You are on page 1of 75

View Article Online

Environmental
View Journal

Science
Water Research & Technology
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: W. Gan, Y. Ge, Y.
Zhong and X. Yang, Environ. Sci.: Water Res. Technol., 2020, DOI: 10.1039/D0EW00231C.
Volume 4
This is an Accepted Manuscript, which has been through the
Environmental
Number 2
February 2018
Pages 101-338
Royal Society of Chemistry peer review process and has been
Science accepted for publication.
Water Research & Technology
rsc.li/es-water
Accepted Manuscripts are published online shortly after acceptance,
before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 2053-1400 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
Jeyong Yoon et al.
Electrochemical lithium recovery and organic pollutant
shall the Royal Society of Chemistry be held responsible for any errors
removal from industrial wastewater of a battery recycling plant

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/es-water
Page 1 of 74 Environmental Science: Water Research & Technology

View Article Online


DOI: 10.1039/D0EW00231C
Water impact

Environmental Science: Water Research & Technology Accepted Manuscript


Chlorine dioxide (ClO2) is a chemical often used in water treatment for

pre-oxidation or disinfection. This work overviews the reactions of ClO2 with

inorganic and organic compounds in water matrix with focus on their reaction kinetics

and mechanisms. Micropollutant transformation during ClO2 oxidation relevant to

water treatment is also included. The information is essential in improving

understanding the ClO2 oxidation chemistry and byproducts formation in water


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

treatment.
Environmental Science: Water Research & Technology Page 2 of 74

1 The Reactions of Chlorine Dioxide with Inorganic and Organic View Article Online
DOI: 10.1039/D0EW00231C

2 Compounds in Water Treatment: Kinetics and Mechanisms

Environmental Science: Water Research & Technology Accepted Manuscript


3

4 Wenhui Gan†,‡,Yuexian Ge†, Yu Zhong†, Xin Yang†,*

6 † School of Environmental Science and Engineering, Guangdong Provincial Key

7 Laboratory of Environmental Pollution Control and Remediation Technology, Sun

8 Yat-sen University, Guangzhou 510275, China


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

9 ‡ Guangdong Provincial Engineering Research Center for Urban Water Recycling and

10 Environmental Safety, Tsinghua Shenzhen International Graduate School, Tsinghua

11 University, Shenzhen 518055, China

12

13

14 * Corresponding author, Tel: +86-2039332690, email: yangx36@mail.sysu.edu.cn

15

16

17

18

19

20

21

22

23

1
Page 3 of 74 Environmental Science: Water Research & Technology

24 Abstract View Article Online


DOI: 10.1039/D0EW00231C

25 Chlorine dioxide (ClO2), as an alternative to chlorine, has been widely applied in

Environmental Science: Water Research & Technology Accepted Manuscript


26 water treatment. In order to better understand the performance of ClO2 in water

27 treatment, the kinetics and mechanisms of ClO2 with inorganic and organic

28 compounds found in water are critically reviewed. In the case of inorganic

29 compounds, ClO2 reacts with I-, CN-, NO2-, SO32-, Fe(II) and Mn(II) rapidly at an

30 apparent second-order reaction rate constants (kapp) of 102–106 M-1s-1 at pH 7.0 and

31 barely reacts with NH4+ and Br-. In the case of organic compounds, ClO2 selectively
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

32 reacts with compounds with electron-rich moieties, such as phenols (kapp = 103–109

33 M-1s-1), anilines (kapp = 105–108 M-1s-1), and thiols (kapp > 108 M-1s-1). ClO2 also shows

34 high reactivity towards aliphatic tertiary amines and heterocyclic nitrogenous

35 compounds (i.e., indoles and piperidines) with kapp at 101–106 M-1s-1 at pH 7.0, but

36 low reaction reactivity with unsaturated structures (i.e., olefins and aldehydes). The

37 kapp values at pH 7.0 in ClO2 oxidation vary over 14 orders of magnitude. Electron

38 transfer is the dominant pathway for ClO2 reactions. Quantitative structure-activity

39 relationships (QSAR) can be used to predict the specie-specific secondary reaction

40 rate constants for ClO2 oxidation of compounds containing phenolic and amine

41 structures. Little modifications are expected on the structure of the parent compounds

42 upon the primary attack of ClO2, but further oxidation generally leads to the formation

43 of quinones, aldehydes and carboxylic acids. Furthermore, the transformation kinetics

44 of inorganic compounds, typical organic compounds and emerging micropollutants

45 are compared and their half-life times under typical water treatment conditions during

46 ClO2 oxidation are calculated.

47

48

2
Environmental Science: Water Research & Technology Page 4 of 74

49 View Article Online


DOI: 10.1039/D0EW00231C

50 1. Introduction

Environmental Science: Water Research & Technology Accepted Manuscript


51 Chlorine dioxide (ClO2) is a powerful oxidant and an alternative disinfectant to

52 chlorine in drinking water treatment. ClO2 is effective in removing algae (e.g.,

53 Chlamydomonas, Phormidium and Chlorella), inactivating planktonic bacteria and

54 some chlorine-resistant pathogens (e.g., Giardia lamblia and Cryptosporidium

55 parvum), and removing remove iron and manganese from source waters.1-3 ClO2 can

56 be used as a pre-oxidant to control the taste and odor problems. Meanwhile, ClO2 can
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

57 be added at the end of the drinking water treatment process to limit the growth of

58 microorganisms throughout the distribution systems. ClO2 has been reported as being

59 used in 8.1% of drinking water treatment plants in the United States and 32.8% of

60 those in China.4, 5 ClO2 is also widely used in drinking water treatment in some

61 European countries.6 ClO2 is also used in wastewater treatment to inactivate

62 microorganisms and to transform micropollutants and reduce the release of organic

63 halogenated byproducts.7

64 As a chemical oxidant, ClO2 is able to oxidize inorganic pollutants in water (e.g.,

65 iron(II), manganese(II), arsenic(III), etc.). Thus, ClO2 is usually applied to treat source

66 waters suffered from such inorganic contaminants. In addition, ClO2 reacts selectively

67 with organic compounds with electron-rich moieties, such as phenols, tertiary amines

68 and anilines. As a result, ClO2 exhibits promising removal of organic contaminants

69 with such moieties during water and wastewater treatment. Some literatures have

70 reported the kinetics, mechanisms and products from ClO2 oxidation of inorganic and

71 organic compounds.8-13 Hoigné and Bader published a nice summary about the

72 apparent second-order reaction rate constants for ClO2 oxidation of inorganic and

73 organic compounds, including phenols and amines.8 Rav-Acha summarized the


3
Page 5 of 74 Environmental Science: Water Research & Technology

74 reaction pathways of ClO2 oxidation of inorganic compounds and a variety DOI:


group ofView Article Online
10.1039/D0EW00231C

75 organic compounds.12

Environmental Science: Water Research & Technology Accepted Manuscript


76 ClO2 oxidation generally proceed via electron transfer pathway.14, 15 That is,

77 chlorine substitution rarely occurs during ClO2 oxidation. Besides, concerns about

78 bromate or brominated DBPs formation can be spared in ClO2 oxidation because

79 bromide ion is unreactive towards ClO2 under typical water treatment conditions.

80 Therefore, tremendous reduction in the formation of halogenated DBPs, such as

81 trihalomethanes (THMs), haloacetic acids (HAAs), and haloacetonitriles (HANs), is


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

82 one of the major advantages of ClO2 over chlorine.16-18 Moreover, ClO2 exhibits

83 stable oxidation and disinfection strength with insignificant hydrolysis under a wide

84 pH range from pH 2.0 to 10.0, which makes ClO2 more practicable than chlorine in

85 treating waters at high pH.

86 In spite of these benefits, the application of ClO2 in water treatment also brings

87 some disadvantages. One major drawback is the formation of chlorite and chlorate.

88 For the dosed ClO2, 30–70% is transformed into chlorite and up to 10% is

89 transformed into chlorate. Considering the adverse health risks of chlorite and

90 chlorate (i.e., hemolytic anemia and liver damage),19 the level of chlorite and chlorate

91 are restricted in drinking waters. For example, the U.S Environmental Protection

92 Agency (USEPA) has set a maximum contaminant level (MCL) for chlorite of 1

93 mg∙L-1.20 The individual threshold of chlorite and chlorate in drinking waters are 0.7

94 mg∙L-1 in China.21 Consequently, the applied dose of ClO2 in drinking water

95 treatments is typically limited to approximately 1.4 mg∙L-1. In Germany, the dose of

96 ClO2 is even restricted at 0.12–0.16 mg∙L-1.22 In the case of wastewater treatment, the

97 applied dose of ClO2 is usually in the range of 2 to 10 mg∙L-1 depending on

98 wastewater properties to guarantee the microbial inactivation efficiency before reuse

4
Environmental Science: Water Research & Technology Page 6 of 74

99 or discharge.23, 24 However, the release of residual ClO2, chlorite and chlorate toView Article Online
DOI: 10.1039/D0EW00231C

100 environment may has adverse influences on the growth of aquatic organisms (e.g.,

Environmental Science: Water Research & Technology Accepted Manuscript


101 algae and fishes).25, 26 Apart from chlorite and chlorate, the organic transformation

102 products from ClO2 oxidation are also observed in treated water, such as aldehydes,

103 carboxylic acids, esters, ketones, and quinones.27 The bioavailability of these

104 byproducts seemed insignificant as ClO2 treatment leads to minor or no increase of

105 assimilable organic carbon (AOC) of waters.28, 29 For micropollutants present in water

106 (e.g., pharmaceuticals and pesticides), the formation of some toxic transformation
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

107 products is reported.30, 31

108 This study presents an updated overview of ClO2 reaction chemistry in water and

109 wastewater treatment, especially the kinetics and mechanisms of ClO2 reactions with

110 inorganic and organic compounds. The literature reports are summarized primarily

111 based on the functional groups of the compounds, i.e., inorganic compounds (e.g., I-,

112 CN-, NO2-, SO32-, Fe(II), Mn(II) and As(III)) and organic compounds (e.g., olefins,

113 amines, phenols, anilines, thiols, amino acids, etc.). Additionally, the transformation

114 of micropollutants during ClO2 oxidation is summarized. Quantitative

115 structure-reactivity relationships (QSARs) are derived for compounds bearing amines

116 and phenols and the ClO2 reactivity with some emerging contaminants are predicted

117 accordingly. Additionally, the transformation of micropollutants during ClO2

118 oxidation is summarized. Finally, the removal of inorganic and organic compounds

119 under water treatment conditions are discussed to provide a picture of their

120 transformation in practical ClO2 treatment.

121

122 2. Physical and chemical properties of ClO2

5
Page 7 of 74 Environmental Science: Water Research & Technology

123 ClO2 is highly soluble in waters (about 3 g∙L-1 at 25°C) and remains as dissolved View Article Online
DOI: 10.1039/D0EW00231C

124 gas in solutions. Thus, hydrolysis of ClO2 is insignificant in wide pH range of 2.0–

Environmental Science: Water Research & Technology Accepted Manuscript


125 10.0. However, ClO2 can undergo disproportionation in alkaline solution to form

126 chlorite and chlorate (Eq. 1).32

127 2𝐶𝑙𝑂2 +2𝑂𝐻 ― = 𝐶𝑙𝑂2― + 𝐶𝑙𝑂3― + 𝐻2𝑂 (1)

128 Due to the instability of ClO2 gas when compressed, ClO2 is generally generated

129 on site. ClO2 is generated via mixing chlorite or chlorate salts with strong acids (e.g.,

130 H2SO4, HCl) or oxidants (e.g., Cl2, HOCl, H2O2).33 The generated ClO2 solution
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

131 inevitably contains unreacted raw reactants (e.g., chlorite or chlorate) or some

132 side-reaction products (e,g., Cl2) as impurities.

133 The molecular structure of ClO2 is shown in Fig. 1. The Cl-O bond length is 1.47

134 Å and the O-Cl-O angle is 117.6°. ClO2 has an absorption peak at 359 nm, with a

135 molar absorptivity of 1250 M-1cm-1. This extinction coefficient is independent of

136 temperature, pH, and ionic strength and is often used for determining ClO2

137 concentrations.

138

139 Fig. 1 The molecular structure of ClO2

140 ClO2 has an odd number of valance electrons and it is a stable radical. ClO2

141 usually reacts with organic compounds via electron transfer pathway. The organic

142 compounds serve as an electron acceptor and ClO2 per se is reduced to the chlorite

143 anion (Eq. 2). By comparison, chlorination of organic compounds usually initiates

144 with an electrophilic attack of HOCl at nucleophilic sites, which leads to formation of

145 chlorinated byproducts.34, 35 Ozonation generally proceeds with ozone addition


6
Environmental Science: Water Research & Technology Page 8 of 74

146 reactions to unsaturated bonds or other electron donating moieties, resulting inView Article Online
DOI: 10.1039/D0EW00231C

147 oxygen-substitution or ring cleavage.36, 37 Therefore, ClO2 oxidation is

Environmental Science: Water Research & Technology Accepted Manuscript


148 mechanistically different from chlorination or ozonation processes.

149 𝐶𝑙𝑂2 + 𝑒 ― = 𝐶𝑙𝑂2― (2)

150 The redox potential for the ClO2/ClO2- couple is pH-dependent38 because of the

151 photolytic equilibrium between chlorite and chlorous acid (pKa = 1.72). In drinking

152 water treatment (pH > 4.0), the E0red(ClO2/ClO2-) is at 0.95 V.39 Chlorite generated

153 from ClO2 can be further reduced to chloride ions (E0red(ClO2-/Cl-) =1.58V) by some
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

154 reductive compounds (Eq. 3).40 Thus the overall reaction will be that 1 mol ClO2

155 accepts 5 mol electrons and is reduced to chloride ions.

156 𝐶𝑙𝑂2― +4𝐻 + +4𝑒 ― = 𝐶𝑙 ― +2𝐻2𝑂

157 (3)

158

159 3. ClO2 oxidation of inorganic and organic compounds

160 The generic chlorine dioxide oxidation reaction is shown in Eq. 4:

161 𝐶𝑙𝑂2 +𝐴→𝑝𝑟𝑜𝑑𝑢𝑐𝑡 (4)

162 where A is an inorganic or organic compound.

163 Previous studies have determined that the kinetics of such reactions can be well

164 described by the second-order kinetics model, as shown in Eq. 5:


𝑑[𝐴]𝑡𝑜𝑡
165 𝑑𝑡 = ― 𝑘𝑎𝑝𝑝[𝐴]𝑡𝑜𝑡[𝐶𝑙𝑂2] (5)

166 where kapp is the apparent second-order reaction rate constant and [A]tot refers to the

167 total concentrations of A species.

168 ClO2 is a neutral compound and does not hydrolyze in water. When A can

169 dissociate according to an acid-base equilibrium controlled by pH, kapp, which

7
Page 9 of 74 Environmental Science: Water Research & Technology

170 characterizes the sum of the reactions of the two species HA and A-, is DOI:
a linear View Article Online
10.1039/D0EW00231C

171 function of the composition (Eq. 6).

Environmental Science: Water Research & Technology Accepted Manuscript


172 𝑘𝑎𝑝𝑝 = (1 ― 𝛼)𝑘𝐻𝐴 +𝛼 ∙ 𝑘𝐴 ― = 𝑘𝐻𝐴 + 𝛼(𝑘𝐴 ― ― 𝑘𝐻𝐴) . (6)

173 kHA and kA - here represent the specific second-order reaction rate constants for the

174 reactions of ClO2 with HA and A-, respectively. α is the fraction of the compound

175 occurring in the deprotonated form (A- ) (Eq. 7).

[𝐴 ― ] 1
176 𝛼 = [𝐴 ― ] + [𝐻𝐴] = (𝑝𝐾𝑎 ― 𝑝𝐻) (7)
1 + 10

177 By using Eq. 5, lots of kinetics was obtained for the reactions between ClO2 and
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

178 diverse groups of inorganic and organic compounds. For the kinetic measurements,

179 stopped-flow method is usually used and the concentrations of precursors are set to be

180 5–10 times higher than ClO2 to achieve the pseudo-first-order reaction conditions.

181 The second-order reaction rates of ClO2 oxidation presented in Section 3.1, 3.2 and

182 Section 4 are mainly obtained via this method.

183 Notably, recent studies confirmed and quantified the formation of HOCl during

184 ClO2 reactions with organic compounds and NOM.15, 41, 42 Trace amount of HOCl can

185 greatly enhance the transformation efficiency of some pollutants, such as atenolol,

186 metoprolol and ciprofloxacin.41 However, most of previous studies were conducted

187 using ClO2 solutions without adding HOCl scavenger (e.g., glycine). Therefore, bias

188 may present in the reported apparent or specie-specific reaction rates for ClO2

189 oxidation because of the interferences from in situ formed HOCl during reaction,

190 especially for compounds (e.g., primary and secondary amines) with higher reactivity

191 towards HOCl than ClO2.

192

193 3.1. Oxidation of inorganic compounds

8
Environmental Science: Water Research & Technology Page 10 of 74

194 ClO2 can oxidize many of the inorganic compounds found in naturalDOI:
waters. View Article Online
10.1039/D0EW00231C

195 ClO2 is firstly reduced to chlorite via the single electron transfer, but chlorite might

Environmental Science: Water Research & Technology Accepted Manuscript


196 not be the final reduction product. Further reduction of chlorite can occur as shown in

197 Eq. 3 with presence of some reductive inorganic compounds (e.g., iodide and

198 cyanide), despite of the much slower redox reaction in slightly acidic to neutral

199 circumstances. The speciation of the oxidation products is governed by reaction pH

200 and the molar ratio of ClO2 and the reactants. Table 1 lists the species-specific and

201 apparent second-order reaction rate constants of ClO2 reactions with inorganic
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

202 compounds commonly found in natural waters.

