N. Y.;

518

Weiss, D.; Burlatsky, S. F. Degradation of Polymer-Electrolyte Membranes in Fuel Cells.

519

J. Electrochem. Soc. 2010, 157 (11), B1542.

520

(29)

Li, H.; Tsay, K.; Wang, H.; Shen, J.; Wu, S.; Zhang, J.; Jia, N.; Wessel, S.; Abouatallah,

521

R.; Joos, N.; et al. Durability of PEM fuel cell cathode in the presence of Fe3+ and Al3+. J.

522

Power Sources 2010, 195 (24), 8089–8093.

523

(30)

Atrazhev, V.; Timokhina, E.; Burlatsky, S. F.; Sultanov, V.; Madden, T.; Gummalla, M.

524

Direct Mechanism of OH Radicals Formation in PEM Fuel Cells. ECS Trans. 2008, 6, 69–

525

74.

526

(31)



Borup, R.; Meyers, J.; Pivovar, B.; Kim, Y. S.; Mukundan, R.; Garland, N.; Myers, D.;

ACS Paragon Plus Environment

28

Page 29 of 41

Environmental Science & Technology

527

Wilson, M.; Garzon, F.; Wood, D.; et al. Scientific aspects of polymer electrolyte fuel cell

528

durability and degradation. Chemical Reviews. American Chemical Society 2007, pp

529

3904–3951.

530

(32)

electrolyte membranes. Int. J. Hydrogen Energy 2006, 31 (13), 1838–1854.

531 532

Collier, A.; Wang, H.; Zi Yuan, X.; Zhang, J.; Wilkinson, D. P. Degradation of polymer

(33)

Hommura, S.; Kawahara, K.; Shimohira, T.; Teraoka, Y. Development of a method for

533

clarifying the perfluorosulfonated membrane degradation mechanism in a fuel cell

534

environment. J. Electrochem. Soc. 2007, 155 (1), A29–A33.

535

(34)

Maurya, S.; Shin, S.; Kim, M.; Yun, S.; Moon, S. Stability of composite anion exchange

536

membranes with various functional groups and their performance for energy conversion. J.

537

Memb. Sci. 2013, 443, 28–35.

538

(35)

Chen, D.; Wang, S.; Xiao, M.; Han, D.; Meng, Y. Sulfonated poly (fluorenyl ether ketone)

539

membrane with embedded silica rich layer and enhanced proton selectivity for vanadium

540

redox flow battery. J. Power Sources 2010, 195 (22), 7701–7708.

541

(36)

Chem. 2006, 75 (4), 473–478.

542 543

(37)

Xiang, Q.; Yu, J.; Wong, P. K. Quantitative characterization of hydroxyl radicals produced by various photocatalysts. J. Colloid Interface Sci. 2011, 357 (1), 163–167.

544 545

Newton, G. L.; Milligan, J. R. Fluorescence detection of hydroxyl radicals. Radiat. Phys.

(38)

Méndez-Díaz, J.; Sánchez-Polo, M.; Rivera-Utrilla, J.; Canonica, S.; Von Gunten, U.

546

Advanced oxidation of the surfactant SDBS by means of hydroxyl and sulphate radicals.

547

Chem. Eng. J. 2010, 163 (3), 300–306.



ACS Paragon Plus Environment

29

Environmental Science & Technology

548

(39)

(40)

Hoffmann, M. R.; Martin, S. T.; Choi, W. Y.; Bahnemann, D. W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95 (1), 69–96.

551 552

Arslan-Alaton, I. A review of the effects of dye-assisting chemicals on advanced oxidation of reactive dyes in wastewater. Color. Technol. 2003, 119 (6), 345–353.

549 550

(41)

Kim, K. W.; Lee, E. H.; Kim, Y. J.; Lee, M. H.; Kim, K. H.; Shin, D. W. A relation

553

between the non-stoichiometry and hydroxyl radical generated at photocatalytic TiO2 on

554

4CP decomposition. J. Photochem. Photobiol. A Chem. 2003, 159 (3), 301–310.

