Iron-Mediated Oxidation of Methoxyhydroquinone under Dark

Jan 20, 2016 - ... made available by participants in Crossref's Cited-by Linking service. ... Production of Hydrogen Peroxide in Groundwater at Rifle,...
1 downloads 0 Views 708KB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

Iron-mediated oxidation of methoxyhydroquinone under dark conditions: kinetic and mechanistic insights Xiu Yuan, James A. Davis, and Peter S. Nico Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b03939 • Publication Date (Web): 20 Jan 2016 Downloaded from http://pubs.acs.org on January 26, 2016

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 34

Environmental Science & Technology

1

Iron-mediated oxidation of methoxyhydroquinone under dark conditions: kinetic and

2

mechanistic insights

3

Xiu Yuan, James A. Davis, Peter S. Nico

4

Earth Sciences Division, Lawrence Berkeley National Laboratory,

5

1 Cyclotron Road, Berkeley, CA 94720, United States

6 7 8 9

Resubmitted

10

Environmental Science & Technology

11

January, 2016

12 13 14 15 16

* Corresponding author. Tel.: +1 (510) 486-7118, Fax: +1 (510) 486-5686

17

Email: [email protected]

18

1 ACS Paragon Plus Environment

Environmental Science & Technology

19

ABSTRACT

20

Despite the biogeochemical significance of the interactions between natural organic matter

21

(NOM) and iron species, considerable uncertainty still remains as to the exact processes

22

contributing to the rates and extents of complexation and redox reactions between these

23

important and complex environmental components. Investigations on the reactivity of low-

24

molecular-weight quinones, which are believed to be key redox active compounds within

25

NOM, towards iron species, could provide considerable insight into the kinetics and

26

mechanisms of reactions involving NOM and iron. In this study, the oxidation of 2-

27

methoxyhydroquinone (MH2Q) by ferric iron (Fe(III)) under dark conditions in the absence

28

and presence of oxygen was investigated within a pH range of 4 – 6. While Fe(III) was

29

capable of stoichiometrically oxidizing MH2Q under anaerobic conditions, catalytic oxidation

30

of MH2Q was observed in the presence of O2 due to further cycling between oxygen,

31

semiquinone radicals, and iron species. A detailed kinetic model was developed to describe

32

the predominant mechanisms, which indicated that both the undissociated and mono-

33

dissociated anions of MH2Q were kinetically active species towards Fe(III) reduction, with

34

the mono-dissociated anion being the key species accounting for the pH dependence of the

35

oxidation. The generated radical intermediates, namely semiquinone and superoxide, are of

36

great importance in reaction chain propagation. The kinetic model may provide critical

37

insight into the underlying mechanisms of the thermodynamic and kinetic characteristics of

38

metal-organic interactions and assist in understanding and predicting factors controlling iron

39

and organic matter transformation and bioavailability in aquatic systems.

2 ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34

40

Environmental Science & Technology

1. INTRODUCTION

41

Natural organic matter (NOM), a chemically diverse mixture of heterogeneous organic

42

compounds with various structural and functional properties, is ubiquitous in terrestrial and

43

aquatic environments. Considerable attention has been drawn to the redox properties of NOM

44

because of its great significance in a wide range of biogeochemical processes including iron

45

bioreduction and contaminant transformation. The electron accepting and donating capacities

46

of NOM have been attributed to the presence of redox-active constituents, such as quinone

47

and phenolic functional groups.1-4 Numerous studies also suggested that quinone moieties in

48

NOM not only contribute to the reduction of organic contaminants, such as alkyl halides and

49

nitroaromatics, but also stimulate the reduction of ferric oxides by iron reducing bacteria.3-9

50

The term quinone collectively refers to organic structures in three oxidation states with one

51

electron reduction of a quinone forming a semiquinone radical that can be further reduced to

52

the hydroquinone form.10 These species are common constituents of many biologically

53

relevant molecules and serve as a vital link in the movement of electrons through cells and

54

tissues.8, 11-13 For example, oxidations of dihydroxybenzenes to benzoquinones and

55

corresponding back reactions are often reversible under neutral and acidic conditions, earning

56

them a prominent place in organic electron transfer chemistry. Further, the frequently

57

observed positive correlation between free radical concentrations in NOM and Fe(III)

58

reduction using electron spin resonance (ESR) spectroscopy has been attributed to the

59

existence of semiquinone radicals, a radical formed either from one-electron oxidation of

60

hydroquinones, or from one-electron reduction of quinones, or from one-electron transfer

61

from hydroquinones to quinones (a process termed comproportionation).1, 3, 13-15 Though

62

commonly presented in very low concentrations due to its high reactivity, semiquinone 3 ACS Paragon Plus Environment

Environmental Science & Technology

63

radicals are kinetically important in the reversible redox chemistry of quinones.9 Previous

64

studies also indicated that quinone moieties are important electron-transfer groups in humic

65

substances (HS) for both biotic/abiotic reduction and oxidation of HS.1, 16 Besides, recent

66

studies by Garg, et al.17, 18 reported that the occurrence of Fe(III) reduction by Suwanee River

67

Fulvic Acid (SRFA, a proxy for terrestrial NOM), in non-irradiated acidic solutions was

68

mainly attributed to the reduced hydroquinone-like moieties intrinsically presented in SRFA.2,

69

4

70

important implications and may provide critical information regarding reaction mechanisms

71

and functional groups that participate in the mobilization/immobilization of a variety of

72

redox-sensitive metals in terrestrial and aquatic environments. For example, hydroquinone

73

mediated reduction of Fe(III) could increase the solubility and bioavailability of this critical

74

trace element,6 while quinone mediated oxidation of uranium (IV) could affect its mobility

75

and thus the remediation of contaminated sites.5

76

Iron, as a micronutrient in both aquatic and terrestrial ecosystems, is essential for virtually all

77

life forms and plays a critical role in many biogeochemical processes. In natural waters, iron

78

exists in two oxidation states, Fe(II) and Fe(III). Despite being more thermodynamically

79

stable, Fe(III) quickly hydrolyzes and precipitates at circumneutral pH and thus becomes

80

unavailable for direct biological uptake,19 while the reduced form, Fe(II), is much more

81

soluble and thus potentially more abundant in a bioavailable form.17 However, in oxygenated

82

waters, Fe(II) is rapidly oxidized to Fe(III) at circumneutral pH, which in turn limits its

83

bioavailability. Despite the biogeochemical functioning of the equilibration/cycling of

84

Fe(II)/Fe(III) couple, fundamental insights into the control and dynamics of interconversion

85

between Fe(II) and Fe(III) in natural waters are strictly limited and how these processes

86

restrict or enhance iron bioavailability are still largely unknown.

Therefore, studies involving the redox behaviour of quinones in metal reduction have

4 ACS Paragon Plus Environment

Page 4 of 34

Page 5 of 34

Environmental Science & Technology

87

Quinones and iron likely coexist in the majority of the natural ecosystem, ranging from the

88

surface water to groundwater, from soils to sediments. Since the highly complex nature of

89

NOM could potentially interfere with efforts to explore the reactivity of quinone moieties on

90

a molecular level, investigations of low-molecular-weight quinones could provide

91

considerable insight into the thermodynamics, kinetics, and mechanisms of reactions

92

involving quinone moieties.20 Though there is a range of different quinone moieties presented

93

in NOM, including benzoquinone-based compounds, such as sorgoleone, and

94

naphthoquinone-based compounds, such as juglone and lawsone, as well as anthraquinone-

95

based compounds,20 a number of previous studies have indicated that those compounds are

96

kinetically closely related to their redox potentials and functional groups, with kinetic

97

understanding of one providing a window to interpretation of associated chemistry and

98

biochemistry of the other.8, 11, 14, 20-22 2-Methoxyhydroquinone (MH2Q, see abbreviations of

99

different quinone species and detailed structure information in Fig.S1 of the Supporting

100

Information), as an intermediate product during lignin degradation, is a common constituent

101

of NOM and has the highest oxidoreductase activity among a series of low-molecular-weight

102

quinones.23, 24 This study therefore investigated the oxidation of MH2Q by Fe(III) under both

103

anaerobic and aerobic conditions over the pH range 4 – 6, with emphasis on development of a

104

rigorous kinetic description of the oxidation mechanism and insights on the role of free

105

radicals, such as semiquinone and superoxide. Thermodynamic characteristics of this system

106

were also discussed by employing the half-cell reduction potentials of different redox couples

107

to verify modeled rate constants of the reactions involving quinone species. Acidic conditions

108

were employed in this study partially to avoid complications associated with the precipitation

109

of iron oxides, but they do reflect conditions of acid mine drainage waters and coastal

110

environments with the presence of acid sulfate soils.