203

9
Page 11 of 74 Environmental Science: Water Research & Technology

204 Table 1 Reaction rate constants of ClO2 with inorganic compounds View Article Online
DOI: 10.1039/D0EW00231C

Species-specific Apparent

Environmental Science: Water Research & Technology Accepted Manuscript


second-order reaction second-order

Compounds rate constants (M-1s-1) reaction rate Refer


pKa T (oC)
(HA/A-) constants at pH ences

kHA kA- 7.0,

kapp (M-1s-1)

Iodide (I-) 1.4×103 1.4×103 22–24 8

Bromide (Br-) < 10-4 < 0.05 22–24 8


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Cyanide (HCN/CN-) 9.4 < 10-5 8.0×103 4.0×102 22–24 8

Nitrite (HNO2/NO2-) 3.3 1.1×102 1.1×102 22–24 8

Sulfite (HSO3-/SO32-) 7.2 2.7×106 2.7×106 25 42

Ammonia
9.3 <10-6 <10-6 <10-6 22–24 8

(NH4 +/NH
3)

Iron (II) (Fe2+/FeOH+) 3×103 3.0×103 22–24 8

Manganese (II)
10.6 < 0.1 ~ 2.0×107 5.0×103 22–24 8

(Mn2+/MnOH+)

205

206 3.1.1. I-, CN-, NO2-, SO32- and HS-

207 During oxidation of aqueous iodide, ClO2 can rapidly oxidize I- to I2 (Eq. 8).38

208 Chlorite produced by the reduction of ClO2 can also react with excess I- to form I2 at

209 pH 4–8 (Eq. 9).44-46 ClO2 has also been reported to oxidize I- to iodate (IO3-), but only

210 at very low pH (1–3.5), which is not representative of the pH of source waters.46 No

211 substantial IO3- formation is observed under practical ClO2 oxidation treatment

212 conditions.47 It should be noted that ClO2 does not react with chloride and barely react

213 with bromide (< 10-4 M-1s-1).8 Since HOCl is formed during ClO2 oxidation,

214 chloramine and hypobromous acid (HOBr) is possible to form. However, only trace

10
Environmental Science: Water Research & Technology Page 12 of 74

215 level of total organic bromine TOBr (< 10 μg∙L-1) was formed when 1.5 mg∙L -1 ClOView Article Online
2
DOI: 10.1039/D0EW00231C

216 reacted with 3 mg∙L-1 Suwannee River NOM solution.48 Therefore, the side-reactions

Environmental Science: Water Research & Technology Accepted Manuscript


217 between in situ formed HOCl and inorganic ions are insignificant.

218 2𝐶𝑙𝑂2 +2𝐼 ― = 2𝐶𝑙𝑂2― + 𝐼2 (8)

219 𝐶𝑙𝑂2― +4𝐻 + +4𝐼 ― = 2𝐼2 + 𝐶𝑙 ― +2𝐻2𝑂 (9)

220 Subsequently, the disproportionation of I2 leads to the formation of HOI (Eq.

221 10)49, 50 and the reaction between I2 and excessive I- results in the formation of I3- (Eq.

222 11).51
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

223 𝐼2 + 𝐻2𝑂⇌𝐻𝑂𝐼 + 𝐼 ― + 𝐻 + (10)

224 𝐼2 + 𝐼 ― ⇌𝐼3― (11)

225 It has been reported that I2, HOI and I3- can react with NOM in water to form

226 iodinated DBPs including iodoform (CHI3) and iodoacetic acid (IAA).44, 51 Thus,

227 formation of iodinated DBPs was observed during ClO2 oxidation of

228 iodine-containing waters.52, 53

229 Similar to iodide, the oxidation of cyanide (CN-) and nitrite (NO2-) by ClO2

230 mainly involves electron transfer reactions which generate cyanate (CNO-) and nitrate

231 (NO3-), respectively (Eq. 12–13).43, 54-56



232 2𝐶𝑙𝑂2 + 𝐶𝑁 ― +2𝑂𝐻 ― = 2𝐶𝑙𝑂2― + 𝐶𝑁𝑂 + 𝐻2𝑂 (12)

233 2𝐶𝑙𝑂2 + 𝑁𝑂2― + 𝐻2𝑂 = 2𝐶𝑙𝑂2― + 𝑁𝑂3― +2𝐻 + (13)

234 Chlorite is primarily produced from these reactions, which may further reduced

235 to chloride via reactions with CN- and NO2- (Eq. 14–15).

236 𝐶𝑙𝑂2― +2𝐶𝑁 ― = 𝐶𝑙 ― +2𝐶𝑁𝑂 ― (14)

237 𝐶𝑙𝑂2― +2𝑁𝑂2― = 𝐶𝑙 ― +2𝑁𝑂3― (15)

238 The mechanism of sulfite oxidation by ClO2 is more complicated due to the

239 pH-dependent distribution of aqueous sulfite (S(IV)) species


11
Page 13 of 74 Environmental Science: Water Research & Technology

240 ([S(IV)]=[H2SO3]+[HSO3-]+[SO32-]).39, 57-59 In addition, the molar ratio of S(IV) andView Article Online
DOI: 10.1039/D0EW00231C

241 ClO2 also influences the speciation of products. The reaction stoichiometry and

Environmental Science: Water Research & Technology Accepted Manuscript


242 products of ClO2 and sulfite are summarized in Scheme 1.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

243

244 Scheme 1 Summary of reaction stoichiometry and products of ClO2 oxidation of

245 sulfur-containing inorganic compounds (a) SO32-; (b) HS-

246 As illustrated in Scheme 1a, in a slightly acidic solution, SO32- is oxidized to

247 sulfate and both chloride and chlorate are produced from ClO2 reduction at ClO2 to

248 SO32- molar ratio of 1:1 (Eq. 16).60 When ClO2 to SO32- molar ratio is 1:3, dithionate,

249 sulfate and chloride are generated (Eq. 17).60

250 2𝐶𝑙𝑂2 +2𝑆𝑂23 ― + 𝐻2𝑂 = 2𝑆𝑂24 ― + 𝐶𝑙 ― + 𝐶𝑙𝑂3― +2𝐻 + (16)

251 2𝐶𝑙𝑂2 +6𝑆𝑂23 ― = 𝑆2𝑂26 ― +4𝑆𝑂24 ― +2𝐶𝑙 ― (17)

252 In alkaline solutions, SO32- is oxidized to sulfate, but the species of ClO2

253 reduction products varied with the ratio of ([S(IV)]0/[ClO2]). Three competitive

254 stoichiometries have been found, with the ratio of ([S(IV)]0/[ClO2]) varying between

255 1.5 and 2.0 (Eq. 18–20):61


12
Environmental Science: Water Research & Technology Page 14 of 74

256 2𝐶𝑙𝑂2 + 𝑆𝑂23 ― +2𝑂𝐻 ― = 𝑆𝑂24 ― +2𝐶𝑙𝑂2― + 𝐻2𝑂 (18)View Article Online
DOI: 10.1039/D0EW00231C

257 2𝐶𝑙𝑂2 +2𝑆𝑂23 ― +2𝑂𝐻 ― = 𝐶𝑙 ― + 𝐶𝑙𝑂3― +2𝑆𝑂24 ― + 𝐻2𝑂 (19)

Environmental Science: Water Research & Technology Accepted Manuscript


258 2𝐶𝑙𝑂2 +3𝑆𝑂23 ― +2𝑂𝐻 ― = 𝐶𝑙𝑂2― + 𝐶𝑙 ― +3𝑆𝑂24 ― + 𝐻2𝑂 (20)

259 ClO2 also reacts with hydrosulfide (HS-) quickly.32 The species distribution of

260 the products depends on the pH and the molar ratios between ClO2 and reactants. As

261 illustrated in Scheme 1b, in alkaline solutions, several reactions might contribute to

262 the overall reaction stoichiometry (Eq. 21–24).62

263 8𝐶𝑙𝑂2 + 𝐻𝑆 ― +9𝑂𝐻 ― = 𝑆𝑂24 ― +8𝐶𝑙𝑂2― +5𝐻2𝑂 (21)


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

264 4𝐶𝑙𝑂2 + 𝐻𝑆 ― +5𝑂𝐻 ― = 𝑆𝑂24 ― +2𝐶𝑙𝑂3― +2𝐶𝑙 ― +3𝐻2𝑂 (22)

265 2𝐶𝑙𝑂2 + 𝐻𝑆 ― + 𝑂𝐻 ― = 𝑆 + 2𝐶𝑙𝑂2― + 𝐻2𝑂 (23)

266 8𝐶𝑙𝑂2 +5𝐻𝑆 ― +13𝑂𝐻 ― = 5𝑆𝑂24 ― +8𝐶𝑙 ― +9𝐻2𝑂 (24)

267 Therefore, ClO2 can rapidly remove the majority of anionic inorganic

268 contaminants under water treatment conditions. For example, hydrosulfide is

269 commonly present in municipal wastewaters, produced from biological reduction of

270 sulfates and the decomposition of sulfur-containing organic moieties. Post-treatment

271 of ClO2 can rapidly destroy hydrosulfide to abate its unpleasant smell in wastewater

272 discharges. Moreover, contamination of CN- may occur when source waters impacted

273 by discharges from electroplating industries. The level of CN- in drinking waters is

274 limited to 0.2 mg∙L-1 by USEPA due to its severe toxicity. Pretreatment with ClO2 can

275 detoxify contaminated source waters by converting CN- to much less toxic CNO-.

276 3.1.2. Fe(II), Mn(II), and As(III)

277 Soluble iron and manganese generally exist in their divalent forms, which

278 impedes drinking water treatment.63 Fe(II) and Mn(II) can be oxidized to insoluble

279 Fe(III) and Mn(III or IV) species, respectively and then removed by coagulation

280 and/or filtration in water treatment. ClO2 is often used as primary oxidant to reduce
13
Page 15 of 74 Environmental Science: Water Research & Technology

281 Fe(II) and Mn(II) level in source waters in this way,, especially in small
DOI: water
View Article Online
10.1039/D0EW00231C

282 treatment plants or in circumstances like a shock increase in soluble iron or

Environmental Science: Water Research & Technology Accepted Manuscript


283 manganese in the summer.64 ClO2 oxidizes Fe(II) and Mn(II) via rapid one-electron

284 transfer leading to the formation of chlorite.64-66 In addition, chlorite (E0red(ClO2-/Cl-)

285 =1.58V) can further react with Fe(II) (E0red(Fe3+/Fe2+) = 0.77V) and Mn(II)

286 (E0red(Mn3+/Mn2+) =1.50V).40 The overall reactions under water treatment conditions

287 can be summarized as in Eq. 25–26.32, 65-67

288 𝐶𝑙𝑂2 +5𝐹𝑒2 + +4𝐻 + = 5𝐹𝑒3 + + 𝐶𝑙 ― +2𝐻2𝑂


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

289 (25)

290 2𝐶𝑙𝑂2 +5𝑀𝑛2 + +6𝐻2𝑂 = 5𝑀𝑛𝑂2(𝑠) + 12𝐻 + +2𝐶𝑙 ― (26)

291 The reaction rate constants of ClO2 oxidation of Fe(II) and Mn(II) increase

292 greatly with increasing pH.65 The reaction rate constants of Mn(II) oxidation increases

293 from 3.2×103 M-1s-1 to 5.0×104 M-1s-1 when pH increases from 6 to 8.40 Results from

294 bench and pilot studies shows that ClO2 pre-treatment incorporated with

295 post-filtration can achieve promising removal of Fe(II) and Mn(II) (> 95%).68-70

296 It should be noted that adding dissolved ferrous sulfate (FeSO4) is an approach to

297 eliminate excess chlorite from waters to meet the regulatory standards (Eq. 27).71

298 Fe(II) rapidly reduced chlorite to chloride (kapp = 1.9×103 M-1s-1 at pH 7.0)72 and the

299 formed Fe(III) colloids can then be removed during subsequent coagulation or

300 filtration.

301 4𝐹𝑒2 + + 𝐶𝑙𝑂2― +10𝐻2𝑂⇌4𝐹𝑒(𝑂𝐻)3(𝑠) + 𝐶𝑙 ― +8𝐻 + (27)

302 In natural waters, soluble arsenic mainly exists as a combination of As(III)

303 [AsO33-] and As(V) [AsO43-].73 ClO2 can oxidize As(III) to As(V) (Eq. 28–29),

304 leading to promoted removal of arsenic in post-filtration or adsorption treatment

305 because removal of As(V) from waters is more accessible than As(III).74-76

14
Environmental Science: Water Research & Technology Page 16 of 74

306 2𝐶𝑙𝑂2 + 𝐴𝑠𝑂33 ― + 𝐻2𝑂 = 𝐴𝑠𝑂34 ― +2𝐶𝑙𝑂2― +2𝐻 + (28)View Article Online
DOI: 10.1039/D0EW00231C

307 2𝐶𝑙𝑂2 +5𝐴𝑠𝑂33 ― + 𝐻2𝑂 = 5𝐴𝑠𝑂34 ― +2𝐶𝑙 ― +2𝐻 + (29)

Environmental Science: Water Research & Technology Accepted Manuscript


308 Effective oxidation of As(III) to less toxic As(V) can be achieved at a pH range of 5

309 to 9 and at a ClO2 to As(III) molar ratio of 1:1. Acidic conditions and higher dose of

310 ClO2 can promote As(III) removal efficiency. 77 ClO2 removes As(III) less effectively

311 than free chlorine or permanganate in demineralized water.75, 77, 78

312

313 3.2. Oxidation of organic compounds


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

314 The molecular structure of ClO2 has a central chlorine atom with a Cl=O double

315 bond and a three-electron bond, where one electron tends to be associated with the

316 chlorine atom, one primarily with the oxygen atom, and one in the region between

317 them (i.e., in the bond). ClO2 therefore tends to react with organic compounds as a

318 one-electron accepter and is reduced to chlorite. That makes ClO2 a selective oxidant

319 whose reactivity generally favors organic molecules with a lone pair of electrons or π

320 electrons (e.g., tertiary amines, phenols and aniline). The DBPs from ClO2 oxidation

321 detected in water treatment are mainly chlorite, chlorate, oxidized by-products (acids,

322 aldehydes and alcohols) and some chlorinated by-products.79, 80 Tracking the

323 precursors and the formation pathways of these oxidized by-products can help to

324 optimize the use of ClO2, avoiding to a great extent the formation of toxic

325 by-products.

326 The formation of HOCl is known to correlate with the total organic chloride

327 (TOCl) formation during ClO2 oxidation of phenols.15 That raises the question on

328 what reactions causing HOCl formation. The organic molecule is generally converted

329 to a free radical after the initial electron transfer to a ClO2 molecule. It can then

330 combine with the oxygen atom of another ClO2 molecule causing the release of

15
Page 17 of 74 Environmental Science: Water Research & Technology

331 HOCl.81 Exploring the reactions of ClO2 with organics from the perspective
DOI:of theView Article Online
10.1039/D0EW00231C

332 release of HOCl can provide new insight into DBP control in ClO2 treatment. This

Environmental Science: Water Research & Technology Accepted Manuscript


333 section will therefore emphasize the origins of typical oxidized organic by-products

334 and the formation of HOCl during ClO2 oxidation in actual treatment situations.

335

336 3.2.1. Reaction with unsaturated olefins

337 ClO2 reactions with olefins were generally slow or negligible with rates in the

338 range of 10-3 to 1.5×102 M-1s-1(Table 2). Unsaturated carboxylic acids (e.g., cinnamic,
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

339 maleic and fumaric acids) are reported to be inert towards ClO2. There is dispute

340 about the extent to which ClO2’s reactions are initiated by hydrogen abstraction or

341 electron transfer. Rav-Acha and his co-workers found no isotope effect on olefins’

342 reactions in H2O and D2O, which argues against hydrogen abstraction. In view of the

343 strong solvent effect and the linear relationship between reaction rate and olefins’

344 ionization potentials, the electron transfer mechanism seems more plausible (Scheme

345 2). ClO2 attacks at the double bonds to form cation radicals, which combine rapidly

346 with another ClO2 molecule (which has unpaired electrons) to form esters. The esters

347 thus formed are unstable and convert to ketones, alcohols or acids through HOCl

348 elimination. During ClO2 oxidation of cyclohexene ( ), cyclohex-1-ene-3-one

349 ( ) and chlorocyclohen-1-ene ( ) were found to be the main products,

350 companied by other chlorinated cyclohexan-1-ol derivatives ( ).82

351 Phenyl-2-chloroethanone ( ) and phenyl-2-chloroethanol ( ) are

352 found to be main products from ClO2 oxidation of styrene ( ).83 ClO2 oxidation

16
Environmental Science: Water Research & Technology Page 18 of 74

View Article Online


DOI: 10.1039/D0EW00231C
353 converts about half of indene ( ) to chlorohydrin ( ),84 suggesting the

Environmental Science: Water Research & Technology Accepted Manuscript


354 significance of HOCl yields during the reaction. Relatively low chlorite formation

355 relative to the ClO2 consumed also supports the substantial formation of HOCl during

356 olefin oxidation. Only a quarter of the ClO2 consumed is reduced to chlorite when the

357 molar ratio of ClO2 to indene is 2:1 at pH 7.0.84 ClO2 oxidation of cyclohexene and

358 isoprene also has relative low chlorite conversion (< 20%).42

359 In general, fractions of olefins in NOM are minor.85 Olefins detected in surface

360 waters and domestic sewages are mainly in forms of branched or cyclic olefins (e.g.,
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

361 polycyclic aromatic hydrocarbon, PAH) derived from pollutions of petroleum

362 industries. However, oxidation of olefins in water treatments are unlikely to be

363 significant because due to the low concentrations of these olefinic pollutants in source

364 waters (<100 ng∙L-1) and sewages (< 500 μg∙L-1).86, 87 In addition, ClO2 reactions with

365 olefins and PAHs generally take hours to accomplish,88 which may likely occurs in

366 distribution systems with adequate ClO2 residual.