555

(42)

Cao, J.; Zhao, H.; Cao, F.; Zhang, J.; Cao, C. Electrocatalytic degradation of 4chlorophenol on F-doped PbO2 anodes. Electrochim. Acta 2009, 54 (9), 2595–2602.

556 557

Page 30 of 41

(43)

Aguilera-Venegas, B.; Speisky, H. Identification of the transition state for fast reactions:

558

The trapping of hydroxyl and methyl radicals by DMPO - A DFT approach. J. Mol. Graph.

559

Model. 2014, 52, 57–70.

560

(44)

Zhao, X.; Qu, J.; Liu, H.; Hu, C. Photoelectrocatalytic degradation of triazine-containing

561

azo dyes at gamma-Bi2MoO6 film electrode under visible light irradiation (lamda > 420

562

nm). Environ. Sci. Technol. 2007, 41 (19), 6802–6807.

563

(45)

Bhattacharjee, S.; Khan, M. N.; Chandra, H.; Symons, M. C. R. Radical cations from

564

nitrone spin-traps: reaction with water to give OH adducts. J. Chem. Soc. Perkin Trans. 2

565

1996, No. 12, 2631.

566

(46)

Sugden, K. D.; Wetterhahn, K. E. Reaction of Chromium (V) with the EPR Spin Traps 5 ,

567

5-Dimethylpyrroline N-Oxide and Phenyl-N-tert-butylnitrone Resulting in Direct

568

Oxidation. Inorg. Chem. 1996, No. V, 651–657.



ACS Paragon Plus Environment

30

Page 31 of 41

569

Environmental Science & Technology

(47)

Roberts, J. G.; Voinov, M. A.; Schmidt, A. C.; Smirnova, T. I.; Sombers, L. A. The

570

Hydroxyl Radical is a Critical Intermediate in the Voltammetric Detection of Hydrogen

571

Peroxide. J. Am. Chem. Soc. 2016, 138 (8), 2516–2519.

572

(48)

Villamena, F. a; Hadad, C. M.; Zweier, J. L. Theoretical study of the spin trapping of

573

hydroxyl radical by cyclic nitrones: a density functional theory approach. J. Am. Chem.

574

Soc. 2004, 126 (6), 1816–1829.

575

(49)

Nosaka, Y.; Komori, S.; Yawata, K.; Hirakawa, T.; Nosaka, A. Y. Photocatalytic ˙OH

576

radical formation in TiO2 aqueous suspension studied by several detection methods. Phys.

577

Chem. Chem. Phys. 2003, 5 (20), 4731–4735.

578

(50) Rafat Husain, S.; Cillard, J.; Cillard, P. Hydroxyl radical scavenging activity of flavonoids. Phytochemistry 1987, 26 (9), 2489–2491.

579 580

(51)

703.

581 582

(52)

(53)

Eberson, L. Formation of Hydroxyl Spin Adducts via Nucleophilic Addition—Oxidation to 5, 5-DimethyI-1-pyrroline N-Oxide. Acta Chem. Scand. 1999, 53, 584–593.

585 586

Chandra, H.; Symons, M. C. R. Hydration of spin-trap cations as a source of hydroxyl adducts. J. Chem. Soc. Chem. Commun. 1986, 16, 1301–1302.

583 584

Forrester, A.; Hepburn, S. Spin traps. A cautionary note. J. Chem. Soc. C Org. 1971, 701–

(54)

Cerri, V.; Frejaville, C.; Vila, F.; Allouche, A. Synthesis, redox behavior and spin-trap

587

properties of 2, 6-di-tert-butylnitrosobenzene (DTBN). J. Org. Chem. 1989, 54 (6), 1447–

588

1450.

589

(55)



Leinisch, F.; Ranguelova, K.; Derose, E. F.; Jiang, J.; Mason, R. P. Evaluation of the

ACS Paragon Plus Environment

31

Environmental Science & Technology

590

Forrester-Hepburn mechanism as an artifact source in ESR spin-trapping. Chem. Res.

591

Toxicol. 2011, 24 (12), 2217–2226.