5 ACS Paragon Plus Environment

Environmental Science & Technology

2. MATERIALS AND METHODS

111 112

A complete description of the materials and methods used is presented in Section SI2 of the

113

Supporting Information (SI).

114

2.1. Reagents

115

All solutions were prepared using 18 MΩ.cm ultrapure Milli-Q water (Thermo Fisher

116

Scientific). Analytical grade chemicals were purchased from VWR International or Sigma-

117

Aldrich (or as otherwise stated) and used without further refinement. Stock solutions were

118

refrigerated at 4 °C in the dark when not in use. All studies were performed in 10 mM NaCl

119

at a controlled room temperature of 22 ± 0.5°C. Experiments were conducted in darkness

120

with the reactor covered in foil for the duration of the reaction. Anaerobic data were collected

121

in an anaerobic glove bag (Coy Laboratory Products, Grass Lake, MI) filled with 5% H2 and

122

95% N2. It is worth mentioning that all the anaerobic experiments conducted in the glove box

123

here were completed within 30 min, during which no interference from H2 was observed.

124

However, slow reduction of 2-methoxybenzoquinone (MBQ) stock solution to MH2Q during

125

an overnight stay in the chamber (more than 17 hr) was observed, suggesting that interference

126

reduction from H2 has to be considered for long-term experiments in the glove box.

127

2.2. Experimental Measurement

128

Measurement of MBQ: Concentrations of MBQ were determined spectrophotometrically

129

using a UV-Vis spectrophotometer (Thermo Fisher Scientific) with ε257 = 1.48 × 104 M−1

130

cm−1.

6 ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34

Environmental Science & Technology

131

Fe(II) Determination: Concentrations of total Fe(II) were determined using a modified FZ

132

method.17

133

H2O2 Determination: Previous studies21, 22 indicated the potential interference from

134

hydroquinones on H2O2 detection using the fluorescence method,25 therefore, a

135

chemiluminescence-based method (the AE-CL method)26 was employed for H2O2

136

measurement in this study.

137

2.3. Kinetic Modeling

138

The kinetic model was fitted to the experimental data over a range of experimental conditions

139

using the program Kintek Explorer.27 The sensitivity of the model to changes in individual

140

rate constant values was also assessed by examining the change in the relative difference

141

between the experimental data and the kinetic model (defined as the relative residual, r) when

142

one rate constant was varied while holding the others fixed at their optimal values. The

143

program Kintecus28 was employed to evaluate the kinetic model using a VBA program to

144

calculate r (see detailed description and Fig.S3 in SI3).

145 146 147 148

7 ACS Paragon Plus Environment

Environmental Science & Technology

149

3. RESULTS AND DISCUSSION

150

3.1 Anaerobic Oxidation of MH2Q by Fe(III)

151

Previous studies have investigated the redox chemistry of hydroquinones and metal species in

152

oxic environments,17, 18, 21, 22, 29 but interaction involving these active components under dark

153

anaerobic conditions is also of great interest, not only because such matrix composition better

154

features the redox circumstances in anoxic environments (groundwater, deep ocean,

155

sediments, et al.), but also the absence of O2 allows full expression of the redox functions of

156

metals and quinones and thus enables a thorough estimation of the generation of semiquinone

157

radicals under dark anaerobic conditions. Further, the reaction mechanism under anaerobic

158

condition is an essential component for the complete kinetic description of the oxidation

159

mechanism occurring in the presence of O2. Therefore, in this study, the oxidation of MH2Q

160

by Fe(III) was firstly investigated under anaerobic conditions, with results shown in Figures 1

161

– 2 (and Fig.S4 and S5). Precipitation of Fe(III) species in the system can be characterized as

162

insignificant in these experiments for several reasons: 1) the acidic conditions which enable

163

the existence of more soluble Fe(III) species, 2) the very low concentrations of Fe(III)

164

employed here, and 3) evidence that > 90% of added Fe(III) was quickly reduced to Fe(II) by

165

MH2Q before any significant nucleation/precipitation occurred (Figures 2, 4, and Fig.S8, also

166

see detailed explanation in section SI4). Therefore, no further consideration was given to

167

precipitation of Fe(III) in this study. Also, Fe(III) and Fe(II) were used as single entities to

168

represent all ferric and ferrous species respectively.

169

Figure 1 illustrates results for anaerobic oxidation of 500 nM MH2Q by different

170

concentrations of Fe(III) at pH 5. The steady-state concentrations of generated MBQ were

171

proportional to added [Fe(III)]0, with a specific ratio of 1:2. This is in accordance with

8 ACS Paragon Plus Environment

Page 8 of 34

Page 9 of 34

Environmental Science & Technology

172

previous studies30-32 reporting the overall stoichiometry for the oxidation of hydroquinone by

173

Fe(III) is: +

( − H ), kanaerobic MH 2 Q + 2Fe(III)  → MBQ + 2Fe(II)

174

(1)

175

which could be further confirmed by Fe(II) generation profiles shown in Fig.S6. At pH 5,

176

while 500 nM Fe(III) was completely reduced to Fe(II) at t = 20 min, only ~250nM of MBQ

177

was generated. pH dependence of MH2Q oxidation by Fe(III) was shown in Figure 2, as well

178

as in Fig.S4 and S5 of the SI. The generation rate of Fe(II), as can be seen from Figure 2,

179

increased significantly with increasing pH. For example, in the presence of 500 nM [MH2Q]0,

180

500 nM Fe(III) was completely reduced to Fe(II) within 5 min at pH 6, while less than 50%

181

of Fe(III) was reduced within the same period at pH 4. Similar trends could also be observed

182

in the generation rate of MBQ in Fig.S4 which showed that 250 nM [Fe(III)]0 was capable of

183

oxidizing ~120 nM MH2Q in the first 2 min at pH 6 but less than 20 nM MH2Q was oxidized

184

in the same duration at pH 4 (Fig.S4, and similar behaviour in Fig.S5). The pH dependence of

185

redox reactions between hydroquinones and ferric or cupric species has been previously

186

observed and was interpreted in terms of reactions via different ionized forms of

187

hydroquinones in various reaction routes.21, 22, 33-37 While the mono-dissociated hydroquinone

188

was postulated as the most kinetically active species towards the reduction of metal species,8,

189

21, 38

190

could also contribute to the reduction of cytochrome c by hydroquinone at acidic pH, which

191

is consistent with the conclusion from a recent study on copper-catalyzed oxidation of

192

naphthohydroquinone.22

studies by Rich and Bendall35, 36 reported that the undissociated hydroquinone species

9 ACS Paragon Plus Environment

Environmental Science & Technology

193

3.2 Aerobic Oxidation of MH2Q by Fe(III)

194

The autoxidation of MH2Q (autoxidation is oxidation in the absence of metal catalysts, see

195

results in section SI6) was found to be negligibly slow under the conditions of this study,

196

principally because of the lower pHs employed here and the high redox potential/pKa's

197

possessed by MH2Q,8, 22 making the first electron transfer from MH2Q to O2 highly

198

endergonic,4 thus this reaction was not further considered in this study. Fe(III)-mediated

199

oxidation of MH2Q in the presence of O2 was inspected, with emphasis on the effects of pH,

200

[Fe(III)]0 and [MH2Q]0, as shown in Figure 3. First and foremost, unlike the results obtained

201

under anaerobic conditions, non-stoichiometric oxidation, or in other words, catalytic

202

oxidation of MH2Q by Fe(III) was observed, especially in Figure 3(D), which demonstrated

203

that at pH 6, 500 nM Fe(III) was capable of completely oxidising 10 µM MH2Q within 2 hr.