367 Table 2 Kinetics of ClO2 oxidation of selected compounds with unsaturated bonds,

368 oxygenated or sulfur-containing moieties

Apparent second-order reaction rate


Compounds T (oC) Reference
constants at pH 7.0, kapp (M-1s-1)

Olefins

Indene 1.1×102 84

(E)-Stilbene 1.3×102 84

Acenaphthylene 6.2 84

α-Methylstyrene 3.1 84

trans-α-Methylstyrene 1.2 25 89

β-Methylstyrene 3.1×101 84

17
Page 19 of 74 Environmental Science: Water Research & Technology

ρ-Methylstyrene 1.6×10-1 25 89 View Article Online


DOI: 10.1039/D0EW00231C

ρ-Phenylstyrene 1.0×10-1 25 89

Environmental Science: Water Research & Technology Accepted Manuscript


Styrene 1.6 84

ρ-Chlorostyrene 4.1×10-1 25 89

Cyclohexene 7.2×10-1 84

Allybenzene 2.4×10-1 84

1-Hexene-3-ol < 10-2 24 8

2,4-Hexadiene-1-ol 8.6 24 8

Ketones
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

2-Pentanone < 3.0×10-5 24 8

Acetone < 10-5 24 8

5,5-Dimethyl-1,3-cyclohexanedione
- 24 8

(dimedone)

1,4-Benzoquinone < 5.0×10-3 90

Aldehydes

Benzaldehyde < 3.0×10-4 23 91

Butyryl aldehyde < 1.0×10-2 92

Other oxygenated compounds

2-Ethanolfuran 3.7×10-1 88

2,5-Dimethylfuran 1.2×102 8

Furfuryl alcohol 3.7×10-1 24 8

Ascorbate 1.0×107 22 13

Anisole < 2.0×10-3 24 8

Phthalic acid < 10-3 24 8

Sulfur-containing compounds

2-Methyl-2-propanethiol 7.5×108 24 8

369

18
Environmental Science: Water Research & Technology Page 20 of 74

View Article Online


DOI: 10.1039/D0EW00231C

Environmental Science: Water Research & Technology Accepted Manuscript


370

371 Note: modified from reference84

372 Scheme 2 Reaction pathway proposed for the ClO2 oxidation of olefins

373

374 3.2.2. Reaction with oxygenated moieties


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

375 3.2.2.1. Aliphatic alcohols, aldehydes, ketones and organic acids

376 ClO2 has limited reactivity towards aliphatic alcohols, aldehydes, ketones and

377 organic acids and the presence of activating electron-donor groups could enhance the

378 reactivity. In practical water treatment conditions, ClO2 is unlikely to react with the

379 alcohols, aldehydes, ketones and organic acids commonly found.

380 Alcohols are generally unlikely to convert to ketones or carboxylic acids at

381 neutral pH and moderate temperature because the relevant reaction rates are only

382 10-4–10-5 M-1s-1. The reaction rate of ClO2 oxidation of furfuryl alcohol is about 0.4

383 M-1s-1 at pH 8.0 and 23oC,8 and that is higher than the rates of other alcohols. The

384 reaction mechanism is thought to start with hydrogen abstraction as illustrated in

385 Scheme 3.

386

387 Scheme 3 Reaction pathway proposed for ClO2 oxidation of alcohols

388 The carbonyl group in aldehydes and ketones is polarized with the carbon atom

389 bearing a partial positive charge. Thus, the reactivity of aldehydes and ketones

390 towards ClO2 is limited because those groups usually favor nucleophilic attack on the
19
Page 21 of 74 Environmental Science: Water Research & Technology

391 C=O bond. Hexanal and 2-methybutanal are reported not to react with DOI:
ClO 2 at
View Article Online
10.1039/D0EW00231C

392 moderate temperatures.93 Some aldehydes, though, can be oxidized to carboxylic

Environmental Science: Water Research & Technology Accepted Manuscript


393 acids by ClO2 in neutral aqueous conditions. Oxidation of butyraldehyde to butyric

394 acid is one example, but the reaction rate is only 10-2 M-1s-1 or less.92 In addition,

395 chlorite can also oxidize aldehydes to carboxylic acids at slightly acidic or neutral

396 pH.94, 95 However, the accumulation of 10–50 μg∙L-1 of aldehydes is observed in

397 ClO2-treated waters.27 This suggests that the oxidation of aldehydes by ClO2 in water

398 treatment is so slow as to be insignificant.


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

399 Ketones show higher reactivity towards ClO2 than aldehydes because of the

400 stronger electron-donating effect of the R groups attached to both sides towards the

401 positive carbonyl carbon. However, low molecular weight ketones like acetone and

402 2-pentanone are generally unreactive towards ClO2, with rates lower than 10-5 M-1s-1.

403 Although the second-order reaction rate between ClO2 and protonated dimedone (pKa

404 =5.2) is about 2×104 M-1s-1, deprotonated dimedone is found to be inert with ClO2

405 oxidation.8 ClO2 can oxidize benzoquinone, but at the relatively slow rate of 5×10-3

406 M-1s-1 at pH 7.0.8 Degradation of benzoquinone to cyclopent-4-ene-1,3-dione

407 derivatives and chlorinated acetones after dealkylation occurs during ClO2

408 oxidation.96

409 Carboxylic acids are usually unreactive towards ClO2 in general and are often

410 found in ClO2-treated water. Some carboxylic acids containing specific reactive

411 groups (e.g., salicylic acids and glyoxylic acid) are exceptions.

412

413 3.2.2.2. Phenolic compounds

414 Humic substances usually have a large fraction of polyphenolics97 which serve

415 as major electron-donating groups in the oxidation of humic substances.98 That

20
Environmental Science: Water Research & Technology Page 22 of 74

416 accelerates the consumption of the oxidant and results in the formation ofView Article Online
DOI: 10.1039/D0EW00231C

417 considerable amounts of DBPs.

Environmental Science: Water Research & Technology Accepted Manuscript


418 Phenols react with ClO2 rapidly with rates of 103 –108 M-1s-1 at neutral pH

419 (Table 3). The reaction rate constants of dissociated phenols with ClO2 are generally

420 about six orders of magnitude higher than those of the undissociated phenols.99 High

421 pH therefore favors ClO2 oxidation of phenols. The substituents on phenols greatly

422 affect their ClO2 oxidation rates. Lee and von Gunten have developed QSARs

423 defining this and negative linear relationships between the Hammett constants
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

424 (∑σ-o,m,p and ∑σ+o,m,p) of the substituents on phenols and the second-order reaction

425 rate constants with ClO2 of dissociated and undissociated substituted phenols. Eq. 30–

426 31 and Fig. 2 show their results.99 The reaction rate constants decrease with increasing

427 Hammett constants, which might be attributed to higher bond dissociation energies

428 when electron-withdrawing substituents are present.100 Based on the established

429 QSAR relationship, the reactivity of emerging micropollutants and other synthetic

430 organic compounds with ClO2 can be predicted. The second-order reaction rates

431 between ClO2 and phenolic micropollutant 17α-ethinylestradiol, bisphenol A,

432 acetaminophen and triclosan can be well predicted by Eq. 31 within a prediction error

433 range of 2/3–1.5 (i.e., the ratio between the differences between prediction rates and

434 measured rates).


+
435 log (𝑘𝑃ℎ𝑂𝐻) = 0.41( ± 0.50) ― 4.69( ± 1.13) ∗ ∑𝜎𝑜,𝑚,𝑝

436 (30)

437 log (𝑘𝑃ℎ𝑂 ― ) = 8.03( ± 0.13) ― 3.24( ± 0.28) ∗ ∑𝜎𝑜,𝑚,𝑝

438 (31)

21
Page 23 of 74 Environmental Science: Water Research & Technology
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


439 Table 3 Kinetics of ClO2 oxidation of phenolic compounds

Species-specific
Apparent second-order rate
Hammett constants second-order rate
Compounds pKa constants at pH 7.0, T (oC) References
constants(M-1s-1)
kapp (M-1s-1)
σ σ+ σ- HA A-
Phenol 0 0 0 10.0 4.0×10-1 4.9×107 4.9×104 24 8

2-Methylphenol -0.13 -0.20 -0.17 10.3 1.6×101 4.4×108 2.2×105 24 8

3-Methylphenol -0.07 -0.07 -0.06 10.1 < 4.0 1.0×108 7.9×104 24 8

4-Methylphenol -0.17 -0.31 -0.17 10.3 4.0×101 5.2×108 2.6×105 24 8

4-Ethylphenol -0.15 -0.3 -0.19 10.3 5.0×101 8.0×108 4.0×105 24 8

4-tert-Butylphenol -0.2 -0.26 -0.13 10.2 1.1×101 1.5×108 9.5×104 24 8

2-Carboxylphenol -0.69 -0.01 0.31 13.6 4.6×101 6.0×107 6.1×101 24 8

4-Carboxylphenol 0 -0.02 0.31 9.2 4.0×101 4.8×107 3.0×105 24 8

4-Acetylphenol 0.5 0.5 0.84 8.1 < 0.5 8.0×105 5.9×104 24 8

4-Cyanophenol 0.66 0.66 1 7.9 < 4.0×103 4.9×102 21 101

2-Chlorophenol 0.4 0.07 0.19 8.5 1.5 3.5×107 1.1×106 24 8

4-Chlorophenol 0.23 0.11 0.19 9.4 <2 3.5×107 1.4×105 24 8

4-Bromophenol 0.23 0.15 0.25 2.7×107 2.7×107 24 8

2-Hydroxyphenol -0.2 -0.60 -0.37 9.1 < 5.0×103 2.0×109 1.6×107 24 8

3-Hydroxyphenol 0.12 -0.04 0.12 9.2 4.0×101 4.8×107 3.0×105 24 8

4-Hydroxyphenol -0.37 -0.92 -0.37 3.9×104 9.0×108 9.0×108 25 43, 81

22
Environmental Science: Water Research & Technology Page 24 of 74
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


440 2-Methoxyphenol -0.27 -0.51 -0.37 9.9 1.0×103 1.2×109 1.5×106 24 8

3-Methoxyphenol 0.12 0.05 0.13 4.9×107 4.9×107 21 101

4-Methoxyphenol -0.27 -0.78 -0.26 10.2 2.50×104 1.7×109 1.1×106 24 8

4-Nitrophenol 0.78 0.79 1.27 7.2 1.40×10-1 4.0×103 1.6×103 24 8

4-Sulfonatophenol 0.35 0.35 0.58 8.8 <1 1.0×106 1.6×104 24 8

2,4-Dimethylphenol -0.3 -0.51 -0.34 10.6 9.0×102 2.1×109 5.3×105 24 8

2,5-Dimethylphenol -0.2 -0.27 -0.23 10.2 1.0×102 6.8×108 4.3×105 24 8

2-Methoxy-4-formylphenol 0.15 0.22 0.66 7.7 - 1.8×108 3.0×107 24 8

2,4-Dichlorophenol 0.63 0.18 0.38 7.8 < 0.6 2.2×107 3.0×106 24 8

2,4,6-Trimethylphenol -0.43 -0.72 -0.51 10.9 3.9×103 4.0×109 5.1×105 24 8

2,4,6-Trichlorophenol 1.03 0.26 0.57 6.2 < 8.0×102 5.3×106 4.6×106 24 8

Pentachlorophenol 1.77 1.06 1.31 4.7 1.4×103 1.4×103 24 8

23
Page 25 of 74 Environmental Science: Water Research & Technology

View Article Online


DOI: 10.1039/D0EW00231C

12
logk=8.03(± 0.13)-3.24(± 0.28)o,m,p

Environmental Science: Water Research & Technology Accepted Manuscript


Dissociated phenols
R2=0.95 Undissociated phenols
10
Micropollutants

8
17a-ethinylestradiol
logk (M s )

Bisphenol A
-1 -1

6 Acetaminophen
Triclosan
4

2 logk=0.41(± 0.50)-4.69(± 1.13)o,m,p


R2=0.86
0
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

-2
-1.0 -0.5 0.0 0.5 1.0 1.5

o,m,p or  o,m,p


441

442 Fig. 2 Correlations between the second-order rate constants (logk) for the reactions of

443 ClO2 with non-dissociated phenols vs. +o,m,p (Eq. 30 and Table 3) and with

444 dissociated phenols vs. -o,m,p (Eq. 31 and Table 3). Data is obtained from ref.8, 99

445

446 The reaction pathways involved in ClO2 oxidation of phenols have been

447 investigated intensively. A two-step reaction mechanism has been proposed (Scheme

448 4a). One-electron transfer from phenolate to a ClO2 molecule first forms a phenoxyl

449 radical and chlorite, which is the rate-determining step of the reaction. Subsequently,

450 rapid oxygen transfer to the benzene ring from a second ClO2 molecule results in the

451 generation of HOCl and benzoquinone.81 The release of HOCl during ClO2 oxidation

452 of phenol has a yield of about 50% of the ClO2 consumed.15 The chlorite yields of

453 mono-hydroxy phenols upon ClO2 oxidation are in the range of 40–60%.42, 81 The

454 mechanisms for di- and tri-hydroxy phenols are similar. Bi- or tri-radicals are formed

455 which subsequently transfer to benzoquinones or ring cleavage products (Scheme 4c).

24
Environmental Science: Water Research & Technology Page 26 of 74

View Article Online


DOI: 10.1039/D0EW00231C

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

456

457 Note: modified from reference81, 102

458 Scheme 4 Summary of reaction scheme of ClO2 oxidation of phenolic compounds (a)

459 phenol; (b) hydroquinone; (c) catechol; (d) vanillin alcohol

460 The reactions of phenols can be classified based on their oxidized products.

461 Hydroquinone and mono-hydroxy phenols (without para-substitution) tend to

462 maintain their ring structure, with quinones as the major products (Scheme 4b).98

463 They are classified as Group I compounds. Di- and tri-hydroxy phenols (e.g.,

464 resorcinol and phloroglucinol) and para-substituted mono-hydroxy phenols (e.g.,

465 p-cresol) are classified as Group II, since ring cleavage is likely to occur together with

466 the formation of carboxylic acids (e.g., oxalic and maleic acids) during ClO2 oxidation

467 (Scheme 4c).96 ClO2 oxidation of phenolic lignins (e.g., vanillin alcohol) is one of the

25
Page 27 of 74 Environmental Science: Water Research & Technology

468 exceptions from that classification. Lactone ester and quinones are identified
DOI:as theView Article Online
10.1039/D0EW00231C

469 predominant products after their demethylation, rather than ring cleavage products

Environmental Science: Water Research & Technology Accepted Manuscript


470 (Scheme 4d).102

471 The products of phenol oxidation vary with the ClO2 dose. When the molar

472 ratio of ClO2 to phenol is above 3, benzoquinone is the main organic product. When

473 phenol was in excess (i.e., ClO2: phenol < 1), chlorinated phenols are also detected in

474 addition to benzoquinone.81 When ClO2 is in large excess, further decomposition of

475 benzoquinone may occur, leading to the formation of ring cleavage products.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

476 Degradation of benzoquinone by ClO2 takes 1–3 hours when the dose of ClO2 is 5

477 times that of the benzoquinone.42 Cyclopent-4-ene-1,3-dione derivatives and

478 chlorinated acetones are detected when excess ClO2 reacts with resorcinol and

479 mono-hydroxy phenols.96

480 Considering the prevalence of phenolic moieties in source waters, the toxicity

481 of their oxidized products is critical for the safety of treated water. Benzoquinone is

482 the most significant product and it can inhibit the growth of Pseudomonas

483 fluorescence.103 Quinones, however, might serve as carcinogens because their

484 metabolites can conjugate with glutathione and cysteine, leading to damage to certain

485 cells (e.g., liver, bladder and kidney).104, 105 In addition, quinones can act as mutagens

486 causing aneuploidy in human cells. Quinones formed in the ClO2 pre-treatment of

487 phenolic compounds can promote the formation of chloral hydrate in

488 post-chlorination.96 It should be noted that halogenated benzoquinones are reported to

489 have greater cytotoxicity than most of the regulated DBPs.106 ClO2 pre-treatment

490 might also increase the formation of halogenated benzoquinones during

491 post-chlorination. That deserves further evaluation.

492

26
Environmental Science: Water Research & Technology Page 28 of 74

493 3.2.3. Reaction with sulfur-containing moieties View Article Online


DOI: 10.1039/D0EW00231C

494 The sulfur content of NOM is relatively low (normally less than 3%) contributed

Environmental Science: Water Research & Technology Accepted Manuscript


495 by sulfur-containing proteins or amino acids. Treatment of some chemical solutions or

496 industrial wastewater may, however, involve oxidation of sulfur-containing organic

497 compounds.

498 Oxidation of thioether (R1–S–R2) to sulfoxides and sulfones (Scheme 5a) has

499 been observed in the ClO2 oxidation of dialkyl, diaryl and dibenzyl sulfides and of the

500 amino acid methionine at moderate temperature.107, 108 At a ClO2 to thioether molar
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

501 ratio of 1:2, the sulfoxide (R1–SO–R2) yield reaches 90–100% without significant

502 formation of sulfones.109 When ClO2 is in excess or at equimolar ratio with thioethers,

503 sulfones (R1–SO2–R2) are the major product.110 The degradation pathways of thioether

504 micropollutants (e.g., methiocarb, ametryn, isoproturon) suggest that sulfoxides are

505 the main oxidation products with small proportions of chlorinated sulfoxides.30, 111

506 From a mechanistic point of view, the sulfur atom contains lone electron pairs which

507 are readily attacked by ClO2. The reaction might initiate through electron transfer

508 from a thioether to ClO2 to yield a cation intermediate and chlorite. Based on the

509 material balance of the elements and the observed formation of chlorinated

510 by-products, HOCl is also released. However, experimental data on the inorganic

511 products from ClO2 oxidation of thioether are too scarce to confirm each step.

512 ClO2 oxidation of thiol-containing compounds occurs rapidly due to the strong

513 nucleophilicity of the thiol group (-SH). However, thiols are less soluble in water than

514 phenols and alcohols because the S-H bond is only slightly polar and has little

515 tendency to hydrogen bond. Reactions with thiols have therefore mainly been

516 investigated in organic solvents.112 Cysteine-based amino acids or peptides (e.g.,

517 glutathione and cystine) are common thiol-containing moieties found in source

27
Page 29 of 74 Environmental Science: Water Research & Technology

518 waters, and they play critical roles in biological activity. ClO2 oxidation of these
DOI: thiols
View Article Online
10.1039/D0EW00231C

519 is initiated by transferring one electron from disassociated RS- anion to ClO2 resulting

Environmental Science: Water Research & Technology Accepted Manuscript


520 in RS• radical and chlorite formation (Scheme 5b). ClO2 oxidation of thiols mainly

521 leads to disulfides, then sulfinic (RSO2H) and sulfonic acids (RSO3H) as further

522 oxidation products with excess ClO2 and at acidic pH (Scheme 5b). It should be noted

523 that the chlorite by-product can oxidize cysteine and glutathione following pathways

524 different from those of ClO2.113, 114 Chlorite is more likely to react with RSH directly,

525 rather than with RS- anion, via oxygen transfer to form corresponding disulfides and
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

526 sulfonic acids. Chloride is then the main reduced product in the oxidation of cysteine,

527 cystine and glutathione, since all can further reduce chlorite. However, more

528 experimental data are needed to clarify to what extent chlorite reacts with other

529 aliphatic and aromatic thiols. ClO2 reactions with other sulfur-containing compounds

530 including thiourea (CS(NH2)2) have also been studied. Sulfinic acid, sulfonic acids

531 and disulfides are reported to be main products.115

532 The thiol concentrations in source waters are quite low. Limited results show

533 that 18.7 M thiol are present per mM dissolved organic carbon (DOC) in Suwannee

534 River NOM.116 In consideration of the 1–10 mg/L DOC levels in source water, the

535 thiol concentrations may exist at sub M to M levels. Their ClO2 consumption

536 should be minor.