592

(56)

Leinisch, F.; Jiang, J.; Derose, E. F.; Khramtsov, V. V.; Mason, R. P. Investigation of

593

spin-trapping artifacts formed by the Forrester-Hepburn mechanism. Free Radic. Biol.

594

Med. 2013, 65, 1497–1505.

595

(57)

Ranguelova, K.; Mason, R. P. The fidelity of spin trapping with DMPO in biological systems. Magn. Reson. Chem. 2011, 49 (4), 152–158.

596 597

(58)

McIntire, G. L.; Blount, H. N.; Stronks, H. J.; Shetty, R. V.; Janzen, E. G. Spin trapping in

598

electrochemistry. 2. Aqueous and nonaqueous electrochemical characterizations of spin

599

traps. J. Phys. Chem. 1980, 84 (8), 916–921.

600

Page 32 of 41

(59)

Hanna, P. M.; Chamulitrat, W.; Mason, P.; Health, Z.; Box, P. O.; Hagi, A.; Nishi, M.;

601

Murakami, A. When are metal ion-dependent hydroxyl and alkoxyl radical adducts of 5,5-

602

dimethyl-1-pyrroline N-oxide artifacts? Arch. Biochem. Biophys. 1992, 296 (2), 640–644.

603

(60)

Jaeger, C. D.; Bard, A. J. Spin trapping and electron spin resonance detection of radical

604

intermediates in the photodecomposition of water at titanium dioxide particulate systems.

605

J. Phys. Chem. 1979, 83 (24), 3146–3152.

606

(61)

dimethyl sulfoxide. Arch. Biochem. Biophys. 1990, 278 (2), 478–481.

607 608

Steiner, M. G.; Babbs, C. F. Quantitation of the hydroxyl radical by reaction with

(62)

Schwarz, P. F.; Turro, N. J.; Bossmann, S. H.; Braun, A. M.; Wahab, A.-M. A. A.; Dürr,

609

H. A New Method To Determine the Generation of Hydroxyl Radicals in Illuminated

610

TiO2 Suspensions. J. Phys. Chem. B 1997, 101 (36), 7127–7134.



ACS Paragon Plus Environment

32

Page 33 of 41

611

Environmental Science & Technology

(63)

Amadelli, R.; De Battisti, A.; Girenko, D. V.; Kovalyov, S. V.; Velichenko, A. B.

612

Electrochemical oxidation of trans-3,4-dihydroxycinnamic acid at PbO2 electrodes: Direct

613

electrolysis and ozone mediated reactions compared. Electrochim. Acta 2000, 46 (2–3),

614

341–347.

615

(64)

Jurva, U.; Wikström, H. V.; Bruins, A. P. Electrochemically assisted Fenton reaction:

616

Reaction of hydroxyl radicals with xenobiotics followed by on-line analysis with high-

617

performance liquid chromatography/tandem mass spectrometry. Rapid Commun. Mass

618

Spectrom. 2002, 16 (20), 1934–1940.

619

(65)

capacity (HOSC) estimation. J. Agric. Food Chem. 2006, 54 (3), 617–626.

620 621

Moore, J.; Yin, J.-J.; Yu, L. L. Novel fluorometric assay for hydroxyl radical scavenging

(66)

Finkelstein, E.; Rosen, G. M.; Rauckman, E. J. Spin trapping. Kinetics of the reaction of

622

superoxide and hydroxyl radicals with nitrones. J. Am. Chem. Soc. 1980, 102 (15), 4994–

623

4999.

624

(67)

Fukui, S.; Hanasaki, Y.; Ogawa, S. High-performance liquid chromatographic

625

determination of methanesulphinic acid as a method for the determination of hydroxyl

626

radicals. J. Chromatogr. A 1993, 630 (1–2), 187–193.

627

(68)

hydroxyl radicals. Anal. Chim. Acta 2001, 434 (2), 169–177.

628 629

Yang, X.-F.; Guo, X.-Q. Study of nitroxide-linked naphthalene as a fluorescence probe for

(69)

Ishibashi, K.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Detection of active oxidative

630

species in TiO2 photocatalysis using the fluorescence technique. Electrochem. commun.