204

Similar tests were also conducted on 1,4-benzohydroquinone and no significant catalytic

205

oxidation was found, presumably because the methoxy group substitution lowers the redox

206

potential of the hydroquinone and makes it more facile for oxidation by Fe(III).8 Since we

207

already confirmed that direct oxidation of MH2Q by O2 is negligible in this system, the rapid

208

oxidation of additional MH2Q must occur, at least partially, with mediation of iron.

209

It is also noteworthy that Figure 3 showed that Fe(III) oxidizes MH2Q to MBQ in a biphasic

210

manner, with initial rapid formation of MBQ followed by a much slower increase in MBQ

211

concentration over time. The initial concentration of MBQ that was rapidly formed was

212

approximately linear-proportional to the added Fe(III) concentration (similar to anaerobic

213

oxidation, where stoichiometric amount of MH2Q was quickly oxidized by added Fe(III)),

214

and thus was ascribed to direct oxidation of MH2Q by added Fe(III). This could be further

215

justified by Figure 4 and Fig.S8 showing Fe(III) was nearly completely reduced to Fe(II) by

216

the end of the initial rapid phase of oxidation, indicating the direct oxidation of MH2Q by

10 ACS Paragon Plus Environment

Page 10 of 34

Page 11 of 34

Environmental Science & Technology

217

Fe(III) is rapid and not rate limiting. Instead, the following much slower process, involving

218

multiple O2-mediated reactions, such as regeneration of Fe(III) by oxidizing Fe(II) at such

219

acidic pH,39, 40 might be more rate-determining in the overall oxidation.

220

On the other hand, it is interesting to notice that in Figure 3, the rate of oxidation is not only

221

correlated with the Fe(III) concentration, but also increased with increasing the initial

222

concentration of the hydroquinone ([MH2Q]0, Fig.S9), similar to behaviour observed in a

223

previous study on copper-mediated oxidation of naphthohydroquinone.22 More specifically,

224

as shown in Figure 3(D), at pH 6, more MH2Q was oxidized per unit of time when [MH2Q]0

225

= 10 µM than that with [MH2Q]0 = 5 µM; and similarly, comparison between Figures 3(A)-

226

(C) and Figure 3(D) indicates that at each pH with [Fe(III)]0 =500 nM, much more MH2Q

227

was oxidized per unit of time when [MH2Q]0 = 5 µM than that with [MH2Q]0 = 500 nM. This

228

was especially observed within the second oxidation phase, indicating that iron-mediated

229

reactions might not be the only factor dominating the “slow oxidation phase”, and instead

230

MH2Q-involved processes must also play a role. Nevertheless, as we declared that the direct

231

interaction between MH2Q and Fe(III) is rapid and thus not likely to be rate limiting, other

232

reactions involving intermediates generated from MH2Q-dependent activities, most likely the

233

semiquinone radicals,21, 22 might be more important in determining the slow oxidation rate.

234

Similar to anaerobic oxidation, there is also apparent pH dependence of the aerobic oxidation

235

of MH2Q by Fe(III) (as shown in Figure 3 and Fig.S7 in the SI), i.e., the overall oxidation

236

rate increased with increasing pH from 4 to 6, and such dependence was observed in both

237

oxidation phases (the initial rapid phase and the later slow phase). Apparently, the

238

mechanism accounting for the anaerobic pH dependence (involvement of different ionized

239

forms of MH2Q in reactions with ferric species), is inevitably contributing here, but other pH

11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 34

240

sensitive processes, such as the oxidation of Fe(II) by dioxygen and superoxide,19, 39, 40 could

241

also impact the overall pH dependence.

242

Besides the generation of MBQ, the transformation of iron in the system is also of great

243

interest. Direct measurement of Fe(II) concentrations over a range of conditions was

244

undertaken in order to assist in determining the mechanism and constraining the fitting

245

parameters. The corresponding results are shown in Figure 4 and Fig.S8. Within the pH range

246

4 – 6, all of Fe(III) added to 500 nM MH2Q solution was quickly reduced to Fe(II), with the

247

reduction rate increasing with increase in pH. As for pH 4 (Fig.S8), the concentration of Fe(II)

248

reached its maximum level after more than 30 min of the reaction initiation, while at pH 6,

249

the highest concentration was quickly achieved within 7 min (Figure 4). Though the re-

250

oxidation of Fe(II) to Fe(III) was happening concurrently,19, 39, 40 this reaction was too slow to

251

compete against the rapid reduction of Fe(III) to Fe(II) by MH2Q. This is to say, all of the

252

Fe(III) regenerated from the oxidation of Fe(II) by O2 and other oxidants (such as superoxide

253

and semiquinone radicals, see detailed mechanism in next section and in Table 1) was

254

immediately re-reduced to Fe(II) by MH2Q, thus almost all of the iron was observed as Fe(II)

255

in the system in spite of the predominating oxygenated conditions. The cycling between

256

Fe(III) and Fe(II) kept proceeding until MH2Q was totally consumed, and subsequently the

257

concentration of Fe(II) started to decline simply because of the oxygenation of Fe(II) to

258

Fe(III). This is clearly the case demonstrated in Figure 4 with the top curve representing the

259

reduction of 500 nM Fe(III) by 500 nM MH2Q at pH 6. After the pseudo-equilibrium

260

between Fe(II) and Fe(III) (8 – 60 min), the obvious decrease of [Fe(II)] began at ~ t = 60

261

min, at which point the added 500 nM of MH2Q was nearly completely transformed to MBQ

262

(Figure 3(C)).

12 ACS Paragon Plus Environment

Page 13 of 34

Environmental Science & Technology

263

In order to gain some insight into the Fe(III)/Fe(II) transformation in the system, the cycling

264

rate and the turnover frequency of iron have been calculated using a modified version of the

265

kinetic model reported in Table 1, with detailed description and results provided in Section

266

SI7 (Fig.S12 and S13). As an example, for the case of 500 nM Fe(III) and 10 µM MH2Q at

267

pH 6, cycling between Fe(III) and Fe(II) occurs at the very beginning of the reaction, with a

268

turnover frequency of ~6 h−1. By examining the contribution of different oxidants on the

269

oxidation of MH2Q (Fig. S13), it is clear that besides Fe(III)/Fe(II), the O2/O2•− couple, which

270

oxidized almost 40% of MH2Q, is another significant force driving the overall oxidation. This

271

also explains the occurrence of non-stoichiometric oxidation of MH2Q by Fe(III) in the

272

presence of O2. The additional MH2Q was oxidized not only by the regenerated Fe(III) that

273

resulted from oxidation of Fe(II), but also by direct reaction with O2•−, generated either from

274

reduction of O2 by semiquinone or by Fe(II) (reactions 6 and 8, respectively, Table 1).

275

Previous studies also suggested that during the metal-catalyzed oxidation of hydroquinones,

276

stoichiometric amounts of hydrogen peroxide (H2O2) would be generated.8, 22, 41-44 H2O2 is of

277

particular interest because of its ubiquitousness both in the cells of organisms as a by-product

278

of oxygen metabolism, and in surface seawaters primarily due to photochemical processes

279

mediated by NOM.45 A recent study by Murphy, et al.46 also reported rapid production of

280

H2O2 (with maximum concentrations of 3.85 µM) when anoxic groundwater mixed with

281

oxidized, iron rich sediments, indicating a metal redox cycling mechanism under dark

282

conditions. More importantly, in the presence of iron, H2O2 may lead to the formation of

283

highly reactive and damaging oxidants such as hydroxyl radical (HO•), a radical which

284

always present in vanishingly small concentration but will in turn induce the oxidation of

285

both NOM and anthropogenic contaminants.46-48 Therefore the generation of H2O2 during the

286

oxidation of MH2Q by Fe(III) at different pH values was also investigated, with results

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 34

287

shown in Figure 5. Apparently, the rate of H2O2 generation, similar to the oxidation of MH2Q,

288

also exhibited strong pH dependence, with more rapid H2O2 generated at pH 6 than at pH 4.