28
Environmental Science: Water Research & Technology Page 30 of 74

View Article Online


DOI: 10.1039/D0EW00231C

Environmental Science: Water Research & Technology Accepted Manuscript


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

537

538 Note: modified from references109, 114

539 Scheme 5 Summary of reaction scheme of ClO2 oxidation of sulfur-containing

540 compounds (a) sulfides and (b) thiols

541

542 3.2.4. Reaction with nitrogen-containing moieties

543 Nitrogenous organic compounds are becoming more prevalent in source waters

544 due to increasing industrial and agricultural discharges,117 which poses more

545 challenges to control nitrogenous DBP formation in water treatment.

546 Table 4 lists the species-specific second-order rate constants (kClO2-A-) and the

547 apparent second-order rate constants (kapp) at pH 7.0 for ClO2 with some aliphatic

548 amines. In general, the reaction rates decrease in the order tertiary amine > secondary

549 amine > primary amine. For tertiary amines, the reaction rate constants are in the

550 range of 103–106 M-1s-1 at neutral pH, which is 2–5 orders of magnitude higher than

551 those of secondary or primary amines. Primary and secondary amines are, however,

552 more reactive in chlorination than tertiary amines.35 ClO2 reacts much faster with

553 deprotonated amines than with neutral species because the deprotonated amines are
29
Page 31 of 74 Environmental Science: Water Research & Technology

554 stronger electron donors.118, 119 Therefore, pH is a critical factor in ClO2 oxidation ofView Article Online
DOI: 10.1039/D0EW00231C

555 amines, and the reaction rates tend to increase with increasing pH. But substituents

Environmental Science: Water Research & Technology Accepted Manuscript


556 induce strong effects on the reaction rate constants. Higher electron density on the

557 nitrogen atom of an amine favors attack ClO2 molecules. The Taft constants of

558 nitrogen substituents are plotted against the ClO2 oxidation rates of amines in Fig. 3.

559 The Taft’s constants were calculated as Eq.32:

560 σ*=∑σ*R1+σ*R2+σ*R3 (32)

561 where ∑σ*R represents the Taft constants of each substituents on the amine’s nitrogen
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

562 atom. The reaction rate constants increase with decreasing ∑σ*, suggesting that

563 amines with stronger electron-donating substituent groups (with smaller σ*) tend to be

564 more reactive with ClO2. Two different correlations are evident between tertiary

565 amines and secondary (and primary) amines because of the large gap between their

566 reactivities towards ClO2. The slope of the correlation for tertiary amines is lower

567 than that for the others, indicating that the effects of substituents on ClO2 reactivity

568 are more significant for secondary and primary amines. Benzylamines, however,

569 deviate from the lines for secondary and primary amines, perhaps due to different

570 reaction mechanisms. The reaction kinetics for ClO2 oxidation of ciprofloxacin

571 (secondary amine) and roxithromycin (tertiary amine) can be well predicted based on

572 the QSAR relationship of amines within a factor of 1/3–3 of measured rates.

573

574 Table 4 Kinetics of ClO2 oxidation of selected nitrogen-containing compounds

Taft kapp (M-1s-1) T


Compounds pKa kClO2-A-(M-1s-1) Ref
constants at pH 7.0 (oC)
Benzylamine 1.25 9.6 3.9×10-2 9.8×10-5 25 10

Methylamine 0.98 10.6 5.0×10-1 1.3×10-4 24 8

2-Aminoethanol 1.23 9.4 1.4×10-2 5.1×10-5 25 10

Urea - < 1×10-6 8

30
Environmental Science: Water Research & Technology Page 32 of 74

View Article Online


N-Methyl-4-methoxybe DOI:
0.73 10.0 2.7×102 2.9×10-1 25 10 10.1039/D0EW00231C
nzylamine
Dimethylamine 0.49 10.7 5.0×102 10.0×10-2 24 8

Environmental Science: Water Research & Technology Accepted Manuscript


Isopropylbenzylamine 0.57 9.7 9.1 1.9×10-2 25 10

Diisopropylamine 0.11 11.0 3.5×102 3.4×10-2 25 10

Diethylamine 0.29 10.5 1.0×103 3.2×10-1 120

Dibenzylamine 1.03 8.4 8.3 3.0×10-1 9

Benzyl-tert-butylamine 0.46 10.2 2.8×102 1.8×10-1 25 10

N,N-Dimethyl-3-metho
0.24 9.0 2.9×104 2.6×102 25 10
xybenzylamine
N,N-Dimethyl-4-metho
0.24 9.3 4.9×104 2.3×102 25 10
xybenzylamine
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

N,N-Dimethyl-4-methyl
0.24 9.2 3.5×104 2.1×102 25 10
benzylamine
N,N-Dimethyl-3-nitrob
0.48 8.2 6.2×103 3.7×102 25 10
enzylamine
N,N-Dimethyl-4-nitrob
0.50 8.1 4.5×103 3.6×102 25 10
enzylamine
N,N-Dimethyl-benzyla
0.27 9.0 2.7×104 2.5×102 25 10
mine
N,N-Dimethyl-tert-buty
-0.3 10.7 2.3×105 4.7×101 25 10
lamine
N,N-Dimethyl-3-Cl-ben
0.39 8.7 1.6×104 3.4×102 25 10
zylamine
N,N-Dimethyl-4-Cl-ben
0.35 8.8 2.0×104 2.9×102 25 10
zylamine
N,N-Dimethyl-4-F-benz
0.32 8.9 2.0×104 2.3×102 25 10
ylamine
Triethylamine -0.3 11.1 2.0×105 1.6×101 120

Trimethylamine 0 9.7 6.0×104 1.2×102 24 8

Dimethylethanolamine 0.25 9.2 9.7×103 6.1×101 119

Aniline 4.5×105 43

4-Aminoaniline 3.5×108 43

N,N-Dimethylaniline 4.4×106 43

3,4-Dicarboxylaniline 1.5×105 121

Luminol 1.0×106 121

Indole 4.9 1.2×104 59

1-Methylindole 1.6×104 59

31
Page 33 of 74 Environmental Science: Water Research & Technology

View Article Online


2-Methylindole 5.7 8.1×105 59
DOI: 10.1039/D0EW00231C

3-Methylindole 5 1.9×106 59

2,3-Dimethylindole 6.1 1.1×108 59

Environmental Science: Water Research & Technology Accepted Manuscript


Piperidine 0.31 1.5×10-1 59

1-Methylpiperidine 8.7×104 122

4-Methylpiperidine 4.9 3.4×106 122

N-Methylpiperidine -0.18 10.4 8.7×104 25 10

1-Hydroxypiperidine 0.31 4.0×104 104

1-Oxylpiperidine 5.0×106 122

Imidazole 1×10-4 24 8

N-tert-Butylpyrrolidine 1.3×106 10

Benzo-2,3-dihydrophth
3.0×105 123
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

alazine-1,4-dione
2,3-dihydrophthalazine-
1.5×105 123
1,4-dione

575

576

577 Fig. 3 Correlations between the apparent second-order reaction rate constants (logkapp)

578 for the reactions of ClO2 with primary and secondary amines and with tertiary amines

579 vs. ∑σ* (Table 4). Data is obtained from ref.8, 99

580

581 The release of HOCl has been observed in ClO2 oxidation of many organic

582 compounds and NOM in recent studies. The released HOCl has been found to play an
32
Environmental Science: Water Research & Technology Page 34 of 74

583 important role in removing atenolol, metoprolol and ciprofloxacin, which areView Article Online
DOI: 10.1039/D0EW00231C

584 secondary amines and reacting 45 orders of magnitude faster with HOCl than with

Environmental Science: Water Research & Technology Accepted Manuscript


585 ClO2.41 Thus, the removal of amines in practical ClO2 treatment of waters is excepted

586 to be better than the predicted removal based on kapp with ClO2, because small

587 fractions of HOCl is often produced during ClO2 oxidation.15, 41 This phenomenon can

588 happen for compounds with very high reaction rate constants with HOCl than with

589 ClO2. By comparing the kinetics data of ClO2 reactions summarized in this study with

590 those of free chlorine reactions summarized in previous studies35, 99, 124, it was found
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

591 that primary amines and secondary amines generally have orders of magnitude higher

592 reaction rate constants with free chlorine than with ClO2 at pH 7.0. Thus, it is

593 reasonable to propose that the presence of HOCl in ClO2 reaction system can

594 contribute greatly to the degradation of primary amines and secondary amines.

595 One-electron oxidation is generally considered to be the initial step in most

596 reactions of aliphatic amines with ClO2, leading to the formation of an aminyl cation

597 radical and chlorite (Scheme 6a). After the rejection of an α-proton, the cation radicals

598 undergo further oxidation and generate aldehydes and an inferior amine. In the case of

599 benzylamines (Scheme 6b), whose α-hydrogens are activated by the adjacent phenyl

600 group, hydrogen abstraction also takes place to initiate the reaction, supported by a

601 significant isotope effect.9 It has been reported that hydrogen abstraction is

602 responsible for 73% of the oxidation of benzylamine and 25% of that of

603 benzyl-tert-amine.9

604

33
Page 35 of 74 Environmental Science: Water Research & Technology

605 Note: adapted from reference9 View Article Online


DOI: 10.1039/D0EW00231C

606 Scheme 6 Reaction pathway proposed for ClO2 oxidation of amines

Environmental Science: Water Research & Technology Accepted Manuscript


607 It should be noted that amines containing a dimethylamine group are considered

608 important precursors of nitrosamines, a potent carcinogen, in water treatment.125 ClO2

609 oxidizes most tertiary aliphatic amines quickly and converts them to secondary

610 amines, and it oxidizes secondary amines to primary amines. Both primary and

611 secondary amines are generally less reactive nitrosamine precursors in subsequent

612 chloramination as they don’t generate nitroso groups. Thus, ClO2 pre-treatment of
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

613 tertiary amines effectively reduces their nitrosamine formation potential.119 However,

614 reduction on nitrosamine formation with ClO2 oxidation is unpredictable when

615 different real waters are being treated. The effect depends on the type of contaminants

616 and the ClO2 oxidation conditions.119, 126, 127 Further investigations covering more

617 types of amines and oxidation conditions might help to improve understanding of the

618 mechanism of nitrosamine formation in ClO2 oxidation.

619 Turning to aromatic amines, they are widely distributed in aqueous environments

620 as degradation products of dyes in industrial wastewater or of herbicides in

621 agricultural discharges.128 Aniline and 4,4-methylenedianiline are listed in

622 Contaminant Candidate List-4 of U.S. EPA.129 Various reaction rate constants for

623 aniline with ClO2 have been reported. Huie and his co-workers determined the rate of

624 the first step of aniline oxidation to be 4.5×105 M-1s-1 at pH 6.9.43 Lee and von Gunten

625 tried to established a QSAR for ClO2 oxidation of anilines based on limited available

626 data,99 thus the credibility of the QSAR is excepted to increase with more studies on

627 this subject. ClO2 oxidation of anilines also follows a one-electron transfer

628 mechanism as its first step (Scheme 7). Since the amino group is directly attached to a

629 benzene ring, that results in a shift of electron density from the nitrogen atom to the

34
Environmental Science: Water Research & Technology Page 36 of 74

630 benzene ring. The reaction pathways and products are then different from thoseView Article Online
DOI: 10.1039/D0EW00231C

631 observed with aliphatic amines. Benzoquinone is the major product with

Environmental Science: Water Research & Technology Accepted Manuscript


632 p-aminophenol and azobenzene dimer as important intermediates.130

633

634 Note: modified from reference130


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

635 Scheme 7 Reaction pathway proposed for ClO2 oxidation of aniline

636

637 Among the amino acids, ClO2 reacts rapidly with cysteine, tyrosine and

638 tryptophan (104–107 M-1s-1) and reacts slowly with histidine, proline, alanine and

639 glycine (10-5–10-2 M-1s-1), as shown in Table 5. Other amino acids do not display

640 measurable reaction rates with ClO2.131 Amino acids are a classic type of primary

641 amine, however, the amino group is unreactive towards ClO2. The ClO2-reactive

642 amino acids contain other reactive groups such as phenolic or sulfur-containing

643 groups, and the oxidation occurs at those sites other than the primary amine groups.

644 For example, the reactive site of cysteine, a most reactive amino acid toward ClO2, is

645 believed to be the thiol group where the reaction involves an electron abstraction

646 (Scheme 8a). The cysteinyl radical forms and rapidly reacts with another ClO2

647 molecule to form a cysteinyl-ClO2 adduct, which then yields cysteic acid at acidic pH

648 or cystine under alkaline conditions.114 ClO2 oxidation of tyrosine occurs

649 predominantly at its phenolic structure, resulting in formation of dopaquinone and

650 dopachrome at pH 6–7 (Scheme 8b).132 For tryptophan, ClO2 first abstracts one

35
Page 37 of 74 Environmental Science: Water Research & Technology

651 electron from the indole ring to form a neutral tryptophan radical. The radical isView Article Online
DOI: 10.1039/D0EW00231C

652 unstable and immediately reacts with a second mole of ClO2 to generate

Environmental Science: Water Research & Technology Accepted Manuscript


653 N-formylkynurenine, kynurenine, 3-hydroxykynurenine and other products.133

654 Table 5 Kinetics of ClO2 oxidation of selected amino acids, peptides, and proteins

Apparent

second-order
Refere
Compounds pKa reaction rate pH T (oC)
nces
constants,

kapp (M-1s-1)
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Amino acids

Cysteine 8.2 1.0×107 7.0 25.0 114

Tyrosine 9.1 1.4×105 7.0 25.0 132

N-Acetyltyrosine 2.5×105 6.2 25.0 132

Tryptophan 9.4 3.4×104 7.0 25.0 133

a-Alanine 9.9 < 1×10-2 8.0 25.0 8

Glycine 9.8 < 1×10-5 8.0 22–24 8

Proline 10.6 3.4×10-2 6.0 25.0 131

Hydroxylproline 6.9×10-2 6.0 25.0 131

Histidine 6.0 5.4×10-2 6.0 25.0 131

Peptides

Glutathione 1.4×108 5.9 25.0 11491

Proteins

Bovine serum albumin 6.4 7.0 25.0 134

Glucose-6-phosphate
9.7 7.0 25.0 134

dehydrogenase

655

656

36
Environmental Science: Water Research & Technology Page 38 of 74

View Article Online


DOI: 10.1039/D0EW00231C

Environmental Science: Water Research & Technology Accepted Manuscript


657
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

658 Note: modified from reference114, 132

659 Scheme 8 Reaction pathway proposed for ClO2 oxidation of amino acids (a) cysteine;

660 (b) tyrosine

661

662 Denaturation of peptides or proteins is caused primarily by covalent oxidative

663 modification of their reactive cysteine, tyrosine and tryptophan residues. For

664 glutathione, a tripeptide formed by glycine, cysteine and glutamic acid, the primary

665 reactive site towards ClO2 is the thiol group and the oxidation products are similar to

666 those of cysteine.114 Treating bovine serum albumin (BSA) and glucose-6-phosphate

667 dehydrogenase (G6PD) with ClO2 converts tryptophan residues to

668 N-formylkynurenine, and tyrosine residues become 3,4-dihydroxyphenylalanine

669 (DOPA) or 2,4,5-trihydroxyphenylalanine (TOPA).134

670 Some heterocyclic nitrogen compounds (e.g., indoles, imidazoles and

671 piperidines) are also reactive towards ClO2. The oxidation rate constant of indoles is

672 about 104–108 M-1s-1,135 but oxidation of imidazole is very slow (10-4 M-1s-1)8 and the

673 rate constants of reaction with piperidine are in the range of 103–106 M-1s-1.10

674 Derivatives of these heterocyclic nitrogen compounds are found extensively in aquatic

37
Page 39 of 74 Environmental Science: Water Research & Technology

675 environments. For instance, piperidine, pyrrolidine and imidazole derivatives can
DOI: beView Article Online
10.1039/D0EW00231C

676 found in alkaloids.136 There are, however, only limited reports about the ClO2

Environmental Science: Water Research & Technology Accepted Manuscript


677 oxidation pathways of heterocyclic nitrogen compounds. Nevertheless, the reaction

678 pathways of ClO2 with histidine and tryptophan (Scheme 9) can give some evidences

679 for the reaction mechanisms of ClO2 oxidation of indole and imidazole structures.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

680

681 Note: modified from reference133, 137

682 Scheme 9 Reaction pathway proposed for ClO2 oxidation of (a) tryptophan and (b)

683 histidine

684

685 4. ClO2 oxidation of emerging contaminants

686 The emerging contaminants (ECs) to be discussed here are primarily synthetic

687 organic chemicals including pharmaceuticals and personal care products (PPCPs),

688 endocrine disrupting compounds (EDCs), pesticides, herbicides, cyanotoxins and

689 antibiotic resistant genes (ARGs), but there are many others. ECs have increasingly

690 been detected in aquatic environments worldwide, including in municipal

691 wastewaters, surface water, ground waters and even in water supply systems, at

692 concentrations ranging from a few ng·L-1 to a few hundred μg·L-1.138-144 The release

38
Environmental Science: Water Research & Technology Page 40 of 74

693 of ECs into the environment may cause ecological risk. ECs can interfere withView Article Online
DOI: 10.1039/D0EW00231C

694 endocrine systems of organisms, promote microbiological resistance, and accumulate

Environmental Science: Water Research & Technology Accepted Manuscript


695 in plants and animals.145-147 Since the early 20th century, ARGs as ECs have been