631

2000, 2 (3), 207–210.



ACS Paragon Plus Environment

33

Environmental Science & Technology

632

(70)

Page 34 of 41

SAV Eremia, D Chevalier-Lucia, GL Radu, J. M. Optimization of hydroxyl radical

633

formation using TiO2 as photocatalyst by response surface methodology. Talanta 2008, 77

634

(2), 858–862.

635

(71)

Nakabayashi, Y.; Nosaka, Y. The pH dependence of OH radical formation in photo-

636

electrochemical water oxidation with rutile TiO2 single crystals. Phys. Chem. Chem. Phys.

637

2015, 17 (45), 30570–30576.

638

(72)

Nakabayashi, Y.; Nosaka, Y. OH Radical Formation at Distinct Faces of Rutile TiO2

639

Crystal in the Procedure of Photoelectrochemical Water Oxidation. J. Phys. Chem. C 2013,

640

117 (45), 23832–23839.

641

(73)

Hirakawa, T.; Nosaka, Y. Properties of O2• and OH• Formed in TiO2 Aqueous

642

Suspensions by Photocatalytic Reaction and the Influence of H2O2 and Some Ions.

643

Langmuir 2002, 18 (8), 3247–3254.

644

(74)

Xiao, Q.; Si, Z.; Zhang, J.; Xiao, C.; Tan, X. Photoinduced hydroxyl radical and

645

photocatalytic activity of samarium-doped TiO2 nanocrystalline. J. Hazard. Mater. 2008,

646

150 (1), 62–67.

647

(75)

Thiruvenkatachari, R.; Kwon, T. O.; Jun, J. C.; Balaji, S.; Matheswaran, M.; Moon, I. S.

648

Application of several advanced oxidation processes for the destruction of terephthalic

649

acid (TPA). J. Hazard. Mater. 2007, 142 (1–2), 308–314.

650

(76)

Shafaei, A.; Nikazar, M.; Arami, M. Photocatalytic degradation of terephthalic acid using

651

titania and zinc oxide photocatalysts: Comparative study. Desalination 2010, 252 (1–3),

652

8–16.



ACS Paragon Plus Environment

34

Page 35 of 41

653

Environmental Science & Technology

(77)

Ai, S.; Wang, Q.; Li, H.; Jin, L. Study on production of free hydroxyl radical and its

654

reaction with salicylic acid at lead dioxide electrode. J. Electroanal. Chem. 2005, 578 (2),

655

223–229.

656

(78)

Montilla, F.; Michaud, P. .; Morallón, E.; Vázquez, J. .; Comninellis, C. Electrochemical

657

oxidation of benzoic acid at boron-doped diamond electrodes. Electrochim. Acta 2002, 47

658

(21), 3509–3513.

659

(79)

Komatsu, M.; Rao, T.; Fujishima, A. Detection of Hydroxyl Radicals Formed on an

660

Anodically Polarized Diamond Electrode Surface in Aqueous Media. Chem. Lett. 2003,

661

32 (4), 396–397.

662

(80)

Ohguri, N.; Nosaka, A.; Nosaka, Y. Detection of OH radicals as the effect of Pt particles

663

in the membrane of polymer electrolyte fuel cells. J. Power Sources 2010, 195 (15),

664

4647–4652.

665

(81)

Salazar-Gastélum, M.; Lin, S. Electrochemical and Spectrometric Studies for the

666

Determination of the Mechanism of Oxygen Evolution Reaction. J. Electrochem. Soc.

667

2016, 163 (5), G37–G43.

668

(82)

Muff, J.; Bennedsen, L. R.; Søgaard, E. G. Study of electrochemical bleaching of p-

669

nitrosodimethylaniline and its role as hydroxyl radical probe compound. J. Appl.

670

Electrochem. 2011, 41 (5), 599–607.

671

(83)

Inactivation of Microorganisms. Environ. Sci. Technol. 2006, 40 (19), 6117–6122.