289

3.3 Kinetic Modelling of MH2Q Oxidation by Fe(III)

290

Since the dissociation states of a hydroquinone is of key importance for its electron-donating

291

capability, the protonation/deprotonation reactions of MH2Q were included in the model as

292

reactions 1 and 2 in Table 1, with previously reported ionization constants.20, 49 Under

293

anaerobic conditions, the oxidation of MH2Q by Fe(III) occurs via direct interactions between

294

Fe(III) and hydroquinone species, and both the undissociated and mono-dissociated

295

hydroquinones have been reported to be active towards reduction of ferric species, as

296

expressed by reactions 3 and 4 in Table 1.34, 36 Although the doubly dissociated species MQ2–

297

is thermodynamically more reactive, its reaction with Fe(III) is negligible in this study

298

considering its extremely low concentration (10–16 – 10–20 M, from pH 6 to 4 under

299

conditions of this study), and thus was not included in the model. The intermediate generated

300

from one electron oxidation of the hydroquinone, semiquinone radical, has been widely

301

proposed to be inevitably involved in the pool with both oxidizing and reducing capibilities.8,

302

18, 21, 34-37

303

included in the model as reaction 5, as well as the oxidation of Fe(II) by MSQ•− (back

304

reaction of reactions 3 and 4) with a reported rate of 1 × 105 M−1·s−1 in recent studies by Garg,

305

et al.17, 18, 29 on iron redox transformations in the presence of natural organic matter.

306

When O2 is present, in addition to those reactions occurring anaerobically, reactions

307

describing the electron transfers between oxygen and semiquinone radical, between

308

superoxide and quinone species, as well as reactions depicting the oxidation of Fe(II), must

309

also be included. The additional reactions included in the model due to the presence of O2

Therefore, the reduction of Fe(III) by 2-methoxysemiquinone radical (MSQ•−) was

14 ACS Paragon Plus Environment

Page 15 of 34

Environmental Science & Technology

310

with corresponding rate constants are listed as reactions 6 – 9 in Table 1. As previously stated,

311

the autoxidation of MH2Q is negligibly slow under conditions of this study, and thus was not

312

required in the model. Because O2 has no charge, Marcus theory of electron transfer predicts

313

that rate constants for reaction of semiquinone radical with O2 will be a function of change in

314

the Gibbs free energy.8, 15 It has also been reported that for quinones with a standard mid-

315

point potential of E0(Quinone/Semiquinone) ≥ − 150 mV, rate constants for the forward

316

reactions follow well the Marcus theory for electron transfer and can be estimated by a

317

reported linear relationship.8, 36, 50 Based on this and E0(MBQ/MSQ•−) = − 0.017 V (see

318

detailed calculations in SI8), the rate constant for the forward reaction can be estimated as k6

319

= 1.5 × 106 M−1·s−1. The equilibrium constant (K) of reaction 6 can also be determined by the

320

standard reduction potential (E0) of both redox couples (MBQ/MSQ•− and O2/O2•−) using the

321

Nernst equation:55

322

0 E(O

- E0

 -(MBQ/MSQ ) 2 /O 2 )

=

RT ln K nF

(2)

323

where R is the gas constant in appropriate units, 8.314 J K−1 mol−1. T is the temperature in

324

Kelvin, F is the Faraday constant 9.6485 × 104 C mol−1 and n is the number of electrons

325

involved in the reaction. Thus, when n = 1, at 25 °C, Equation (4) can be written as:

326

log K =

0 E(O

- E0

 -(MBQ/MSQ ) 2 /O 2 )

(3)

0.059

327

Here we employed the values of E0(O2/O2•−) = − 0.155 V56-58 and E0(MBQ/MSQ•−) = − 0.017

328

V in equation (3) to calculate K6 = 4.58 × 10−3, thus k−6 = k6/K6 = 3.3 × 108 M−1·s−1. These

329

values agree very well with previous reported rate constants of methyl-hydroquinone (k6 =

330

1.2 × 106 M−1·s−1 and k−6 = 7.3 × 108 M−1·s−1),50 which has close redox potentials to MH2Q.

331

Further, because of the substituent effect, the redox potential of MH2Q lies between that of

332

1,4-benzohydroquinone and 1,4-naphthohydroquinone, both the forward and reverse rate 15 ACS Paragon Plus Environment

Environmental Science & Technology

333

constants of reaction 6 are also between these two analogous aromatics,21, 22 which again

334

confirms that the redox potential of the quinone/semiquinone couple governs the reaction of

335

semiquinone with dioxygen to form superoxide.8

336

The comproportionation reaction between MH2Q and MBQ was not important in the system,

337

as a result of the extremely low concentrations of the doubly dissociated hydroquinone

338

species (MQ2−, the well-established dominant active species towards comproportionation

339

reaction,15, 34, 36, 49 having a concentration of ~10–16 – 10–20 M from pH 6 to 4 under

340

conditions of this study). Since MSQ•− can be both oxidizing and reducing, it will readily

341

disproportionate to corresponding hydroquinone and quinone. However, due to its low

342

concentration and its rapid removal by dioxygen and iron species (reactions 3 – 6, Table 1),

343

the disproportionation of MSQ•− plays no significant role here. This could be further verified

344

by the result that inclusion of these two reactions in the kinetic model with varying rate

345

constants from 10 – 1010 M−1·s−1 has no observable effect on the overall model fitting

346

(Fig.S3(G) and (H) in SI). Therefore, these two reactions were not included in the model.

347

The oxidation of MH2Q by superoxide, which was expected to have a relatively low reaction

348

rate,58 was included in the model as reaction 7. Previous studies58-60 also suggested that

349

reactions involving H2O2 in this system (such as the backward reaction of reaction 7, Table 1)

350

were not important, presumably because the reaction rate between H2O2 and semiquinones

351

(~104 − 105 M−1·s−1) is low.61, 62 Therefore, this reaction is unlikely to compete effectively

352

with the rapid removal of semiquinones by dioxygen (reaction 6, Table 1). Further evidence

353

for this was given by the sensitivity analysis, which demonstrated that inclusion of this

354

reaction in the kinetic model and varying the rate constant from 10 – 105 M−1·s−1 had

355

negligible effect on the model fitting (see Fig. S3(I) in SI). As such, this reaction was

356

excluded from the model. The disproportionation of O2•−, with rates of 6 × 106 and 6 × 108 16 ACS Paragon Plus Environment

Page 16 of 34

Page 17 of 34

Environmental Science & Technology

357

M−1·s−1 at pH 6 and 4 respectively,63, 64 was not important here because of its low

358

concentration and rapid reaction with iron and quinone species (reactions 6 – 9 in Table 1).

359

This was also confirmed by the sensitivity analysis (see Fig.S3(J) in SI). Cross-dismutation

360

between MSQ•− and O2•− might also occur, but little information on this reaction has been

361

previously reported and the rate constant is not known. Inclusion of this reaction in

362

simulations and varying the rate constant over a reasonable range revealed minimal effect on

363

the fitting (Fig. S3(K) in SI), justifying exclusion of this reaction from the model. The

364

oxidation of Fe(II) (reactions 8 and 9), as the key route for regeneration of Fe(III) in the

365

presence of O2, has been extensively investigated at acidic and circumneutral pHs. Therefore,

366

here we employed the well-established pH-dependent rates from previous studies19, 39, 40, 52-54

367

for interactions between O2, Fe(II)/Fe(III) and O2•−.

368

Due to the limited information available in the literature, rate constants for the interactions of

369

Fe(III) and quinone species (reactions 3, 4 and 5, Table 1) and the reaction between MH2Q

370

and O2•− (reaction 7, Table 1) were determined empirically by best fit of the model to the

371

experimental data. Results of model fitting over the pH range 4 – 6 using reactions 1 – 9

372

(Table 1) are illustrated in Figures 1 – 5 (and Fig.S4-S9 in the SI). In general, the model

373

provides reasonable fits to the experimental data over a range of conditions. The good

374

agreement between the model output and the experimental data validates our assumptions of

375

the proposed mechanism and the assignment of individual rate constants.