696 found in municipal wastewaters,148, 149 surface waters and associated sediments,150,

697 151 and drinking water distribution systems.152, 153 During water treatment,

698 conventional water treatment processes like coagulation, flocculation, sedimentation

699 and filtration are generally not effective to deactivate or degrade ARGs. Oxidation by

700 ClO2 is, however, effective to remove some ECs from waters.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

701 The diverse moieties in ECs lead to various reaction rate constants of ECs toward

702 ClO2. Based on chlorine dioxide’s reactivity with the main functional groups, the

703 reactivity at pH 7.0 usually decreases in the order: phenols>anilines>tertiary amines

704 > secondary amines, as shown in Table 6. ClO2 reacts rapidly with ECs containing

705 electron-rich phenolic groups, such as 17α-ethinylestradiol, and the kapp at pH 7.0 for

706 these ECs generally exceed 104 M-1s-1.99, 154, 155 ECs containing amine groups

707 generally have reaction rate constants with ClO2 at around 103 M-1s-1 for anilines such

708 as sulfonamides,156 102 M-1s-1 for tertiary amines such as roxithromycin,99 and 10

709 M-1s-1 for secondary amines such as ciprofloxacin.157 The kapp of pharmaceuticals at

710 pH 7.0 generally range from high to low in the order tetracyclines, estrogens,

711 sulfonamides, macrolides and then fluoroquinolones. The reaction rate constants

712 within the same group also vary depending on the structures involved. For example,

713 among the fluoroquinolones, the reaction rate constants of ofloxacin and enrofloxacin,

714 which contain tertiary amines on their piperazine moieties (77.5 and 62.7 M-1s-1,

715 respectively) are higher than those of ciprofloxacin, norfloxacin, lomefloxacin or

716 pipemidic acid which contain secondary amines (1.5–13.0 M-1s-1).157 This is in good

717 agreement with the trend of faster oxidation of tertiary amines than secondary amines

39
Page 41 of 74 Environmental Science: Water Research & Technology

718 by ClO2. For phenylurea herbicides containing secondary amines, diuron shows
DOI: much
View Article Online
10.1039/D0EW00231C

719 higher reactivity towards ClO2 than chlortoluron or isoproturon due to the different

Environmental Science: Water Research & Technology Accepted Manuscript


720 functional groups on the aromatic ring.158 Regarding the algal toxin microcystin-LR,

721 the reaction rate constant with ClO2 is low at 1.24 M-1s-1 at pH 7.0.159
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

40
Environmental Science: Water Research & Technology Page 42 of 74
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


722 Table 6 Kinetics of ClO2 reactions with emerging contaminants

Species-specific second-order rate constants (M-1s-1) Apparent second-order


T Refere
Compound pKa reaction rate constants, pH
H2A+ HA A- A2- (oC) nces
kapp (M-1s-1)

Phenols (PPCPs and EDCs)

Acetaminophen 9.7 2.6 1.1×108 2.1×105 7.0 99

Bisphenol A 9.6, 10.2 4.3 2.8×108 1.8×105 7.0 99

Triclosan 8.1 1.2 8.5×106 6.3×105 7.0 99

4-n-Nonylphenol 10.7 5.9 3.6×108 7.1×104 7.0 99

17α-ethinylestradiol 10.4 <2×102 4.6×108 2.0×105 7.0 20 154

17-estradiol 10.71 ~105 7.0 20 154

Estrone 10.8 ~105 7.0 20 154

Amoxicillin >104 8.0 20 155

Cefadroxil >104 8.0 20 155

Tetracycline 3.3, 7.7, 7.2×106 2.7×107 1.3×106 7.0 22 160

41
Page 43 of 74 Environmental Science: Water Research & Technology
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


9.3

3.3, 7.3,
Oxytetracycline 3.7×106 1.6×107 1.2×106 7.0 22 160

9.1

3.3, 7.4,
Chlorotetracycline 1.05×106 2.57×107 3.2×105 7.0 22 160

9.3

3.5, 6.9,
Iso-chlorotetracycline 3.05×105 2.13×107 2.2×105 7.0 22 160

9.4

Anilines (PPCPs)

Sulfamethoxazole 1.9, 5.6 negligible 4.8×101 6.5×103 6.1×103 7.0 20 156

Sulfamethizole 1.9, 5.3 negligible 9.0×101 3.8×103 3.9×103 7.0 20 156

Sulfadimethoxine 2.1, 6.1 negligible 3.5×101 5.3×103 4.4×103 7.0 20 156

Sulfamethazine 2.1, 7.5 negligible 9.5×102 1.2×104 4.1×103 7.0 20 156

Sulfamerazine 2.1, 6.9 negligible 9.4×102 8.9×103 5.6×103 7.0 20 156

Diclofenac 4.2 1.1×104 1.1×104 7.0 20 154

Iopromide <1.0×10-2 7.0 20 154

Tertiary amines (PPCPs)

Roxithromycin 8.8 1.4×104 2.2×102 7.0 20 154

42
Environmental Science: Water Research & Technology Page 44 of 74
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


Enrofloxacin 6.1, 7.7 <1.0×10-4 1.6×101 3.4×102 6.3×101 7.0 22 157

Ofloxacin 6.2, 7.8 <1.0×10-4 1.8×101 5.4×102 7.8×101 7.0 22 157

Dimethylaminophenazo
>102 7.4 20 154

ne

Secondary amines (PPCPs)

Ciprofloxacin 6.1, 8.6 <1.0×10-4 4.5 1.9×102 7.9 7.0 22 157

Norfloxacin 6.2, 8.5 <1.0×10-4 1.01×101 1.78×102 1.3×101 7.0 22 157

Lomefloxacin 5.5, 8.8 <1.0×10-4 3.9 1.9×102 6.8 7.0 22 157

Pipemidic acid 5.4, 8.2 <1.0×10-4 1.54×10-3 24.4 1.5 7.0 22 157

Atenolol 9.6 1.0 8.0 20 41

Metoprolol 1.3 8.0 20 41

Ifosfamide <1 7.4 20 154

Others (PPCPs)

Trimethoprim 3.2, 7.1 3.76 7.00 34.1 1.9×101 7.0 22 161

Gemifibrozil 5.9×101 7.0 99

Naproxen 6.1×102 7.0 99

Antipyrine 1.4 4.8×10-1 7.0 25 162

43
Page 45 of 74 Environmental Science: Water Research & Technology
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


Penicillin-G <0.1 6.4 20 155

Bezafibrate 3.6 <1.0×10-2 7.0 20 154

Carbamazepine <1.5×10-2 7.0 20 154

Diazepam <2.5×10-2 7.0 20 154

Ibuprofen 4.9 <1.0×10-2 7.0 20 154

Caffeine <1 7.4 20 154

Clofibric acid <1 7.4 20 154

Cyclophosphamide <1 7.4 20 154

Fenoprofen <1 7.4 20 154

Glibenclamide <10 7.4 20 154

Ketoprofen <1 7.4 20 154

Pentoxifylline >102 7.4 20 154

Propylphenazone >102 7.4 20 154

Herbicides

Chlortoluron 1.4 8.0 25 158

Isoproturon 2.1 8.0 25 158

44
Environmental Science: Water Research & Technology Page 46 of 74
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

Environmental Science: Water Research & Technology Accepted Manuscript


Pesticides

Methiocarb 1.1 7.4 22 30

Cyanotoxins

Cylindrospermopsin 0.9 8.0 20 163

Anatoxin-a 9.36 4.7 1.2 8.0 20 163

Microcystin-LR 1.1 7.0 25 159

45
Page 47 of 74 Environmental Science: Water Research & Technology

724 pH has a significant impact on the reactivity of ECs with ClO2. The secondary View Article Online
DOI: 10.1039/D0EW00231C

725 reaction rate constants of most ECs increase rapidly from pH 5.0 to 9.0, and then the

Environmental Science: Water Research & Technology Accepted Manuscript


726 increase slows above pH 9.0, as shown in Fig. 4. For example, the secondary rate

727 constant of ciprofloxacin increases by more than 3 orders of magnitude from pH 5.0

728 to 10.0 and those of tetracyclines (e.g., tetracycline, oxytetracycline and

729 chlorotetracycline) increase by 4-6 orders of magnitude from pH 2.5 to 10.5. The

730 reaction rate constant of 17α-ethinylestradiol increases linearly with pH. Diclofenac,

731 however, shows no pH dependence and its second-order rate constant with ClO2
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

732 remains 1.05×104 M-1s-1 at any pH over 5.0. The reaction rates of penicillin-G

733 potassium salt with ClO2 at pH 3.5, 5.4 and 6.4 are 3.5, 0.35, and < 0.1 M-1s-1,

734 respectively, decreasing as the pH increases.155 For phenylurea herbicides, the

735 degradation rates of chlortoluron and isoproturon increase progressively with

736 increasing pH, while the degradation rate of diuron decreases.158 For antipyrine,

737 higher pH accelerates the reaction with ClO2 when the pH is less than 9.0, but in a

738 strongly alkaline environment (pH > 9.0) the oxidation process slows with increasing

739 pH.

740 Such pH-dependent reactivity can be explained by the protonation state of the

741 reactive groups involved, like the phenolic group of estrogens, the aniline group of

742 sulfonamides and the secondary or tertiary amino group of fluoroquinolones. The

743 specific rate constants of the deprotonated species (A-) are generally higher than those

744 of the neutral species (HA), and the reaction rate constants between cation species

745 (H2A+) and ClO2 are generally negligible. Therefore, the shifting to anionic species

746 (A-) with increasing pH from 5.0 to 10.0 generally enhances the rates of ClO2’s

747 reactions with phenolic-containing and amine-containing ECs such as estrogens,

748 sulfonamides and fluoroquinolones. Diclofenac with a pKa of 4.0 contains a carboxyl

46
Environmental Science: Water Research & Technology Page 48 of 74

749 group acting as a proton donor, and its reaction rate with ClO2 varies littleDOI:at10.1039/D0EW00231C
pHsView Article Online

750 higher than 5.0. Higher pH results in higher oxidation-reduction potential of ClO2

Environmental Science: Water Research & Technology Accepted Manuscript


751 solution, making it easier to oxidize substrates.164 However, ClO2 may decompose to

752 chlorite and chlorate when the pH exceeds 9.0, causing a sharp decline in ClO2

753 concentration, which inhibits the degradation process.165 That might explain the

754 slower increase in degradation rate observed with most ECs at pH > 9.0. Nevertheless,

755 the species-specific reaction rate constants shown in Table 6 can be applied to

756 calculate apparent second-order reaction rate constants of ECs at different pHs.
757
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

758

759 Fig. 4 pH dependence of the apparent second-order reaction rate constants (logkapp)

760 and the half-life times (logt1/2) for reactions of ClO2 with selected emerging

761 contaminants at 20–25oC. The half-lives are calculated at a chlorine dioxide

762 concentration of 1 mg∙L-1 (15 μM).

763 The reactions between ClO2 and ECs usually proceed via electron transfer

764 pathways. The EC provides one electron to ClO2 and is activated to a radical. The

47
Page 49 of 74 Environmental Science: Water Research & Technology

765 newly-formed radical can couple with ClO2 to further produce oxidation products. View Article Online
DOI: 10.1039/D0EW00231C

766 The reactive sites of typical ECs for ClO2 attack are presented in Fig. 5.

Environmental Science: Water Research & Technology Accepted Manuscript


767 Taking 17α-ethinylestradiol, a highly reactive phenolic compound, for an

768 example, the phenolic group of 17α-ethinylestradiol is the first attack site for ClO2

769 producing single-electron transfers from the electron-rich phenolic group to ClO2. The

770 oxidation of 17α-ethinylestradiol by ClO2 may produce quinone-type products such as

771 hydroquinone, which is similar to those produced through ozone oxidation.166 The

772 aniline group of sulfamethoxazole is reported to be attacked initially by ClO2 resulting


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

773 in the formation of aminyl radical cations.167 The radical cations then deprotonate and

774 hydroxylate to form hydroxylamine, which can be further oxidized to form nitroso

775 groups.168 Moreover, the cleavage of C-S and S-N bonds may also occur, and the

776 resulting intermediates may further oxidize to sulfate ion, nitrate ion, oxalic acid and

777 other products.156 As with sulfamethoxazole, aniline moiety of diclofenac is the

778 principal reactive site involved in hydroxylation and chlorination followed by the

779 formation of aldehyde intermediates.31, 169, 170 Finally, ring structures are opened and

780 further oxidation continues. With roxithromycin, a macrolide antibiotic, ClO2 first

781 attacks the tertiary amino group, and leads to the cleavage of the C-N bond. After

782 further oxidation, roxithromycin loses the entire amino group.154 With the

783 fluoroquinolones, the piperazine ring is the major reaction site and ClO2 first attacks

784 tertiary amines (in enrofloxacin and ofloxacin) or secondary amines (in ciprofloxacin,

785 norfloxacin, lomefloxacin and pipemidic acid), generating a radical.171 Further

786 oxidation of the amine group by another ClO2 molecule results in fragmentation of the

787 piperazine ring and the formation of imine intermediates. Subsequent hydrolysis of

788 the imine intermediates leads to the formation of the dealkylated product observed in

789 the oxidation of amines by ClO2.9, 172 In spite of bearing a primary amine group in

48
Environmental Science: Water Research & Technology Page 50 of 74

790 trimethoprim, the pyrimidine ring is suspected to be most susceptible for ClODOI:
2 attack.
View Article Online
10.1039/D0EW00231C

791 Oxidation of trimethoprim by ClO2 leads mostly to chlorine and hydroxyl

Environmental Science: Water Research & Technology Accepted Manuscript


792 substitutions at the diaminopyrimidinyl moiety as well as transformation of amino

793 groups (-NH2) to nitroso groups (-NO).161


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

794

795 Note: Red colors indicate the reactive sites for ClO2 attack.

796 Fig. 5 Typical emerging contaminants

797 In case of pesticides containing secondary amine group such as chlortoluron or

798 diuron, the electron-rich nitrogen atom first undergoes single-electron transfer and the

799 free radical intermediates transform into hydroxyl derivatives.111, 137 The subsequent

800 breaking of the N-C bond in the intermediates gives rise to smaller and more

801 hydrophilic products, mainly carboxylic acids and aldehydes.158 Methiocarb (MC) is a

802 commonly used sulfur-containing pesticide. During the ClO2 oxidation, addition of

803 ClO2 and OH- around the sulfur center results in an intermediate adduct MC–ClO2–

804 OH. With its cleavage, the more toxic methiocarb sulfoxide is generated by losing

49
Page 51 of 74 Environmental Science: Water Research & Technology

805 HClO2, while the less toxic methiocarb sulfone is generated by HOCl loss.111, 173 InView Article Online
DOI: 10.1039/D0EW00231C

806 alkaline conditions, methiocarb sulfoxide and methiocarb sulfone can be further

Environmental Science: Water Research & Technology Accepted Manuscript


807 oxidized to 2,6-dimethylbenzoquinone, 2,6-dimethylhydroquinone,

808 4-chloro-3,5-dimethylphenol and other unidentified products.30

809 Microcystin-LR is an abundant algal toxin. Its conjugated double bonds are

810 reported to be the main reactive site during ClO2 oxidation. The major degradation

811 products are dihydroxy isomers of microcystin-LR, which are less toxic than the

812 parent compound.159


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

813 The degradation efficiency of ClO2 on ARGs is quite low. The kapp values for

814 ClO2 degradation of four 266—1017 bp amplicons of a chromosomal ARG of

815 multidrug-resistant Bacillus subtilis 1A189 ranged from 0.010 to 0.013 M−1s−1 per bp

816 during ClO2 treatment, which are lower than the rates for hydroxyl radical, ozone and

817 free chlorine.174 ARGs degradation rates in ClO2 oxidation correlates strongly with

818 the contents of 5’-guanine.174 This observation suggests that guanine is the most

819 susceptible site in ARGs degradation by ClO2 and the apparent second-order reaction

820 rate for ClO2 oxidation of guanosine 5’-monophosphate is about 4.5×102 M−1s−1 at pH

821 7.0.175 Therefore, theoretically, the ARGs degradation efficiency in ClO2 oxidation

822 depends on the nucleotide contents of specific ARGs.

823 The removal efficiency of ECs is highly dependent on the water and wastewater

824 matrix. In a study evaluating the removal of 56 PPCPs during ClO2 treatment of

825 biologically treated wastewater, better removal efficiency was observed in low COD

826 (35 mg/L COD) wastewater from the plant than that in high COD wastewater (55

827 mg/L COD) at the same ClO2 dose.176 Additionally, the removal is also highly

828 dependent on the functional groups in PPCPs (Table 5). In the 56 PPCPs’ study, about

829 1/3 of the PPCPs barely degraded even at a ClO2 dose of 20 mg/L, whereas others

50
Environmental Science: Water Research & Technology Page 52 of 74

830 were reduced by over 90% at a ClO2 dose of 0.5 mg/L.176 One advantage DOI:
of 10.1039/D0EW00231C
ClOView
2
Article Online

831 treatment is that the removal efficiency of ECs is barely affected by the presence of

Environmental Science: Water Research & Technology Accepted Manuscript


832 ammonia and nitrite in water and wastewater.177 Contractively, great reduction in

833 ECs’ removal efficiency is found with increasing concentrations of ammonia and

834 nitrite during chlorination treatment.177

835

836 5. Implications and future research perspectives

837 In order to give a whole picture of the transformation of diverse groups of


Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

838 compounds in water, Fig. 6 summarizes the available reaction rate constants for ClO2

839 and also presents their half-lives at a typical ClO2 dose of 1 mg∙L-1 and at pH 7.0. In

840 consideration of a typical recommended disinfection period of 30 min for ClO2 as

841 primary disinfectant, phenolic, anilines and heterocyclic nitrogen moieties in waters

842 are decomposed to great extent, while oxidation of olefins and ketones are more likely

843 to occur when ClO2 residual is maintained in long-distance water distribution system.