672 673

Jeong, J.; Kim, J. E. E. Y. The Role of Reactive Oxygen Species in the Electrochemical

(84)



Ou, B.; Hampsch-Woodill, M.; Flanagan, J.; Deemer, E. K.; Prior, R. L.; Huang, D. Novel

ACS Paragon Plus Environment

35

Environmental Science & Technology

674

Fluorometric Assay for Hydroxyl Radical Prevention Capacity Using Fluorescein as the

675

Probe. J. Agric. Food Chem. 2002, 50 (10), 2772–2777.

676

(85)

Lindsey, M. E.; Tarr, M. A. Quantitation of hydroxyl radical during Fenton oxidation following a single addition of iron and peroxide. Chemosphere 2000, 41 (3), 409–417.

677 678

(86)

Satoh, A. Y.; Trosko, J. E. Methylene Blue Dye Test for Rapid Qualitative Detection of

679

Hydroxyl Radicals Formed in a Fenton’s Reaction Aqueous Solution. Environ. Sci.

680

Technol. 2007, 41 (8), 2881–2887.

681

(87)

Liu, J.; Lagger, G.; Tacchini, P.; Girault, H. H. Generation of OH radicals at palladium

682

oxide nanoparticle modified electrodes, and scavenging by fluorescent probes and

683

antioxidants. J. Electroanal. Chem. 2008, 619–620, 131–136.

684

Page 36 of 41

(88)

Kishimoto, N.; Sugimura, E. Feasibility of an electrochemically assisted Fenton method

685

using Fe2+/HOCl system as an advanced oxidation process. Water Sci. Technol. 2010, 62

686

(10), 2321–2329.

687

(89)

Montilla, F.; Michaud, P. .; Morallón, E.; Vázquez, J. .; Comninellis, C. Electrochemical

688

oxidation of benzoic acid at boron-doped diamond electrodes. Electrochim. Acta 2002, 47

689

(21), 3509–3513.

690

(90)

Regoli, F.; Winston, G. W.; Radicals, P.; Radi-, H.; Phar-, G. W. T. A. Quantification of

691

Total Oxidant Scavenging Capacity of Antioxidants for Peroxynitrite , Peroxyl Radicals ,

692

and Hydroxyl Radicals. Reactions 1999, 105, 96–105.

693

(91)

Page, S. E.; Arnold, W. A.; McNeill, K. Terephthalate as a probe for photochemically generated hydroxyl radical. J. Environ. Monit. 2010, 12 (9), 1658.

694



ACS Paragon Plus Environment

36

Page 37 of 41

695

Environmental Science & Technology

(92)

Houk, L. L.; Johnson, S. K.; Feng, J.; Houk, R. S.; Johnson, D. C. Electrochemical

696

incineration of benzoquinone in aqueous media using a quaternary metal oxide electrode

697

in the absence of a soluble supporting electrolyte. J. Appl. Electrochem. 1998, 28 (11),

698

1167–1177.

699

(93)

Incineration of Benzoquinone. J. Electrochem. Soc. 1995, 142 (11), 3626.

700 701

Feng, J. Electrocatalysis of Anodic Oxygen-Transfer Reactions: The Electrochemical

(94)

Nien Shuchmann, M.; Bothe, E.; Justus, von S.; von Sonntag, C. Reaction of OH radicals

702

with benzoquinone in aqueous solutions . A pulse radiolysis study. Spectrum 1998, 0 (4),

703

791–796.

704

(95)

radicals. J. Phys. Chem. 1966, 70 (6), 2660–2662.

705 706

Anbar, M.; Meyerstein, D.; Neta, P. Reactivity of aromatic compounds toward hydroxyl

(96)

Jeong, J.; Kim, C.; Yoon, J. The effect of electrode material on the generation of oxidants

707

and microbial inactivation in the electrochemical disinfection processes. Water Res. 2009,

708

43 (4), 895–901.