376

In accordance with previous studies,34-36 both the undissociated and singly dissociated MH2Q

377

species (MH2Q0 and MHQ−) were found to be kinetically active towards Fe(III) reduction

378

with intrinsic rate constants of 2.4 × 103 and 3.0 × 108 M−1·s−1, respectively. Since the

379

standard one-electron reduction potential of the MSQ•−/MH2Q0 couple is much higher than

380

that of the MSQ•−/MHQ− couple (see detailed calculations in SI8), reaction 3 is less 17 ACS Paragon Plus Environment

Environmental Science & Technology

381

favourable in a thermodynamic sense and occurs via a rather slower rate than that of reaction

382

4. This again confirms that the mono-dissociated hydroquinone anion is a substantially more

383

facile electron donor to ferric species.8, 34, 36 Although MHQ− only occupies a small fraction

384

of the total hydroquinone species at acidic pH, the protonation/deprotonation equilibrium is

385

faster than that of redox reactions and thus is not rate-limiting.35, 65 It is also worth

386

mentioning that, although thermodynamically less favourable, reaction 3 is still feasible. We

387

can conclude here that the back reaction is more rapid than the forward reaction, since the

388

rate of reaction between Fe(II) and semiquinone radical is almost two orders of magnitude

389

higher than that of the forward reaction. Moreover, the rapid removal of the product, MSQ•−,

390

by Fe(III) and O2 (reactions 5 and 6, respectively) is another significant force driving the

391

oxidation reaction to the right.8, 35 Thus, the presence of O2 in the system not only promotes

392

Fe(III) re-supply by continuously oxidizing Fe(II), but also induces the rapid removal of

393

MSQ•− with resultant generation of O2•−. The model also indicates that the rate of MSQ•−

394

removal by O2 > rate of removal by Fe(III) > rate of removal by Fe(II) > rate of removal by

395

disproportionation, consistent with previously reported conclusion.36 However, unlike the

396

study by Uchimiya and Stone,11 there is no back reduction of MBQ by Fe(II) observed in this

397

study, presumably due to the acidic pH and low concentrations of both reactants employed in

398

this study.

399

The model predicted rate of 1.2 × 107 M−1·s−1 for reduction of Fe(III) by MSQ•− (reaction 5,

400

Table 1) is higher than the value of 2 × 106 M−1·s−1 reported by Yamazaki and Ohnishi34 on

401

the reduction of ferricytochrome c by semiquinones, but lower than the value of 2 × 107

402

M−1·s−1 reported in a recent study on the reduction of cupric species by 1,4-semiquinone.21

403

Furthermore, MSQ•−, with a standard one-electron reduction potential of E0(MBQ/MSQ•−) =

404

− 0.017 V, is more reducing than oxidizing. This is to say, in a system with both Fe(III) and

18 ACS Paragon Plus Environment

Page 18 of 34

Page 19 of 34

Environmental Science & Technology

405

Fe(II) present, instead of shifting back to MH2Q by oxidizing Fe(II), MSQ•− is more likely to

406

transfer to MBQ by reducing Fe(III), consolidating the model predicted k5 > k-3. As indicated

407

in the study by Eyer,58 the oxidation of MH2Q by superoxide, was expected to have a

408

relatively low reaction rate of ~104 – 105 M−1·s−1.66, 67, 68 Our model predicted rate of 7 × 105

409

M−1·s−1 is also close to the value of 2 × 105 M−1·s−1 reported in recent studies by Garg, et al.17

410

on iron redox transformations in the presence of NOM under acid conditions.

411

Despite the capability of the kinetic model in providing a good description of the various

412

reactions between the quinone triad (MBQ, MSQ•− and MH2Q), Fe(III)/Fe(II), and O2 over a

413

range of experimental conditions with a relatively well-constrained set of rate constants, the

414

model is limited by the assumption that the composite terms Fe(II) and Fe(III) adequately

415

represent all inorganic iron in their +2 and +3 oxidation states, respectively. Therefore, care

416

must be taken when interpreting the model results, since the model is condition-specific by

417

default. For example, the presence of complexing organic ligands would influence the

418

reactivity of various iron species and as such the composite “intrinsic” rate constants of

419

various reactions with iron will need to be re-evaluated.

420

3.4 Implication of Findings

421

In this study, we investigated the oxidation of MH2Q by Fe(III) under dark conditions in the

422

absence and presence of oxygen. While Fe(III) was capable of stoichiometrically oxidizing

423

MH2Q under anaerobic conditions, catalytic oxidation of MH2Q was observed in the presence

424

of O2 due to further cycling between iron, oxygen and semiquinone radicals. Based on the

425

experimental observations, a detailed kinetic model was developed to describe the

426

predominant mechanisms operating in this system and to provide insights into the observed

427

pH dependence and the role of generated radical intermediates. Our results also showed that

19 ACS Paragon Plus Environment

Environmental Science & Technology

428

the direct oxidation of MH2Q by O2 is spin-restricted and very slow under conditions of this

429

study, but the Fe(III)/Fe(II) couple could overcome such restriction and perform as an

430

electron transfer agent between the hydroquinone and O2 (similar to previous observed

431

behaviours of copper species21, 22), leading to the existence of pseudo-stable but high

432

concentrations of reduced Fe(II) under oxic and dark conditions that could potentially

433

promote the generation of ROS. One particular example in this study is the generation of O2•−

434

and H2O2 from the redox cycling of iron between its +3 and +2 oxidation states in the

435

presence of MH2Q and O2 confirms the essential role of NOM and reduced metals in non-

436

photochemical pathway for ROS formation in natural aquatic environments,69 which, at the

437

same time, are particularly important for the oxidative degradation of NOM itself and any

438

further cascade of radical-promoted reactions with other constituents in natural waters.

439

Though the presence of O2 is imperative for iron cycling and chain propagating to reach

440

catalytic oxidation of MH2Q in this study, in deep ocean/lakes and groundwater, where O2

441

deficiency prevails, other components presented in such anoxic/anaerobic environments

442

could serve to bridge Fe(II) and Fe(III), and thus facilitate iron cycling and further redox of

443

quinones and NOM. For instance, Uchimiya and Stone11 reported the oxidation of Fe(II) by

444

various quinones at circumneutral pH could serve as a vital link in networks of electron

445

transport in suboxic and anoxic environments. Studies also showed chromium(VI) and

446

manganese(III, IV) (hydr)oxides were capable of oxidising Fe(II) under anoxic conditions.70,

447

71

448

communities may also contribute, especially in low pH environments.72 Therefore, the

449

coexistence of quinones and iron could potentially facilitate the redox cycling of both species

450

in allowable environments, and separately or in concert, iron species (both dissolved and

At the same time, the presence of fluctuating redox conditions and (micro)biological

20 ACS Paragon Plus Environment

Page 20 of 34

Page 21 of 34

Environmental Science & Technology

451

precipitated) and quinones are believed to be pivotal participants in a wide range of

452

environmental redox reactions.

453

Due to the close correlations between kinetic/thermodynamic characteristics of quinone triads

454

and their structures and functional groups, the kinetic model and associated rate constants

455

reported in this study could provide insight into the rate and extent of interactions between

456

other quinone members and metals that have different electrophilicity and redox potentials.

457

On the other hand, while quinone moieties are recognized to dominate the redox properties of

458

NOM, care must be taken when interpreting the results observed in pure quinone solutions to

459

the natural aquatic environment, where the presence of an extended network of complexing

460

organic ligands would influence the reactivity of various iron species and the stability of

461

radicals. For example, similarities and differences in iron redox transformations between a

462

simple quinone and the quinone moieties presented in SRFA were identified in recent studies

463

by Jiang et al.73 and Garg, et al.,18 where increased stability of semiquinone radicals were

464

found in SRFA solution compared to that in pure hydroquinone solution, indicating quinone

465

moieties presented in highly conjugated and substituted environment of NOM will play a

466

much more important role in the rate and extent of metal redox transformations.