844 It should be noted that the half-lives of tertiary amines are around 30 min, indicating

845 that only partial of tertiary amines are decomposed in that disinfection period,

846 whereas primary and secondary amines are barely transformed. ClO2 can also lead to

847 promising removal of some micropollutants bearing reactive groups. Sulfonamides

848 (aniline structures) and tetracyclines (tertiary-amine and phenol-like structures) can be

849 quickly decomposed at the typical ClO2 reaction conditions. Fluoroquinolones,

850 -lactams and other pharmaceuticals have varied half-lives, dependent on their

851 specific structures. The degradation of pesticides (e.g., antipyrine and metoprolol) and

852 algal toxins (e.g., microsystins) generally takes hours to complete. As illustrated in

853 Fig. 6, the degradation extents of organic compounds bearing diverse functional

854 groups can be estimated. Therefore, we will know what types of compounds are

51
Page 53 of 74 Environmental Science: Water Research & Technology

855 degraded when ClO2 is used for primary disinfection (0.5–1.4 mg∙L-1, 15–30 min) andView Article Online
DOI: 10.1039/D0EW00231C

856 secondary disinfection (0.2–0.5 mg∙L-1, 30–120 min). It will also give us guidance

Environmental Science: Water Research & Technology Accepted Manuscript


857 about the doses and reaction time suitable for target pollutants removal. For example,

858 when ClO2 is applied as secondary disinfectant, enhanced decomposition of tertiary

859 amines, olefins and some micropollutants (e.g., fluoroquinolones and trimethoprim)

860 are expected, while degradation of ketones, pesticides (e.g., antipyrine and

861 metoprolol) and algal toxins (e.g., microsystins) generally takes hours to complete,

862 which are more likely to occur in long-distance water distribution system with
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

863 adequate ClO2 residual.

1010 10-7
109 10-6
108 10-5
107 10-4
106 10-3
105 10-2
104 10-1
kapp (M-1s-1)

103 100
102 30min 101 t1/2 (min)
101 120min
102
100 103
10-1 104
10-2 105
10-3 106
10-4 107
10-5 108
10-6 Organics Emerging contaminants 109
10-7
ic es es es ins es ols es en es ms es es ics als es ins
an in in in f d n in g n a id in t c id t
org am am am Ole ehy he nil itro olo act am ycl ibio uti tic cys
In ry y y l d P A n uin -l n c nt c e es ro
a ar iar d
a ic
cl roq β ulfo etra r a ma P ic
im nd rt an cy luo S T the ar M
Pr eco Te s ro F O ph
S ne te ic
eto He i ot
K ib
nt
n-a
No
864

865 Fig. 6 The apparent second-order reaction rate constants (kapp) and the half-life times

866 (t1/2) for reactions of ClO2 with model compounds bearing different functional groups.

867 t1/2 was calculated by assuming a ClO2 concentration of 1 mg∙L-1 (15 μM) at pH 7.0.

52
Environmental Science: Water Research & Technology Page 54 of 74

868 The horizontal dash lines represent the typical ClO2 treatment period, 30min andView Article Online
DOI: 10.1039/D0EW00231C

869 120min, respectively.

Environmental Science: Water Research & Technology Accepted Manuscript


870 As discussed above, ClO2 oxidation is selective and reacts with water

871 components mainly via electron transfer pathway. Despite of the limited formation of

872 halogenated DBPs, formation of chlorite and chlorate is the biggest obstacle for ClO2

873 application. The applied dose of ClO2 needs to be carefully controlled. One should

874 bear in mind that the reactions of moieties in waters is predicted based on laboratory

875 results with model compounds, which may not fully represent the situation in real
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

876 water treatment. Effects of competing reactions, variation of treatment conditions and

877 the level of pollutants in waters should be taken into consideration in practice. For

878 example, ClO2 is sometimes applied to remove Fe(II) and Mn(II) in water. Though

879 chloride ions, not chlorite, are the products from ClO2 reduction by Fe(II) or Mn(II)

880 based on the reaction mechanisms illustrated in section 3.1, ClO2 also competitively

881 reacts with NOM (e.g., phenolic moieties) and quickly reduce ClO2 to chlorite. In

882 addition, the presence of NOM may affect ECs’ degradation via two ways: competing

883 consumption of ClO2 and direct interaction with micropollutants. Accelerated

884 consumption of ClO2 in real water matrixes can reduce the removal efficiency of ECs.

885 For example, the kapp of ClO2 oxidation of ciprofloxacin is about 13.2 M-1 s-1 in

886 ultrapure water at pH 7.0, whereas the measured kapp in surface water matrixes

887 reduced to 6.4 M-1s-1.157 Moreover, complexes between NOM and ECs are likely

888 formed due to the electrostatic attraction between the negatively charged NOM

889 surface and the positively charged functional groups of ECs (e.g., amine groups). The

890 formation of NOM-EC complexes may hinder the contact between ClO2 and ECs,

891 leading to impaired removal performance for ECs.178 Still, the ClO2 dose needs to be

892 handled in great care. On the other perspective, phenolic moieties in NOM are found

53
Page 55 of 74 Environmental Science: Water Research & Technology

893 to be the dominant precursors of chlorite in the beginning stage of ClO2 reaction (< 5View Article Online
DOI: 10.1039/D0EW00231C

894 min).42 Transforming phenolic groups by some oxidation processes, such as

Environmental Science: Water Research & Technology Accepted Manuscript


895 permanganate, ozone or ferrate, ahead of ClO2 application may be helpful to reduce

896 chlorite formation, but this deserves future justification.

897 ClO2 is helpful in reducing some DBP precursors during subsequent chlorination

898 or chloramination in drinking water treatment. ClO2 is confirmed to be very effective

899 to reduce THM precursors, predominantly due to the deterioration of phenolic

900 precursors in water.98 Therefore, in waters suffering from meeting the THM
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

901 regulations, ClO2 pretreatment is a useful method. ClO2 pretreatment ahead of

902 chloramination is generally useful to reduce N-nitrosamine precursors mainly due to

903 the transformation of tertiary amines unless the water is impacted by hydrazine

904 pollutants.179 ClO2 also changes the formation of nitrogenous DBPs, such as

905 haloacetonitriles, halonitromethanes, and haloacetamides, but the changes don’t have

906 a consistent trend and vary with water matrix.180 Further characterization of

907 transformation products during ClO2 oxidation and their influences on the formation

908 of emerging disinfection byproducts, especially brominated, iodinated, or aromatic

909 DBPs, in subsequent chlorination and chloramination deserves further investigation.

910 From the view of safety of ClO2-treated waters, the main concern is the

911 formation of quinones that mainly derived from oxidation of phenolic compounds and

912 undesirable inorganic by-products chlorite and chlorate. In general, waters treated by

913 ClO2 alone exhibited much lower chronic mammalian cell cytotoxicity and

914 genotoxicity than chlorinated waters, according to data from European HIWATE

915 project.6 Cautious still need to be taken that when ClO2 is applied in combination with

916 other oxidants (e.g., chloramine) because a synergistic effect of organic DBP and

917 chlorite/chlorate could contributed to higher toxicity of treated samples.48 It was

54
Environmental Science: Water Research & Technology Page 56 of 74

918 reported that treatment of wastewaters in high ammonia levels by mixing oxidant ofView Article Online
DOI: 10.1039/D0EW00231C

919 ClO2 might exhibit higher cytotoxicity than using ClO2 alone.7 Thus, further

Environmental Science: Water Research & Technology Accepted Manuscript


920 investigations into featured organic products from ClO2 oxidation of different water

921 qualities are in urgent need to know better what are the suitable circumstances to

922 apply ClO2 treatment.

923 ClO2 has been studied for several decades from the aspects of reaction kinetics,

924 mechanisms, products and so on. However, the presence of HOCl in the reactions of

925 ClO2 with organic compounds and NOM is experimentally confirmed until recently.15,
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

926 41 Although the in situ formed HOCl is responsible for chlorinated by-products

927 formation during ClO2 oxidation, only about 7% of dosed ClO2 was involved with

928 chlorine incorporation to organic moieties overall.15 This explains the low yields of

929 chlorinated by-products during ClO2 oxidation. On the other hand, the presence of

930 trace amount of HOCl has been found to contribute to ECs’ removal.41 More studies

931 should be conducted to explore further the role of HOCl in ClO2 reaction system.

932

933 6. Conclusions

934 A varied group of inorganic and organic compounds can be transformed by ClO2.

935 ClO2 exhibits fast transformation rates of inorganic compounds including I-, CN-,

936 NO2-, SO32-, Fe(II) and Mn(II), but it does not react with ammonia or Br-. In typical

937 water treatment conditions, phenols, tertiary amines and thiols are reactive toward

938 ClO2, whereas alcohols, aldehydes, ketones, benzoquinone are less reactive.

939 Dissociated phenols and amines have faster reaction rate constants than the

940 un-dissociated forms. Quantitative relationships between structure and reaction rate

941 constants have been established for phenols and amines with electron-donating

942 substituents exhibiting faster reaction rates. For many such compounds and many

55
Page 57 of 74 Environmental Science: Water Research & Technology

943 emerging contaminants, mineralization is minimal and often little modification


DOI:of theView Article Online
10.1039/D0EW00231C

944 parent compounds is observed, with the formation of quinones, aldehydes and

Environmental Science: Water Research & Technology Accepted Manuscript


945 carboxylic acids. Electron transfer is the dominant reaction pathway, and ClO2 is first

946 converted to chlorite. The subsequent reaction steps are sometimes associated with the

947 release of HOCl or the formation of chloride from further reduction of chlorite by the

948 parent compounds. Except for chlorophenols, chlorinated organic products are

949 seldomly reported. Significant information on the reactivity of ClO2 with groups of

950 inorganic and organic compounds is available from the literatures. Such information
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

951 is essential for improving understanding of the transformation of emerging

952 contaminants and humic substances bearing diverse functional groups. It is also

953 helpful in understanding the formation trends of chlorite and the dominant organic

954 byproduct in ClO2-treated water and wastewater.

955

956 Acknowledgements:

957 This work was supported by the National Key Research and Development

958 Program of China (2017YFE0133200), National Natural Science Foundation of China

959 (21876210 and 21622706), Guangdong Provincial Science and Technology Planning

960 Project (2019A050503006 and 2017B020216005) and the China Postdoctoral Science

961 Foundation (2019M660671).

962

963 References

964 1. J. Huang, L. Wang, N. Ren, F. Ma and Juli, Disinfection effect of chlorine dioxide
965 on bacteria in water, Water Res., 1997, 31, 607–613.
966 2. H. Junli, W. Li, R. Nenqi, L. X. Li, S. R. Fun and Y. Guanle, Disinfection effect of
967 chlorine dioxide on viruses, algae and animal planktons in water, Water Res., 1997,
968 31, 455–460.

56
Environmental Science: Water Research & Technology Page 58 of 74

969 3. C. P. Chauret, C. Z. Radziminski, M. Lepuil, R. Creason and R. C. Andrews, View Article Online
DOI: 10.1039/D0EW00231C

970 Chlorine dioxide inactivation of Cryptosporidium parvum oocysts and bacterial spore

Environmental Science: Water Research & Technology Accepted Manuscript


971 indicators, Appl. Environ. Microb., 2001, 67, 2993–3001.
972 4. AWWA Water Quality Division Disinfection Systems Committee, Committee
973 report: Disinfection at large and medium-size systems, J. Am. Water Work. Assoc.,
974 2000, 92, 32–43.
975 5. J. Zhang, X. Zhang, J. Chen, C. Deng, N. Xu, W. Shi and P. Cheng, Highly
976 selective luminescent sensing of xylene isomers by a water stable Zn-organic
977 framework, Inorg. Chem. Commun., 2016, 69, 1–3.
978 6. C. H. Jeong, E. D. Wagner, V. R. Siebert, S. Anduri, S. D. Richardson, E. J.
979 Daiber, A. B. McKague, M. Kogevinas, C. M. Villanueva, E. H. Goslan, W. Luo, L.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

980 M. Isabelle, J. F. Pankow, R. Grazuleviciene, S. Cordier, S. C. Edwards, E. Righi, M.


981 J. Nieuwenhuijsen and M. J. Plewa, Occurrence and toxicity of disinfection
982 byproducts in European drinking waters in relation with the HIWATE epidemiology
983 study, Environ. Sci. Technol., 2012, 46, 12120–12128.
984 7. Y. Zhong, W. Gan, Y. Du, H. Huang, Q. Wu, Y. Xiang, C. Shang and X. Yang,
985 Disinfection byproducts and their toxicity in wastewater effluents treated by the
986 mixing oxidant of ClO2/Cl2, Water Res., 2019, 162, 471–481.
987 8. J. Hoigné and H. Bader, Kinetics of reactions of chlorine dioxide (OClO) in
988 water—I. Rate constants for inorganic and organic compounds, Water Res., 1994, 28,
989 45–55.
990 9. L. Hull, G. Davis, D. Rosenblatt, H. Williams and R. Weglein, Oxidations of
991 amines. III. Duality of mechanism in the reaction of amines with chlorine dioxide, J.
992 Am. Chem. Soc., 1967, 89, 1163–1170.
993 10. D. Rosenblatt, L. Hull, D. De Luca, G. Davis, R. Weglein and H. Williams,
994 Oxidations of amines. II. Substituent effects in chlorine dioxide oxidations, J. Am.
995 Chem. Soc., 1967, 89, 1158–1163.
996 11. C. Rav-Acha, E. Choshen and S. Sarel, Chlorine dioxide as an electron-transfer
997 oxidant of olefins, Helv. Chim. Acta, 1986, 69, 1728–1733.
998 12. C. Rav-Acha, Transformation of aqueous pollutants by chlorine dioxide:
999 Reactions, mechanisms and products. Quality and Treatment of Drinking Water II,
1000 Springer, 1998, 143–175.

57
Page 59 of 74 Environmental Science: Water Research & Technology

1001 13. I. Ganiev, Q. Timergazin, N. Kabalnova, V. Shereshovets and G. Tolstikov, View Article Online
DOI: 10.1039/D0EW00231C

1002 Reactions of chlorine dioxide with organic compounds, Eurasian Chem. Tech. J.,

Environmental Science: Water Research & Technology Accepted Manuscript


1003 2005, 7, 1–31.
1004 14. J. K. Leigh, J. Rajput and D. E. Richardson, Kinetics and mechanism of styrene
1005 epoxidation by chlorite: Role of chlorine dioxide, Inorg. Chem., 2014, 53, 6715–6727.
1006 15. V. Rougé, S. Allard, J. Croué and U. von Gunten, In situ formation of free
1007 chlorine during ClO2 treatment: Implications on the formation of disinfection
1008 byproducts, Environ. Sci. Technol., 2018, 52, 13421–13429.
1009 16. S. E. Barrett, S. W. Krasner and G. L. Amy, Natural organic matter and
1010 disinfection by-products: Characterization and control in drinking water—an
1011 overview, ACS Symp. Ser., 2000, 761, 2–14.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1012 17. D. J. Gates, G. Ziglio and K. Ozekin, State of the science of chlorine dioxide in
1013 drinking water, Water Research Foundation/Fondazione, 2009.
1014 18. X. Zhang, S. Echigo, R. A. Minear and M. J. Plewa, Characterization and
1015 comparison of disinfection by-products of four major disinfectants, ACS Symp. Ser.,
1016 2000, 761, 15–23.
1017 19. D. Couri, M. S. Abdel-Rahman and R. J. Bull, Toxicological effects of chlorine
1018 dioxide, chlorite and chlorate, Environ. Health Persp., 1982, 46, 13–17.
1019 20. USEPA, National primary drinking water regulations: Stage 2 disinfectants and
1020 disinfection byproducts rule, Fed. Reg., 2006, 71, 388–493.
1021 21. Ministry of Health, Standards for drinking water quality (GB5749-2006). In
1022 Ministry of Health of the People's Pepublic of China: Beijing, 2006.
1023 22. T. Schwartz, S. Hoffmann and U. Obst, Formation of natural biofilms during
1024 chlorine dioxide and UV disinfection in a public drinking water distribution system, J.
1025 Appl. Microbiol., 2003, 95, 591–601.
1026 23. J. De Koning, D. Bixio, A. Karabelas, M. Salgot and A. Schäfer, Characterisation
1027 and assessment of water treatment technologies for reuse, Desalination, 2008, 218,
1028 92–104.
1029 24. A. Bischoff, J. Fan, P. Cornel, M. Wagner and L. Ma, Disinfection of treated
1030 wastewater as an essential purification step for safe urban reuse: A comparative pilot
1031 study of UV-and ClO2-disinfection systems for urban reuse applications in China, J.
1032 Water Reuse Desal., 2013, 3, 325–335.
1033 25. R. K. Chhetri, A. Baun and H. R. Andersen, Algal toxicity of the alternative
1034 disinfectants performic acid (PFA), peracetic acid (PAA), chlorine dioxide (ClO2) and
58
Environmental Science: Water Research & Technology Page 60 of 74

1035 their by-products hydrogen peroxide (H2O2) and chlorite (ClO2−), Int. DOI:
J. 10.1039/D0EW00231C
Hyg.View Article Online
1036 Environ. Health, 2017, 220, 570–574.