709

(97)

Han, W.; Zhang, P.; Zhu, W.; Yin, J.; Li, L. Photocatalysis of p-chlorobenzoic acid in

710

aqueous solution under irradiation of 254 nm and 185 nm UV light. Water Res. 2004, 38

711

(19), 4197–4203.

712

(98)

Cho, M.; Chung, H.; Choi, W.; Yoon, J. Linear correlation between inactivation of E. coli

713

and OH radical concentration in TiO2 photocatalytic disinfection. Water Res. 2004, 38 (4),

714

1069–1077.

715

(99)



Chen, C.-Y.; Jafvert, C. T. Photoreactivity of carboxylated single-walled carbon

ACS Paragon Plus Environment

37

Environmental Science & Technology

Page 38 of 41

716

nanotubes in sunlight: reactive oxygen species production in water. Environ. Sci. Technol.

717

2010, 44 (17), 6674–6679.

718

(100) Lee, M.; Yun, H. J.; Yu, S.; Yi, J. Enhancement in photocatalytic oxygen evolution via

719

water oxidation under visible light on nitrogen-doped TiO2 nanorods with dominant

720

reactive {102} facets. Catal. Commun. 2014, 43, 11–15.

721

(101) Yun, H. J.; Lee, D. M.; Yu, S.; Yoon, J.; Park, H. J.; Yi, J. Effect of valence band energy

722

on the photocatalytic performance of N-doped TiO2 for the production of O2 via the

723

oxidation of water by visible light. J. Mol. Catal. A Chem. 2013, 378, 221–226.

724

(102) Wu, D.; You, H.; Jin, D.; Li, X. Enhanced inactivation of Escherichia coli with Ag-coated

725

TiO2 thin film under UV-C irradiation. J. Photochem. Photobiol. A Chem. 2011, 217 (1),

726

177–183.

727

(103) Wu, D.; You, H.; Zhang, R.; Chen, C.; Lee, D. J. Inactivation of Amphidinium sp. in

728

ballast waters using UV/Ag-TiO2+O3 advanced oxidation treatment. Bioresour. Technol.

729

2011, 102 (21), 9838–9842.

730

(104) Donaghue, A.; Chaplin, B. P. Effect of select organic compounds on perchlorate

731

formation at boron-doped diamond film anodes. Environ. Sci. Technol. 2013, 47 (21),

732

12391–12399.

733

(105) Al-Shemmeri, T. Engineering Fluid Mechanics; Bookboon, 2012.

734

(106) Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.;

735

Barone, V.; Mennucci, B.; Petersson, G. Gaussian09. Gaussian. Inc.: Wallingford CT

736

2009.



ACS Paragon Plus Environment

38

Page 39 of 41

737

Environmental Science & Technology

(107) Bauschlicher, C. W.; Partridge, H. A modification of teh Gaussian-2 approach using density functional theory. J. Chem. Phys. 1995, 105 (5), 1788.

738 739

(108) Marenich, A. V; Cramer, C. J.; Truhlar, D. G. Universal solvation model based on solute

740

electron density and on a continuum model of the solvent defined by the bulk dielectric

741

constant and atomic surface tensions. J. Phys. Chem. B 2009, 113 (18), 6378–6396.

742

(109) Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Aqueous solvation free energies of ions and

743

ion-water clusters based on an accurate value for the absolute aqueous solvation free

744

energy of the proton. J. Phys. Chem. B 2006, 110 (32), 16066–16081.

745

(110) Tissandier, M. D.; Cowen, K. A.; Feng, W. Y.; Gundlach, E.; Cohen, M. H.; Earhart, A.

746

D.; Coe, J. V.; Tuttle, T. R. The proton’s absolute aqueous enthalpy and Gibbs free energy

747

of solvation from cluster-ion solvation data. J. Phys. Chem. A 1998, 102 (40), 7787–7794.

748

(111) Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Single-ion solvation free energies and the

749

normal hydrogen electrode potential in methanol, acetonitrile, and dimethyl sulfoxide. J.

750

Phys. Chem. B 2007, 111 (2), 408–422.

751

(112) Bard, A.; Faulkner, L. Electrochemical methods: fundamentals and applications, 2nd Editio.; Wiley, 2000.