21 ACS Paragon Plus Environment

Environmental Science & Technology

467

4. ACKNOWLEDGEMENTS

468

The work reported here is supported as part of the Sustainable Systems (SS) Scientific Focus

469

Area (SFA) program at LBNL, funded by the U.S. Department of Energy, Office of Science,

470

Office of Biological and Environmental Research, Subsurface Biogeochemical Research

471

Program, under Award Number DE-AC02-05CH11231. The authors gratefully acknowledge

472

the assistance from Dr. Christopher Miller at the University of New South Wales (Australia)

473

for the model sensitivity calculations.

474

ASSOCIATED CONTENT

475

Supporting Information

476

Additional detailed information on experimental methods, manipulative experiments,

477

sensitivity analysis and extra experimental data and calculation. This material is available

478

free of charge via the Internet at http://pubs.acs.org.

479 480 481 482 483

22 ACS Paragon Plus Environment

Page 22 of 34

Page 23 of 34

Environmental Science & Technology

484

REFERENCES

485 486 487

1. Scott, D. T.; McKnight, D. M.; Blunt-Harris, E. L.; Kolesar, S. E.; Lovley, D. R., Quinone moieties act as electron acceptors in the reduction of humic substances by humicsreducing microorganisms. Environ. Sci. Technol. 1998, 32, (19), 2984-2989.

488 489

2. Aeschbacher, M.; Graf, C.; Schwarzenbach, R. P.; Sander, M., Antioxidant properties of humic substances. Environ. Sci. Technol. 2012, 46, (9), 4916-4925.

490 491 492

3. Nurmi, J. T.; Tratnyek, P. G., Electrochemical Properties of Natural Organic Matter (NOM), Fractions of NOM, and Model Biogeochemical Electron Shuttles. Environ. Sci. Technol. 2002, 36, (4), 617-624.

493 494 495

4. Aeschbacher, M.; Sander, M.; Schwarzenbach, R. P., Novel electrochemical approach to assess the redox properties of humic substances. Environ. Sci. Technol. 2010, 44, (1), 8793.

496 497

5. Chen, J.; Gu, B.; Royer, R. A.; Burgos, W. D., The roles of natural organic matter in chemical and microbial reduction of ferric iron. Sci. Total Environ. 2003, 307, (1–3), 167-178.

498 499 500

6. Lovley, D. R.; Coates, J. D.; Blunt-Harris, E. L.; Phillips, E. J.; Woodward, J. C., Humic substances as electron acceptors for microbial respiration. Nature 1996, 382, (6590), 445-448.

501 502

7. Lovley, D. R.; Woodward, J. C.; Chapelle, F., Stimulated anoxic biodégradation of aromatic hydrocarbons using Fe (in) ligands. Nature 1994, 370, 128-131.

503 504 505

8. Song, Y.; Buettner, G. R., Thermodynamic and kinetic considerations for the reaction of semiquinone radicals to form superoxide and hydrogen peroxide. Free Radical Biol. Med. 2010, 49, (6), 919-962.

506 507

9. Uchimiya, M.; Stone, A. T., Reversible redox chemistry of quinones: Impact on biogeochemical cycles. Chemosphere 2009, 77, (4), 451-458.

508 509 510 511

10. Laviron, E., Electrochemical reactions with protonations at equilibrium: Part VIII. The 2e, 2H+ reaction (nine-member square scheme) for a surface or for a heterogeneous reaction in the absence of disproportionation and dimerization reactions. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1983, 146, (1), 15-36.

512 513

11. Uchimiya, M.; Stone, A. T., Reduction of substituted p-benzoquinones by Fe II near neutral pH. Aquat Geochem 2010, 16, (1), 173-188.

514 515 516

12. Hong, Y.; Sengupta, S.; Hur, W.; Sim, T., Identification of Novel ROS Inducers: Quinone Derivatives Tethered to Long Hydrocarbon Chains. J. Med. Chem. 2015, 58, (9), 3739-3750.

23 ACS Paragon Plus Environment

Environmental Science & Technology

517 518 519

13. Paul, A.; Stösser, R.; Zehl, A.; Zwirnmann, E.; Vogt, R. D.; Steinberg, C. E. W., Nature and abundance of organic radicals in natural organic matter: Effect of pH and irradiation. Environ. Sci. Technol. 2006, 40, (19), 5897-5903.

520 521

14. Uchimiya, M.; Stone, A. T., Redox reactions between iron and quinones: Thermodynamic constraints. Geochim. Cosmochim. Acta 2006, 70, (6), 1388-1401.

522 523 524

15. Roginsky, V.; M. Pisarenko, L.; Bors, W.; Michel, C., The kinetics and thermodynamics of quinone-semiquinone-hydroquinone systems under physiological conditions. J. Chem. Soc., Perkin Trans. 2 1999, (4), 871-876.

525 526 527

16. Aeschbacher, M.; Vergari, D.; Schwarzenbach, R. P.; Sander, M., Electrochemical analysis of proton and electron transfer equilibria of the reducible moieties in humic acids. Environ. Sci. Technol. 2011, 45, (19), 8385-8394.

528 529 530

17. Garg, S.; Ito, H.; Rose, A. L.; Waite, T. D., Mechanism and kinetics of dark iron redox transformations in previously photolyzed acidic natural organic matter solutions. Environ. Sci. Technol. 2013, 47, (4), 1861-1869.

531 532 533

18. Garg, S.; Jiang, C.; David Waite, T., Mechanistic insights into iron redox transformations in the presence of natural organic matter: Impact of pH and light. Geochim. Cosmochim. Acta 2015, 165, 14-34.

534 535 536

19. Pham, A. N.; Waite, T. D., Oxygenation of Fe(II) in natural waters revisited: Kinetic modeling approaches, rate constant estimation and the importance of various reaction pathways. Geochim. Cosmochim. Acta 2008, 72, (15), 3616-3630.

537 538

20. Uchimiya, M.; Stone, A. T., Aqueous oxidation of substituted dihydroxybenzenes by substituted benzoquinones. Environ. Sci. Technol. 2006, 40, (11), 3515-3521.

539 540 541

21. Yuan, X.; Pham, A. N.; Miller, C. J.; Waite, T. D., Copper-catalyzed hydroquinone oxidation and associated redox cycling of copper under conditions typical of natural saline waters. Environ. Sci. Technol. 2013, 47, (15), 8355-8364.

542 543 544

22. Yuan, X.; Miller, C. J.; Pham, A. N.; Waite, T. D., Kinetics and mechanism of autoand copper-catalyzed oxidation of 1,4-naphthohydroquinone. Free Radical Biol. Med. 2014, 71, (0), 291-302.

545 546

23. Kirk, T. K.; Lorenz, L. F., Methoxyhydroquinone, an Intermediate of Vanillate Catabolism by Polyporus dichrous. Applied Microbiology 1973, 26, (2), 173-175.

547 548 549

24. Buswell, J. A.; Hamp, S.; Eriksson, K. E., Intracellular quinone reduction in Sporotrichum pulverulentum by a NAD(P)H: Quinone oxidoreductase: Possible role in vanillic acid catabolism. FEBS Lett. 1979, 108, (1), 229-232.

550 551 552 553

25. Zhou, M.; Diwu, Z.; Panchuk-Voloshina, N.; Haugland, R. P., A stable nonfluorescent derivative of resorufin for the fluorometric determination of trace hydrogen peroxide: Applications in detecting the activity of phagocyte NADPH oxidase and other oxidases. Anal. Biochem. 1997, 253, (2), 162-168.

24 ACS Paragon Plus Environment

Page 24 of 34

Page 25 of 34

Environmental Science & Technology

554 555 556 557

26. King, D. W.; Cooper, W. J.; Rusak, S. A.; Peake, B. M.; Kiddle, J. J.; O'Sullivan, D. W.; Melamed, M. L.; Morgan, C. R.; Theberge, S. M., Flow injection analysis of H2O2 in natural waters using acridinium ester chemiluminescence:  Method development and optimization using a kinetic model. Anal. Chem. 2007, 79, (11), 4169-4176.

558 559 560

27. Johnson, K. A.; Simpson, Z. B.; Blom, T., Global Kinetic Explorer: A new computer program for dynamic simulation and fitting of kinetic data. Anal. Biochem. 2009, 387, (1), 20-29.

561

28.