Environmental Science: Water Research & Technology Accepted Manuscript


1037 26. G. Svecevičius, J. Šyvokienė, P. Stasiūnaitė and L. Mickėnienė, Acute and
1038 chronic toxicity of chlorine dioxide (ClO2) and chlorite (ClO2−) to rainbow trout
1039 (oncorhynchus mykiss) Environ. Sci. Pollut. Res., 2005, 12, 302–305.
1040 27. A. Dabrowska, J. Swietlik and J. Nawrocki, Formation of aldehydes upon ClO2
1041 disinfection, Water Res., 2003, 37, 1161–1169.
1042 28. J. Świetlik, U. Raczyk-Stanisławiak and J. Nawrocki, The influence of
1043 disinfection on aquatic biodegradable organic carbon formation, Water Res., 2009, 43,
1044 463–473.
1045 29. M. K. Ramseier, A. Peter, J. Traber and U. von Gunten, Formation of assimilable
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1046 organic carbon during oxidation of natural waters with ozone, chlorine dioxide,
1047 chlorine, permanganate, and ferrate, Water Res., 2011, 45, 2002–2010.
1048 30. F. Tian, Z. Qiang, C. Liu, T. Zhang and B. Dong, Kinetics and mechanism for
1049 methiocarb degradation by chlorine dioxide in aqueous solution, Chemosphere, 2010,
1050 79, 646–651.
1051 31. Y. Wang, H. Liu, G. Liu and Y. Xie, Oxidation of diclofenac by aqueous chlorine
1052 dioxide: Identification of major disinfection byproducts and toxicity evaluation, Sci.
1053 Total Environ., 2014, 473, 437–445.
1054 32. E. M. Aieta and J. D. Berg, A review of chlorine dioxide in drinking water
1055 treatment, J. Am. Water Work. Assoc., 1986, 78, 62–72.
1056 33. D. J. Gates, The chlorine dioxide handbook. American Water Works Association:
1057 Denver, 1998.
1058 34. G. V. Korshin, C. W. Li and M. M. Benjamin, Use of UV spectroscopy to study
1059 chlorination of natural organic matter, ACS sym. Ser., 1996, 649, 182–195.
1060 35. M. Deborde and U. von Gunten, Reactions of chlorine with inorganic and organic
1061 compounds during water treatment—Kinetics and mechanisms: A critical review,
1062 Water Res., 2008, 42, 13–51.
1063 36. E. Mvula and C. von Sonntag, Ozonolysis of phenols in aqueous solution, Org.
1064 Biomol. Chem., 2003, 1, 1749–1756.
1065 37. C. Von Sonntag and U. Von Gunten, Chemistry of ozone in water and wastewater
1066 treatment. IWA, 2012.
1067 38. I. Fábián and G. Gordon, The kinetics and mechanism of the chlorine
1068 dioxide−iodide ion reaction, Inorg. Chem., 1997, 36, 2494–2497.
59
Page 61 of 74 Environmental Science: Water Research & Technology

1069 39. J. A. Dean, Lange's handbook of chemistry, McGraw-Hill, 1999. View Article Online
DOI: 10.1039/D0EW00231C

1070 40. W. M. Haynes, CRC handbook of chemistry and physics, CRC press: Boca Raton,

Environmental Science: Water Research & Technology Accepted Manuscript


1071 FL, 2014.
1072 41. J. Terhalle, P. Kaiser, M. Jütte, J. Buss, S. Yasar, R. Marks, H. Uhlmann, T. C.
1073 Schmidt and H. V. Lutze, Chlorine dioxide—Pollutant transformation and formation
1074 of hypochlorous acid as a secondary oxidant, Environ. Sci. Technol., 2018, 52, 9964–
1075 9971.
1076 42. W. Gan, S. Huang, Y. Ge, T. Bond, P. Westerhoff, J. Zhai and X. Yang, Chlorite
1077 formation during ClO2 oxidation of model compounds having various functional
1078 groups and humic substances, Water Res., 2019, 159, 348–357.
1079 43. R. E. Huie and P. Neta, Kinetics of one-electron transfer reactions involving
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1080 chlorine dioxide and nitrogen dioxide, J. Phys. Chem., 1986, 90, 1193–1198.
1081 44. H. Fukutomi and G. Gordon, Kinetic study of the reaction between chlorine
1082 dioxide and potassium iodide in aqueous solution, J. Am. Chem. Soc., 1967, 89, 1362–
1083 1366.
1084 45. D. M. Kern and C. H. Kim, Iodine catalysis in the chlorite-iodide reaction, J. Am.
1085 Chem. Soc., 1965, 87, 5309–5313.
1086 46. I. Lengyel, J. Li, K. Kustin and I. Epstein, Rate constants for reactions between
1087 iodine- and chlorine-containing species: A detailed mechanism of the chlorine
1088 dioxide/chlorite-iodide reaction, J. Am. Chem. Soc., 1996, 118, 3708–3719.
1089 47. G. Hua and D. A. Reckhow, Comparison of disinfection byproduct formation
1090 from chlorine and alternative disinfectants, Water Res., 2007, 41, 1667–1678.
1091 48. J. Han and X. Zhang, Evaluating the comparative toxicity of DBP mixtures from
1092 different disinfection scenarios: A new approach by combining freeze-drying or
1093 rotoevaporation with a marine polychaete bioassay, Environ. Sci. Technol., 2018, 52,
1094 10552–10561.
1095 49. I. Lengyel, I. R. Epstein and K. Kustin, Kinetics of iodine hydrolysis, Inorg.
1096 Chem., 1993, 32, 5880–5882.
1097 50. I. Lengyel, G. Rabai and I. R. Epstein, Experimental and modeling study of
1098 oscillations in the chlorine dioxide-iodine-malonic acid reaction, J. Am. Chem. Soc.,
1099 1990, 112, 9104–9110.
1100 51. E. M. Smith, M. J. Plewa, C. L. Lindell, S. D. Richardson and W. A. Mitch,
1101 Comparison of byproduct formation in waters treated with chlorine and iodine:
1102 relevance to point-of-use treatment, Environ. Sci. Technol., 2010, 44, 8446–8452.
60
Environmental Science: Water Research & Technology Page 62 of 74

1103 52. T. Ye, B. Xu, Y. L. Lin, C. Y. Hu, L. Lin, T. Y. Zhang and N. Y. Gao, Formation View Article Online
DOI: 10.1039/D0EW00231C

1104 of iodinated disinfection by-products during oxidation of iodide-containing waters

Environmental Science: Water Research & Technology Accepted Manuscript


1105 with chlorine dioxide, Water Res., 2013, 47, 3006–3014.
1106 53. T. Y. Zhang, B. Xu, C. Y. Hu, Y. L. Lin, L. Lin, T. Ye and F. X. Tian, A
1107 comparison of iodinated trihalomethane formation from chlorine, chlorine dioxide and
1108 potassium permanganate oxidation processes, Water Res., 2015, 68, 394–403.
1109 54. D. M. Stanbury, R. Martinez, E. Tseng and C. E. Miller, Slow electron transfer
1110 between main-group species: Oxidation of nitrite by chlorine dioxide, Inorg. Chem.,
1111 1988, 27, 4277–4280.
1112 55. J. R. Parga and D. L. Cocke, Oxidation of cyanide in a hydrocyclone reactor by
1113 chlorine dioxide, Desalination, 2001, 140, 289–296.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1114 56. J. R. Parga, S. S. Shukla and F. R. Carrillo-Pedroza, Destruction of cyanide waste


1115 solutions using chlorine dioxide, ozone and titania sol, Waste Manage., 2003, 23,
1116 183–191.
1117 57. J. Halperin and H. Taube, The transfer of oxygen atoms in oxidation—reduction
1118 reactions. III. The reaction of halogenates with sulfite in aqueous solution, J. Am.
1119 Chem. Soc., 1952, 74, 375–380.
1120 58. K. D. Fogelman, D. M. Walker and D. W. Margerum, Nonmetal redox kinetics:
1121 Hypochlorite and hypochlorous acid reactions with sulfite, Inorg. Chem., 1989, 28,
1122 986–993.
1123 59. G. Merenyi, J. Lind and X. Shen, Electron transfer from indoles, phenol, and
1124 sulfite (SO32-) to chlorine dioxide (ClO2), J. Phys. Chem., 1988, 92, 134–137.
1125 60. A. K. Horváth and I. Nagypál, Kinetics and mechanism of the oxidation of sulfite
1126 by chlorine dioxide in a slightly acidic medium, J. Phys. Chem. A, 2006, 110, 4753–
1127 4758.
1128 61. K. Suzuki and G. Gordon, Stoichiometry and kinetics of the reaction between
1129 chlorine dioxide and sulfur(IV) in basic solutions, Inorg. Chem., 1978, 17, 3115–
1130 3118.
1131 62. G. Csekö, C. Pan, Q. Gao and A. K. Horváth, Kinetics of the two-stage oxidation
1132 of sulfide by chlorine dioxide, Inorg. Chem., 2018, 57, 10189–10198.
1133 63. Wong and J. M., Chlorination-filtration for iron and manganese removal, J. Am.
1134 Water Work. Assoc., 1984, 76, 76–79.

61
Page 63 of 74 Environmental Science: Water Research & Technology

1135 64. W. R. Knocke, J. E. Van Benschoten, M. J. Kearney, A. W. Soborski and


DOI: D. A.View Article Online
10.1039/D0EW00231C

1136 Reckhow, Kinetics of manganese and iron oxidation by potassium permanganate and

Environmental Science: Water Research & Technology Accepted Manuscript


1137 chlorine dioxide, J. Am. Water Work. Assoc., 1991, 83, 80–87.
1138 65. J. E. Van Benschoten, W. Lin and W. R. Knocke, Kinetic modeling of
1139 manganese(II) oxidation by chlorine dioxide and potassium permanganate, Environ.
1140 Sci. Technol., 1992, 26, 1327–1333.
1141 66. L. Wang, I. N. Odeh and D. W. Margerum, Chlorine dioxide reduction by
1142 aqueous iron(II) through outer-sphere and inner-sphere electron-transfer pathways,
1143 Inorg. Chem., 2004, 43, 7545–7551.
1144 67. R. Henderson, K. Carlson and D. Gregory, The impact of ferrous ion reduction of
1145 chlorite ion on drinking water process performance, Water Res., 2001, 35, 4464–4473.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1146 68. L. Chen, J. J. Zhang and X. L. Zheng, Coupling technique for deep removal of
1147 manganese and iron from potable water, Environ. Eng. Sci., 2016, 33, 261–269.
1148 69. C. Hall, E. R. LaBerge and S. J. Duranceau, Comparing potassium permanganate,
1149 chlorine dioxide, and chlorine oxidation for manganese control of a volcanic island
1150 surface water treated with a conventional coagulation, sedimentation, and filtration
1151 process, Desalin. Water Treat., 2016, 57, 14355–14363.
1152 70. G. K. Khadse, P. M. Patni and P. K. Labhasetwar, Removal of iron and
1153 manganese from drinking water supply, Sustain. Water Resour. Manag., 2015, 1,
1154 157–165.
1155 71. G. H. Hurst and W. R. Knocke, Evaluating ferrous iron for chlorite ion removal,
1156 J. Am. Water Work. Assoc., 1997, 89, 98–105.
1157 72. M. G. Ondrus and G. Gordon, Oxidation of hexaaquoiron (II) by chlorine (III) in
1158 aqueous solution, Inorg. Chem., 1972, 11, 985–989.
1159 73. W. Cullen and K. Reimer, Arsenic speciation in the environment, Chem. Rev.,
1160 1988, 89, 713–764.
1161 74. G. Ghurye and D. A. Clifford, Laboratory study on the oxidation of arsenic III to
1162 arsenic V, In Office of Research and Development, USEPA, 2001.
1163 75. G. Ghurye and D. Clifford, As(III) oxidation using chemical and solid-phase
1164 oxidants, J. Am. Water Work. Assoc., 2004, 96, 84–96.
1165 76. O. X. Leupin, S. J. Hug and A. B. M. Badruzzaman, Arsenic removal from
1166 Bangladesh tube well water with filter columns containing zerovalent iron filings and
1167 sand, Environ. Sci. Technol., 2005, 39, 8032–8037.

62
Environmental Science: Water Research & Technology Page 64 of 74

1168 77. S. Sorlini and F. Gialdini, Conventional oxidation treatments for the removal ofView Article Online
DOI: 10.1039/D0EW00231C

1169 arsenic with chlorine dioxide, hypochlorite, potassium permanganate and

Environmental Science: Water Research & Technology Accepted Manuscript


1170 monochloramine, Water Res., 2010, 44, 5653–5659.
1171 78. V. K. Sharma, P. K. Dutta and A. K. Ray, Review of kinetics of chemical and
1172 photocatalytical oxidation of arsenic(III) as influenced by pH, J. Environ. Sci. Health,
1173 Part A: Toxic/Hazard. Subst. Environ. Eng., 2007, 42, 997–1004.
1174 79. J. Han, X. Zhang, J. Liu, X. Zhu and T. Gong, Characterization of halogenated
1175 DBPs and identification of new DBPs trihalomethanols in chlorine dioxide treated
1176 drinking water with multiple extractions, J. Environ. Sci., 2017, 58, 83–92.
1177 80. R. J. Bull and F. C. Kopfler, Health effects of disinfectants and disinfection
1178 by-products, Amer Water Works Assn, 1991.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1179 81. J. E. Wajon, D. H. Rosenblatt and E. P. Burrows, Oxidation of phenol and


1180 hydroquinone by chlorine dioxide, Environ. Sci. Technol., 1982, 16, 396–402.
1181 82. B. O. Lindgren, C. M. Svahn and G. Widmark, Chlorine dioxide oxidation of
1182 cyclohexene, Acta. Chem. Scand., 1965, 19, 7–13.
1183 83. I. V. Loginova, I. Y. Chukicheva and A. V. Kuchin, Reaction of styrene with
1184 chlorine dioxide, Russ. J. Gen. Chem., 2018, 88, 825–828.
1185 84. C. Rav-Acha, E. Choshen and S. Sarel, Chlorine dioxide as an electron-transfer
1186 oxidant of olefins, Helv. Chim. Acta, 1986, 69, 1728–1733.
1187 85. K. A. Thorn, D. W. Folan and P. MacCarthy, Characterization of the
1188 International Humic Substances Society standard and reference fulvic and humic
1189 acids by solution state carbon-13 (13C) and hydrogen-1 (1H) nuclear magnetic
1190 resonance spectrometry, US Department of the Interior, US Geological Survey, 1991.
1191 86. A. Tor, M. Aydin, Y. Çengeloglu and S. Özcan, Using n-alkanes for
1192 identification of oils in domestic wastewaters, Environ. Technol., 2005, 26, 1289–
1193 1296.
1194 87. M. B. Yunker, S. Backus, E. G. Pannatier, D. Jeffries and R. Macdonald, Sources
1195 and significance of alkane and PAH hydrocarbons in Canadian arctic rivers, Estuar.
1196 Coast. Shelf S., 2002, 55, 1–31.
1197 88. C. Rav-Acha and R. Blits, The different reaction mechanisms by which chlorine
1198 and chlorine dioxide react with polycyclic aromatic hydrocarbons (PAH) in water,
1199 Water Res., 1985, 19, 1273–1281.

63
Page 65 of 74 Environmental Science: Water Research & Technology

1200 89. E. Choshen, R. Elits and C. Rav-Acha, The formation of cation-radicals


DOI:by theView Article Online
10.1039/D0EW00231C

1201 action of chlorine dioxide on ρ-substituted styrenes usd other alkenes, Tetrahedron

Environmental Science: Water Research & Technology Accepted Manuscript


1202 Lett., 1986, 27, 5989–5992.
1203 90. P. G. Tratnyek and J. Hoigné, Kinetics of reactions of chlorine dioxide (OCIO) in
1204 water—II. Quantitative structure-activity relationships for phenolic compounds,
1205 Water Res., 1994, 28, 57–66.
1206 91. V. Vaida and J. D. Simon, The photoreactivity of chlorine dioxide, Science, 1995,
1207 268, 1443–1448.
1208 92. J. Hoigne and H. Bader, Kinetics of reactions of chlorine dioxide with
1209 representative micropollutants in water, Vom Wasser, 1982, 59, 253–267.
1210 93. J. R. Kastner, K. C. Das, C. Hu and R. McClendon, Effect of pH and temperature
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1211 on the kinetics of odor oxidation using chlorine dioxide, J. Air Waste Manage., 2003,
1212 53, 1218–1224.
1213 94. E. Dalcanale and F. Montanari, Selective oxidation of aldehydes to carboxylic
1214 acids with sodium chlorite-hydrogen peroxide, J. Org. Chem., 1986, 51, 567–569.
1215 95. H. F. Launer and Y. Tomimatsu, Stoichiometry of chlorite-aldehyde reactions.
1216 Analytical procedures, Anal. Chem., 1959, 31, 1385–1390.
1217 96. W. Gan, Y. Ge, H. Zhu, H. Huang and X. Yang, ClO2 pre-oxidation changes the
1218 yields and formation pathways of chloroform and chloral hydrate from phenolic
1219 precursors during chlorination, Water Res., 2019, 148, 250–260.
1220 97. J. D. Ritchie and E. M. Perdue, Proton-binding study of standard and reference
1221 fulvic acids, humic acids, and natural organic matter, Geochim. Cosmochim. Ac.,
1222 2003, 67, 85–96.
1223 98. J. Wenk, M. Aeschbacher, E. Salhi, S. Canonica, U. von Gunten and M. Sander,
1224 Chemical oxidation of dissolved organic matter by chlorine dioxide, chlorine, and
1225 ozone: effects on its optical and antioxidant properties, Environ. Sci. Technol., 2013,
1226 47, 11147–11156.
1227 99. Y. Lee and U. Von Gunten, Quantitative structure–activity relationships (QSARs)
1228 for the transformation of organic micropollutants during oxidative water treatment,
1229 Water Res., 2012, 46, 6177–6195.
1230 100. D. J. V. A. dos Santos, A. S. Newton, R. Bernardino and R. C. Guedes,
1231 Substituent effects on O–H and S–H bond dissociation enthalpies of disubstituted
1232 phenols and thiophenols, Int. J. Quantum. Chem., 2008, 108, 754–761.