752 753

(113) Vaissier, V.; Barnes, P.; Kirkpatrick, J.; Nelson, J. Influence of polar medium on the

754

reorganization energy of charge transfer between dyes in a dye sensitized film. Phys.

755

Chem. Chem. Phys. 2013, 15 (13), 4804–4814.

756

(114) Peng, C.; Bernhard Schlegel, H. Combining Synchronous Transit and Quasi-Newton Methods to Find Transition States. Isr. J. Chem. 1993, 33 (4), 449–454.

757



ACS Paragon Plus Environment

39

Environmental Science & Technology

758

(115) Peng, C. Y.; Ayala, P. Y.; Schlegel, H. B.; Frisch, M. J. Using redundant internal

759

coordinates to optimize equilibrium geometries and transition states. Journal of

760

Computational Chemistry. 1996, pp 49–56.

761

(116) Visser, S. de. The accuracy of density functional theory calculations in biocatalysis. J. Proteomics Enzymol. 2013.

762 763

Page 40 of 41

(117) Shah, A. H.; Zaid, W.; Shah, A.; Rana, U. A.; Hussain, H.; Ashiq, M. N.; Qureshi, R.;

764

Badshah, A.; Zia, M. A.; Kraatz, H.-B. pH Dependent Electrochemical Characterization,

765

Computational Studies and Evaluation of Thermodynamic, Kinetic and Analytical

766

Parameters of Two Phenazines. J. Electrochem. Soc. 2014, 162 (3), H115–H123.

767

(118) Chen, G. Electrochemical technologies in wastewater treatment. Sep. Purif. Technol. 2004, 38 (1), 11–41.

768 769

(119) Iniesta, J. Electrochemical oxidation of phenol at boron-doped diamond electrode. Electrochim. Acta 2001, 46 (23), 3573–3578.

770 771

(120) Houk, L. L.; Johnson, S. K.; Feng, J.; Houk, R. S.; Johnson, D. C. Electrochemical

772

incineration of benzoquinone in aqueous media using a quaternary metal oxide electrode

773

in the absence of a soluble supporting electrolyte. J. Appl. Electrochem. 1998, 28 (11),

774

1167–1177.

775

(121) Zaky, A. M.; Chaplin, B. P. Mechanism of p-Substituted Phenol Oxidation at a Ti4O7 Reactive Electrochemical Membrane. Environ. Sci. Technol. 2014, 48 (10), 5857–5867.

776 777

(122) Bard, A.; Mirkin, M. Scanning Electrochemical Microscopy; CRC Press, Inc., 2012.

778

(123) Smith, J. M. Chemical engineering kinetics; No. TP149 S58, 1981.



ACS Paragon Plus Environment

40

Page 41 of 41

779

Environmental Science & Technology

(124) Chaplin, B.; Hubler, D.; Farrell, J. Understanding anodic wear at boron doped diamond film electrodes. Electrochim. Acta 2013, 89, 122–131.

780 781

(125) Ohguri, N.; Nosaka, A.; Nosaka, Y. Detection of OH radicals formed at PEFC electrodes

782

by means of a fluorescence probe. Electrochem. Solid-State Lett. 2009, 12 (6), B94–B96.

783

(126) Brunmark, A. Formation of electronically excited states during the interaction of p-

784

benzoquinone with hydrogen peroxide. J. Biolumin. Chemilumin. 1989, 4 (1), 219–225.

785

(127) Bunnett, J. F.; Zahler, R. E. Aromatic Nucleophilic Substitution Reactions. Chem. Rev. 1951, 49 (2), 273–412.

786 787

(128) Oliveira, C. M.; Ferreira, A. C. S.; De Freitas, V.; Silva, A. M. S. Oxidation mechanisms occurring in wines. Food Res. Int. 2011, 44 (5), 1115–1126.

788 789

(129) Kutyrev, A. A. Nucleophilic reactions of quinones. Tetrahedron 1991, 47 (38), 8043– 8065.

790 791 792



ACS Paragon Plus Environment

41