562 563 564

29. Garg, S.; Jiang, C.; Miller, C. J.; Rose, A. L.; Waite, T. D., Iron redox transformations in continuously photolyzed acidic solutions containing natural organic matter: Kinetic and mechanistic insights. Environ. Sci. Technol. 2013, 47, (16), 9190-9197.

565 566

30. Pracht, J.; Boenigk, J.; Isenbeck-Schröter, M.; Keppler, F.; Schöler, H. F., Abiotic Fe(III) induced mineralization of phenolic substances. Chemosphere 2001, 44, (4), 613-619.

567 568 569

31. Baxendale, J.; Hardy, H., Kinetics and equilibria in solutions containing ferrous ion, ferric ion, and some substituted hydroquinones and Quinones. Transactions of the Faraday Society 1954, 50, 808-814.

570 571

32. Anschutz, A. J.; Penn, R. L., Reduction of crystalline iron(III) oxyhydroxides using hydroquinone: Influence of phase and particle size. Geochem. Trans. 2005, 6, (3), 60-66.

572 573

33. Williams, G. R., The reduction of cytochrome c by hydroquinone. Can. J. Biochem. Phys. 1963, 41, (1), 231-237.

574 575 576

34. Yamazaki, I.; Ohnishi, T., One-electron-transfer reactions in biochemical systems I. Kinetic analysis of the oxidation-reduction equilibrium between quinol-quinone and ferroferricytochrome c. Biochim. Biophys. Acta 1966, 112, (3), 469-481.

577 578 579

35. Rich, P. R.; Bendall, D. S., A mechanism for the reduction of cytochromes by quinols in solution and its relevance to biological electron transfer reactions. FEBS Lett. 1979, 105, (2), 189-194.

580 581 582

36. Rich, P. R.; Bendall, D. S., The kinetics and thermodynamics of the reduction of cytochrome c by substituted p-benzoquinols in solution. Biochim. Biophys. Acta 1980, 592, (3), 506-518.

583 584

37. Rich, P. R., Electron transfer reactions between quinols and quinones in aqueous and aprotic media. Biochim. Biophys. Acta 1981, 637, (1), 28-33.

585 586

38. Baxendale, J.; Hardy, H.; Sutcliffe, L., Kinetics and equilibria in the system ferrous ion+ ferric ion+ hydro-quinone+ quinone. Trans. Faraday Soc. 1951, 47, 963-973.

587 588 589

39. Jones, A. M.; Griffin, P. J.; Waite, T. D., Ferrous iron oxidation by molecular oxygen under acidic conditions: The effect of citrate, EDTA and fulvic acid. Geochim. Cosmochim. Acta 2015, 160, 117-131.

Ianni, J. C., Kintecus, Windows version 3.9. 2006.

25 ACS Paragon Plus Environment

Environmental Science & Technology

590 591 592

40. Jones, A. M.; Griffin, P. J.; Collins, R. N.; Waite, T. D., Ferrous iron oxidation under acidic conditions – The effect of ferric oxide surfaces. Geochim. Cosmochim. Acta 2014, 145, (0), 1-12.

593 594 595

41. Li, Y.; Kuppusamy, P.; Zweier, J. L.; Trush, M. A., ESR evidence for the generation of reactive oxygen species from the copper-mediated oxidation of the benzene metabolite, hydroquinone: role in DNA damage. Chem.-Biol. Interact. 1995, 94, (2), 101-120.

596 597

42. Mandal, S.; Kazmi, N. H.; Sayre, L. M., Ligand dependence in the copper-catalyzed oxidation of hydroquinones. Arch. Biochem. Biophys. 2005, 435, (1), 21-31.

598 599

43. Li, Y. B.; Trush, M. A., Oxidation of hydroquinone by copper: Chemical mechanism and biological effects. Arch. Biochem. Biophys. 1993, 300, (1), 346-355.

600 601 602

44. Li, Y. B.; Trush, M. A., DNA damage resulting from the oxidation of hydroquinone by copper: role for a Cu(II)/Cu(I) redox cycle and reactive oxygen generation. Carcinogenesis 1993, 14, (7), 1303-1311.

603 604 605

45. Miller, C. J.; Vincent Lee, S. M.; Rose, A. L.; Waite, T. D., Impact of natural organic matter on H2O2-mediated oxidation of Fe(II) in coastal seawaters. Environ. Sci. Technol. 2012, 46, (20), 11078-11085.

606 607 608

46. Murphy, S. A.; Solomon, B. M.; Meng, S.; Copeland, J. M.; Shaw, T. J.; Ferry, J. L., Geochemical production of reactive oxygen species from biogeochemically reduced Fe. Environ. Sci. Technol. 2014, 48, (7), 3815-3821.

609 610 611

47. Pignatello, J. J.; Oliveros, E.; MacKay, A., Advanced Oxidation Processes for Organic Contaminant Destruction Based on the Fenton Reaction and Related Chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, (1), 1-84.

612 613 614

48. Miller, C. J.; Rose, A. L.; Waite, T. D., Hydroxyl radical production by H2O2mediated oxidation of Fe(II) complexed by suwannee river fulvic acid under circumneutral freshwater conditions. Environ. Sci. Technol. 2012, 47, (2), 829-835.

615 616

49. Bishop, C.; Tong, L., Equilibria of substituted semiquinones at high pH. J. Am. Chem. Soc. 1965, 87, (3), 501-505.

617 618

50. Meisel, D., Free energy correlation of rate constants for electron transfer between organic systems in aqueous solutions. Chem. Phys. Lett. 1975, 34, (2), 263-266.

619 620 621

51. King, D. W.; Lounsbury, H. A.; Millero, F. J., Rates and Mechanism of Fe(II) Oxidation at Nanomolar Total Iron Concentrations. Environ. Sci. Technol. 1995, 29, (3), 818824.

622 623 624

52. Rush, J. D.; Bielski, B. H., Pulse radiolytic studies of the reactions of HO2/O2 with Fe (II)/Fe (III) ions. The reactivity of HO2/O2 with ferric ions and its implication on the occurrence of the Haber-Weiss reaction. J. Phys. Chem. 1985, 89, (23), 5062-5066.

625 626

53. Wehrli, B., Redox reactions of metal ions at mineral surfaces. Aquatic chemical kinetics 1990, 311-336. 26 ACS Paragon Plus Environment

Page 26 of 34

Page 27 of 34

Environmental Science & Technology

627 628 629

54. Jones, A. M.; Griffin, P. J.; Collins, R. N.; Waite, T. D., Ferrous iron oxidation under acidic conditions – The effect of ferric oxide surfaces. Geochim. Cosmochim. Acta 2014, 145, 1-12.

630 631

55. Morel, F. M. M.; Hering, J. G., Principles and applications of aquatic chemistry. Wiley: New York, 1993.

632 633

56. Meisel, D.; Czapski, G., One-electron transfer equilibriums and redox potentials of radicals studied by pulse radiolysis. J. Phys. Chem. 1975, 79, (15), 1503-1509.

634 635 636

57. Buettner, G. R., The pecking order of free-radicals and antioxidants - lipidperoxidation, alpha-tocopherol, and ascorbate. Arch. Biochem. Biophys. 1993, 300, (2), 535543.

637 638

58. Eyer, P., Effects of superoxide dismutase on the autoxidation of 1,4-hydroquinone. Chem.-Biol. Interact. 1991, 80, (2), 159-176.

639 640

59. Ishii, T.; Fridovich, I., Dual effects of superoxide dismutase on the autoxidation of 1,4-naphthohydroquinone. Free Radical Biol. Med. 1990, 8, (1), 21-24.

641 642 643

60. Mochizuki, M.; Yamazaki, S.-i.; Kano, K.; Ikeda, T., Kinetic analysis and mechanistic aspects of autoxidation of catechins. Biochimica et Biophysica Acta (BBA) General Subjects 2002, 1569, (1-3), 35-44.

644 645 646

61. Kalyanaraman, B.; Sealy, R. C.; Sinha, B., An electron spin resonance study of the reduction of peroxides by anthracycline semiquinones. Biochimica et Biophysica Acta (BBA)General Subjects 1984, 799, (3), 270-275.

647 648 649

62. Zhu, B.-Z.; Zhao, H.-T.; Kalyanaraman, B.; Frei, B., Metal-independent production of hydroxyl radicals by halogenated quinones and hydrogen peroxide: an ESR spin trapping study. Free Radical Biol. Med. 2002, 32, (5), 465-473.

650 651 652

63. Zafiriou, O. C., Chemistry of superoxide ion-radical (O2-) in seawater. I. pKasw* (HOO) and uncatalyzed dismutation kinetics studied by pulse radiolysis. Mar. Chem. 1990, 30, 31-43.

653 654 655

64. Yuan, X.; Pham, A. N.; Xing, G.; Rose, A. L.; Waite, T. D., Effects of pH, chloride, and bicarbonate on Cu(I) oxidation kinetics at circumneutral pH. Environ. Sci. Technol. 2012, 46, (3), 1527-1535.

656 657 658

65. Adams, G.; Michael, B., Pulse radiolysis of benzoquinone and hydroquinone. Semiquinone formation by water elimination from trihydroxy-cyclohexadienyl radicals. Trans. Faraday Soc. 1967, 63, 1171-1180.

659 660

66. Roginsky, V.; Barsukova, T., Kinetics of oxidation of hydroquinones by molecular oxygen. Effect of superoxide dismutase. J. Chem. Soc., Perkin Trans. 2 2000, (7), 1575-1582.

661 662 663

67. Kitagawa, S.; Fujisawa, H.; Sakurai, H., Scavenging effects of dihydric and polyhydric phenols on superoxide anion radicals, studied by electron-spin-resonance spectrometry. Chem. Pharm. Bull. (Tokyo) 1992, 40, (2), 304-307. 27 ACS Paragon Plus Environment

Environmental Science & Technology

664 665 666 667

68. Deeble, D. J.; Parsons, B. J.; Phillips, G. O.; Schuchmann, H.-P.; Von Sonntag, C., Superoxide radical reactions in aqueous solutions of pyrogallol and N-propyl gallate: The involvement of phenoxyl radicals. A pulse radiolysis study. Int. J. Radiat. Biol. 1988, 54, (2), 179-193.

668 669 670

69. Page, S. E.; Kling, G. W.; Sander, M.; Harrold, K. H.; Logan, J. R.; McNeill, K.; Cory, R. M., Dark Formation of Hydroxyl Radical in Arctic Soil and Surface Waters. Environ. Sci. Technol. 2013, 47, (22), 12860-12867.

671 672

70. Buerge, I. J.; Hug, S. J., Influence of Organic Ligands on Chromium(VI) Reduction by Iron(II). Environ. Sci. Technol. 1998, 32, (14), 2092-2099.

673 674 675

71. Postma, D.; Appelo, C. A. J., Reduction of Mn-oxides by ferrous iron in a flow system: column experiment and reactive transport modeling. Geochim. Cosmochim. Acta 2000, 64, (7), 1237-1247.

676 677 678

72. Borch, T.; Kretzschmar, R.; Kappler, A.; Cappellen, P. V.; Ginder-Vogel, M.; Voegelin, A.; Campbell, K., Biogeochemical Redox Processes and their Impact on Contaminant Dynamics. Environ. Sci. Technol. 2010, 44, (1), 15-23.

679 680 681

73. Jiang, C.; Garg, S.; Waite, T. D., Hydroquinone-mediated redox cycling of iron and concomitant oxidation of hydroquinone in oxic waters under acidic conditions: Comparison with iron-natural organic matter interactions. Environ. Sci. Technol. 2015.

682 683 684 685 686 687 688 689 690

28 ACS Paragon Plus Environment

Page 28 of 34

Page 29 of 34

Environmental Science & Technology

Table 1. Modeled reactions and rate constants of MH2Q oxidation by Fe(III)

691

692 693

a

Unless noted otherwise.

694 695 696 697 698 699 29 ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 34

700

Figure Captions

701

Figure 1. Anaerobic oxidation of 500 nM MH2Q by Fe(III) at pH 5, () [Fe(III)]0 = 100 nM,

702

() [Fe(III)]0 = 250 nM, and () [Fe(III)]0 = 500 nM. Error bars (some of which are too

703

small to be visible) are the standard errors from triplicate measurements, and solid lines

704

represent the model fit using reactions 1 – 5 (Table 1).

705

Figure 2. Generation of Fe(II) from anaerobic oxidation of MH2Q by Fe(III) at different pH,

706

[MH2Q]0 = 500 nM, () [Fe(III)]0 = 250 nM at pH 4, () [Fe(III)]0 = 500 nM at pH 4, ()

707

[Fe(III)]0 = 500 nM at pH 5, and () [Fe(III)]0 = 500 nM at pH 6. Error bars (some of which

708

are too small to be visible) are the standard errors from triplicate measurements, and solid

709

lines represent the model fit using reactions 1 – 5 (Table 1).

710

Figure 3. Aerobic oxidation of MH2Q by Fe(III) at different pH. (A) – (C): [MH2Q]0 = 500

711

nM at (A) pH4, (B) pH5 and (C) pH6 with () [Fe(III)]0 = 100 nM, () [Fe(III)]0 = 250 nM

712

and () [Fe(III)]0 = 500 nM; (D): [Fe(III)]0 = 500 nM at (◇) pH 4 with [MH2Q]0 = 5 µM,

713

() pH 5 with [MH2Q]0 = 5 µM, () pH 6 with [MH2Q]0 = 5 µM, and () pH6 with

714

[MH2Q]0 = 10 µM. Error bars (some of which are too small to be visible) are the standard

715

errors from triplicate measurements, and solid lines represent the model fit using reactions 1

716

– 9 (Table 1).

717

Figure 4. Generation of Fe(II) from aerobic oxidation of MH2Q by Fe(III) at pH 6, [MH2Q]0

718

= 500 nM with () [Fe(III)]0 = 100 nM, () [Fe(III)]0 = 250 nM and () [Fe(III)]0 = 500

719

nM. Error bars (some of which are too small to be visible) are the standard errors from

720

triplicate measurements, and solid lines represent the model fit using reactions 1 – 9 (Table 1).

721

Figure 5. Generation of H2O2 from aerobic oxidation of MH2Q by Fe(III) at different pH,

30 ACS Paragon Plus Environment

Page 31 of 34

Environmental Science & Technology

722

[MH2Q]0 = 500 nM and [Fe(III)]0 = 500 nM at () pH 4, () pH 5 and () pH 6. Error bars

723

(some of which are too small to be visible) are the standard errors from triplicate

724

measurements, and solid lines represent the model fit using reactions 1 – 9 (Table 1).

725

31 ACS Paragon Plus Environment

Environmental Science & Technology

726

Page 32 of 34

Figure 1

250

[M B Q ], n M

200 150 100 50 0 0

5

10

15

20

25

30

T im e , m in

727 728 729

Figure 2

500

[ F e ( II) ] , n M

400 300 200 100 0 0

730

5

10

15

20

T im e , m in

731 732 733 32 ACS Paragon Plus Environment

25

Page 33 of 34

734

Environmental Science & Technology

Figure 3

500

500

(A )

(B )

400

[M B Q ], n M

[M B Q ], n M

400 300 200 100

300 200 100

0

0 0

30

60

90

120

150

180

0

30

60

90

T im e , m in

500

10000

(C )

300 200 100

180

(D )

6000 4000 2000

0

0 0

30

60

90

120

150

180

0

20

T im e , m in

40

60

Figure 4

500

[ F e ( II) ] , n M

400 300 200 100 0 0

80

T im e , m in

735

737

150

8000

[M B Q ], n M

[M B Q ], n M

400

736

120

T im e , m in

30

60

90

120

150

T im e , m in

738 739 33 ACS Paragon Plus Environment

180

100

120

Environmental Science & Technology

740

Page 34 of 34

Figure 5

300

[ H 2 O 2 ], n M

250 200 150 100 50 0 0

10

20

30

40

50

T im e , m in

741 742 743 744 745

Table of Contents Graphic

746

34 ACS Paragon Plus Environment

60