64
Environmental Science: Water Research & Technology Page 66 of 74

1233 101. Z. B. Alfassi, R. E. Huie and P. Neta, Substituent effects on rates of one-electron View Article Online
DOI: 10.1039/D0EW00231C

1234 oxidation of phenols by the radicals chlorine dioxide, nitrogen dioxide, and

Environmental Science: Water Research & Technology Accepted Manuscript


1235 trioxosulfate (1-), J. Phys. Chem., 1986, 90, 4156–4158.
1236 102. Y. Ni, X. Shen and A. R. P. van Heiningen, Studies on the reactions of phenolic
1237 and non-phenolic lignin model compounds with chlorine dioxide, J. Wood Chem.
1238 Technol., 1994, 14, 243–262.
1239 103. J. Trevors and J. Basaraba, Toxicity of benzoquinone and hydroquinone in
1240 short-term bacterial bioassays, B. Environ. Contam. Tox., 1980, 25, 672–675.
1241 104. H. Du, J. Li, B. Moe, C. F. McGuigan, S. Shen and X. Li, Cytotoxicity and
1242 oxidative damage induced by halobenzoquinones to T24 bladder cancer cells,
1243 Environ. Sci. Technol., 2013, 47, 2823–2830.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1244 105. J. L. Bolton, M. A. Trush, T. M. Penning, G. Dryhurst and T. J. Monks, Role of


1245 quinones in toxicology, Chem. Res. Toxicol., 2000, 13, 135–160.
1246 106. J. Li, W. Wang, B. Moe, H. Wang and X. Li, Chemical and toxicological
1247 characterization of halobenzoquinones, an emerging class of disinfection byproducts,
1248 Chem. Res. Toxicol., 2015, 28, 306–318.
1249 107. I. V. Loginova, K. S. Rodygin, S. A. Rubtsova, P. A. Slepukhin, A. V. Kuchin
1250 and V. A. Polukeev, Oxidation of polyfunctional sulfides with chlorine dioxide, Russ.
1251 J. Org. Chem., 2011, 47, 124–130.
1252 108. I. V. Loginova, S. A. Rubtsova and A. V. Kuchin, Oxidation by chlorine dioxide
1253 of methionine and cysteine derivatives to sulfoxides, Chem. Nat. Compd., 2008, 44,
1254 752–754.
1255 109. A. Kutchin, S. Rubtsova and I. Loginova, Reactions of chlorine dioxide with
1256 organic compounds. Selective oxidation of sulfides to sulfoxides by chlorine dioxide,
1257 Russ. Chem. B, 2001, 50, 432–435.
1258 110. A. V. Kutchin, S. A. Rubtsova, D. V. Sudarikov and M. Y. Demakova,
1259 Chlorine dioxide in chemo- and stereoselective oxidation of sulfides, Russ. Chem. B,
1260 2013, 62, 1–5.
1261 111. A. Lopez, G. Mascolo, G. Tiravanti and R. Passino, Degradation of
1262 herbicides (ametryn and isoproturon) during water disinfection by means of two
1263 oxidants (hypochlorite and chlorine dioxide), Water Sci. Technol., 35, 129–136.
1264 112. M. Z. Yakupov, V. V. Shereshovets, U. B. Imashev and F. R. Ismagilov,
1265 Liquid-phase oxidation of thiols with chlorine dioxide, Russ. Chem. B, 2001, 50,
1266 2352–2355.
65
Page 67 of 74 Environmental Science: Water Research & Technology

1267 113. J. Darkwa, R. Olojo, E. Chikwana and R. H. Simoyi, Antioxidant chemistry: View Article Online
DOI: 10.1039/D0EW00231C

1268 Oxidation of L-cysteine and its metabolites by chlorite and chlorine dioxide, J. Phys.

Environmental Science: Water Research & Technology Accepted Manuscript


1269 Chem. A, 2004, 108, 5576–5587.
1270 114. A. Ison, I. N. Odeh and D. W. Margerum, Kinetics and mechanisms of
1271 chlorine dioxide and chlorite oxidations of cysteine and glutathione, Inorg. Chem.,
1272 2006, 45, 8768–8775.
1273 115. G. Rábai, R. T. Wang and K. Kustin, Kinetics and mechanism of the oxidation of
1274 thiourea by chlorine dioxide, Int. J. Chem. Kinet., 1993, 25, 53–62.
1275 116. C. Joe-Wong, E. Shoenfelt, E. J. Hauser, N. Crompton and S. C. B. Myneni,
1276 Estimation of reactive thiol concentrations in dissolved organic matter and bacterial
1277 eell membranes in aquatic systems, Environ. Sci. Technol., 2012, 46, 9854–9861.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1278 117. P. Westerhoff and H. Mash, Dissolved organic nitrogen in drinking water
1279 supplies: a review, J. Water Supply Res. T., 2002, 51, 415–448.
1280 118. M. Selbes, D. Kim and T. Karanfil, The effect of pre-oxidation on NDMA
1281 formation and the influence of pH, Water Res., 2014, 66, 169–179
1282 119. C. Lee, C. Schmidt, J. Yoon and U. von Gunten, Oxidation of
1283 N-nitrosodimethylamine (NDMA) precursors with ozone and chlorine dioxide:
1284 Kinetics and effect on NDMA formation potential, Environ. Sci. Technol., 2007, 41,
1285 2056–2063.
1286 120. D. H. Rosenblatt, A. J. Hayes Jr, B. L. Harrison, R. A. Streaty and K. A. Moore,
1287 The reaction of chlorine dioxide with triethylamine in aqueous solution, J. Org.
1288 Chem., 1963, 28, 2790–2794.
1289 121. T. E. Eriksen, J. Lind and G. Merényi, Oxidation of luminol by chlorine dioxide.
1290 Formation of 5-aminophthalazine-1, 4-dione, J. Chem. Soc., Faraday Trans. 1, 1981,
1291 77, 2125–2135.
1292 122. P. Neta, R. E. Huie and A. B. Ross, Rate constants for reactions of inorganic
1293 radicals in aqueous solution, J. Phys. Chem. Ref. Data, 1988, 17, 1027–1284.
1294 123. G. Merenyi, J. Lind and T. E. Erikson, Nucleophilic addition to diazaquinones.
1295 Formation and breakdown of tetrahedral intermediates in relation to luminol
1296 chemiluminescence, J. Am. Chem. Soc., 1986, 108, 7716–7726.
1297 124. L. Abia, A. X.L., M. Canle, M. V. Garcia and J. A. Santaballa, Oxidation of
1298 aliphatic amines by aqueous chlorine, Tetrahedron, 1998, 54, 521–530.

66
Environmental Science: Water Research & Technology Page 68 of 74

1299 125. M. Selbes, D. Kim, N. Ates and T. Karanfil, The roles of tertiary amine structure, View Article Online
DOI: 10.1039/D0EW00231C

1300 background organic matter and chloramine species on NDMA formation, Water Res.,

Environmental Science: Water Research & Technology Accepted Manuscript


1301 2013, 47, 945–953.
1302 126. Y. Y. Zhao, J. M. Boyd, M. Woodbeck, R. C. Andrews, F. Qin, S. E. Hrudey and
1303 X. F. Li, Formation of N-nitrosamines from eleven disinfection treatments of seven
1304 different surface waters, Environ. Sci. Technol., 2008, 42, 4857–4862.
1305 127. A. D. Shah, S. W. Krasner, C. F. T. Lee, U. von Gunten and W. A. Mitch,
1306 Trade-offs in disinfection byproduct formation associated with precursor preoxidation
1307 for control of N-nitrosodimethylamine formation, Environ. Sci. Technol., 2012, 46,
1308 4809–4818.
1309 128. X. Lin, J. Zhang, X. Luo, C. Zhang and Y. Zhou, Removal of aniline using lignin
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1310 grafted acrylic acid from aqueous solution, Chem. Eng. J., 2011, 172, 856–863.
1311 129. S. D. Richardson and S. Y. Kimura, Water analysis: Emerging contaminants and
1312 current issues, Anal. Chem., 2016, 88, 546–582.
1313 130. C. A. H. Aguilar, J. Narayanan, N. Singh and P. Thangarasu, Kinetics and
1314 mechanism for the oxidation of anilines by ClO2: a combined experimental and
1315 computational study, J. Phys. Org. Chem., 2014, 27, 440–449.
1316 131. H. K. Tan, W. B. Wheeler and C. I. Wei, Reaction of chlorine dioxide with
1317 amino acids and peptides: kinetics and mutagenicity studies, Mutat. Res./Genet.
1318 Toxicol., 1987, 188, 259–266.
1319 132. M. J. Napolitano, B. J. Green, J. S. Nicoson and D. W. Margerum, Chlorine
1320 dioxide oxidations of tyrosine, N-acetyltyrosine, and dopa, Chem. Res. Toxicol., 2005,
1321 18, 501–508.
1322 133. D. J. Stewart, M. J. Napolitano, E. V. Bakhmutova-Albert and D. W. Margerum,
1323 Kinetics and mechanisms of chlorine dioxide oxidation of tryptophan, Inorg. Chem.,
1324 2008, 47, 1639–1647.
1325 134. N. Ogata, Denaturation of protein by chlorine dioxide: Oxidative modification of
1326 tryptophan and tyrosine residues, Biochemstry, 2007, 46, 4898–4911.
1327 135. X. Shen, J. Lind and G. Merenyi, One-electron oxidation of indoles and
1328 acid-base properties of the indolyl radicals, J. Phys. Chem., 1987, 91, 4403–4406.
1329 136. T. Aniszewski, Alkaloids-Secrets of Life: Aklaloid Chemistry, Biological
1330 Significance, Applications and Ecological Role, Elsevier, 2007.
1331 137. S. Navalon, M. Alvaro and H. Garcia, Chlorine dioxide reaction with selected
1332 amino acids in water, J. Hazard. Mater., 2009, 164, 1089–1097.
67
Page 69 of 74 Environmental Science: Water Research & Technology

1333 138. Y. Yang, Y. S. Ok, K. H. Kim, E. E. Kwon and Y. F. Tsang, Occurrences andView Article Online
DOI: 10.1039/D0EW00231C

1334 removal of pharmaceuticals and personal care products (PPCPs) in drinking water and

Environmental Science: Water Research & Technology Accepted Manuscript


1335 water/sewage treatment plants: A review, Sci. Total. Environ., 2017, 596–597, 303–
1336 320.
1337 139. S. D. Kim, J. Cho, I. S. Kim, B. J. Vanderford and S. A. Snyder, Occurrence and
1338 removal of pharmaceuticals and endocrine disruptors in South Korean surface,
1339 drinking, and waste waters, Water Res., 2007, 41, 1013–1021.
1340 140. D. J. Lapworth, N. Baran, M. E. Stuart and R. S. Ward, Emerging organic
1341 contaminants in groundwater: A review of sources, fate and occurrence, Environ.
1342 Pollut., 2012, 163, 287–303.
1343 141. V. de Jesus Gaffney, C. M. Almeida, A. Rodrigues, E. Ferreira, M. J. Benoliel
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1344 and V. V. Cardoso, Occurrence of pharmaceuticals in a water supply system and


1345 related human health risk assessment, Water Res., 2015, 72, 199–208.
1346 142. X. Yang, R. C. Flowers, H. S. Weinberg and P. C. Singer, Occurrence and
1347 removal of pharmaceuticals and personal care products (PPCPs) in an advanced
1348 wastewater reclamation plant, Water Res., 2011, 45, 5218–5228.
1349 143. M. J. Benotti, R. A. Trenholm, B. J. Vanderford, J. C. Holady, B. D. Stanford
1350 and S. A. Snyder, Pharmaceuticals and endocrine disrupting compounds in US
1351 drinking water, Environ. Sci. Technol., 2009, 43, 597–603.
1352 144. R. P. Schwarzenbach, B. I. Escher, K. Fenner, T. B. Hofstetter, C. A. Johnson, U.
1353 Von Gunten and B. Wehrli, The challenge of micropollutants in aquatic systems,
1354 Science, 2006, 313, 1072–1077.
1355 145. M. G. Pintado-Herrera, C. Wang, J. Lu, Y. Chang, W. Chen, X. Li and P. A.
1356 Lara-Martín, Distribution, mass inventories, and ecological risk assessment of legacy
1357 and emerging contaminants in sediments from the Pearl River Estuary in China, J.
1358 Hazard. Mater., 2017, 323, 128–138.
1359 146. L. C. Pereira, A. O. de Souza, M. F. F. Bernardes, M. Pazin, M. J. Tasso, P. H.
1360 Pereira and D. J. Dorta, A perspective on the potential risks of emerging contaminants
1361 to human and environmental health, Environ. Sci. Pollut. Res., 2015, 22, 13800–
1362 13823.
1363 147. M. Stuart, D. Lapworth, E. Crane and A. Hart, Review of risk from potential
1364 emerging contaminants in UK groundwater, Sci. Total Environ., 2012, 416, 1–21.
1365 148. Y. Ma, C. A. Wilson, J. T. Novak, R. Riffat, S. Aynur, S. Murthy and A. Pruden,
1366 Effect of various sludge digestion conditions on sulfonamide, macrolide, and
68
Environmental Science: Water Research & Technology Page 70 of 74

1367 tetracycline resistance genes and class I integrons, Environ. Sci. Technol., 2011, 45,View Article Online
DOI: 10.1039/D0EW00231C

1368 7855–7861.

Environmental Science: Water Research & Technology Accepted Manuscript


1369 149. Y. Luo, F. Yang, J. Mathieu, D. Mao, Q. Wang and P. J. J. Alvarez, Proliferation
1370 of multidrug-resistant new delhi metallo-β-lactamase genes in municipal wastewater
1371 treatment plants in Northern China, Environ. Sci. Technol. Lett., 2014, 1, 26–30.
1372 150. N. Czekalski, T. Berthold, S. Caucci, A. Egli and H. Bürgmann, Increased levels
1373 of multiresistant bacteria and resistance genes after wastewater treatment and their
1374 dissemination into lake geneva, Switzerland, Front. Microb., 2012, 3, 106–106.
1375 151. R. Pei, S. C. Kim, K. H. Carlson and A. Pruden, Effect of river landscape on the
1376 sediment concentrations of antibiotics and corresponding antibiotic resistance genes
1377 (ARG), Water Res., 2006, 40, 2427–2435.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1378 152. C. Xi, Y. Zhang, C. F. Marrs, W. Ye, C. Simon, B. Foxman and J. Nriagu,
1379 Prevalence of antibiotic resistance in drinking water treatment and distribution
1380 systems, Appl. Environ. Microb., 2009, 75, 5714–5718.
1381 153. L. Ma, B. Li, X. T. Jiang, Y. L. Wang, Y. Xia, A. D. Li and T. Zhang, Catalogue
1382 of antibiotic resistome and host-tracking in drinking water deciphered by a large scale
1383 survey, Microbiome, 2017, 5, 154–166.
1384 154. M. M. Huber, S. Korhonen, T. A. Ternes and U. V. Gunten, Oxidation of
1385 pharmaceuticals during water treatment with chlorine dioxide, Water Res., 2005, 39,
1386 3607–3617.
1387 155. S. Navalon, M. Alvaro and H. Garcia, Reaction of chlorine dioxide with
1388 emergent water pollutants: product study of the reaction of three beta-lactam
1389 antibiotics with ClO2, Water Res., 2008, 42, 1935–1942.
1390 156. W. Ben, Y. Shi, W. Li, Y. Zhang and Z. Qiang, Oxidation of sulfonamide
1391 antibiotics by chlorine dioxide in water: Kinetics and reaction pathways, Chem. Eng.
1392 J., 2017, 327, 743–750.
1393 157. P. Wang, Y. He and C. Huang, Oxidation of fluoroquinolone antibiotics and
1394 structurally related amines by chlorine dioxide: Reaction kinetics, product and
1395 pathway evaluation, Water Res., 2010, 44, 5989–5998.
1396 158. F. Tian, B. Xu, T. Zhang and N. Gao, Degradation of phenylurea herbicides by
1397 chlorine dioxide and formation of disinfection by-products during subsequent chlor
1398 (am) ination, Chem. Eng. J., 2014, 258, 210–217.
1399 159. T. P. Kull, P. H. Backlund, K. M. Karlsson and J. A. Meriluoto, Oxidation of the
1400 cyanobacterial hepatotoxin microcystin-LR by chlorine dioxide: reaction kinetics,
69
Page 71 of 74 Environmental Science: Water Research & Technology

1401 characterization, and toxicity of reaction products, Environ. Sci. Technol., 2004, 38,View Article Online
DOI: 10.1039/D0EW00231C

1402 6025–6031.

Environmental Science: Water Research & Technology Accepted Manuscript


1403 160. Y. Chen, H. Li, Z. Wang, H. Li, T. Tao and Y. Zuo, Photodegradation of selected
1404 blockers in aqueous fulvic acid solutions: Kinetics, mechanism, and product analysis,
1405 Water Res., 2012, 46, 2965–2972.
1406 161. P. Wang, Y. He and C. H. Huang, Oxidation of antibiotic agent trimethoprim by
1407 chlorine dioxide: Reaction kinetics and pathways, J. Environ. Eng., 2012, 138, 360–
1408 366.
1409 162. X. H. Jia, L. Feng, Y. Z. Liu and L. Q. Zhang, Oxidation of antipyrine by
1410 chlorine dioxide: Reaction kinetics and degradation pathway, Chem. Eng. J., 309,
1411 646–654.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1412 163. E. Rodríguez, G. D. Onstad, T. P. Kull, J. S. Metcalf, J. L. Acero and U. von


1413 Gunten, Oxidative elimination of cyanotoxins: Comparison of ozone, chlorine,
1414 chlorine dioxide and permanganate, Water Res., 2007, 41, 3381–3393.
1415 164. H. Roques, Chemical water treatment: principles and practice, VCH Publishers.
1416 Inc., New York, 1996, 620.
1417 165. M. Wu, J. Liu, S. You, L. Wang, J. Huang and Y. Tian, Effects and kinetics of
1418 chlorine dioxide for removal of benzo[α]pyrene in water, Environ. Eng. Sci., 29, 133–
1419 138.
1420 166. C. Rav-Acha, The reactions of chlorine dioxide with aquatic organic materials
1421 and their health effects, Water Res., 1984, 18, 1329–1341.
1422 167. S. Willach, H. V. Lutze, K. Eckey, K. Loeppenberg, M. Lüling, J. Terhalle, J.
1423 Wolbert, M. A. Jochmann, U. Karst and T. C. Schmidt, Degradation of
1424 sulfamethoxazole using ozone and chlorine dioxide-compound-specific stable isotope
1425 analysis, transformation product analysis and mechanistic aspects, Water Res., 2017,
1426 122, 280–289.
1427 168. A. G. Trovó, R. F. Nogueira, A. Agüera, A. R. Fernandez-Alba, C. Sirtori and S.
1428 Malato, Degradation of sulfamethoxazole in water by solar photo-Fenton. Chemical
1429 and toxicological evaluation, Water Res., 2009, 43, 3922–3931.
1430 169. Y. Wang, H. Liu, Y. Xie, T. Ni and G. Liu, Oxidative removal of diclofenac by
1431 chlorine dioxide: Reaction kinetics and mechanism, Chem. Eng. J., 2015, 279, 409–
1432 415.
1433 170. S. Othman, V. Mansuy-Mouries, C. Bensoussan, P. Battioni and D. Mansuy,
1434 Hydroxylation of diclofenac: An illustration of the complementary roles of
70
Environmental Science: Water Research & Technology Page 72 of 74

1435 biomimetic metalloporphyrin catalysts and yeasts expressing human cytochromes View Article Online
DOI: 10.1039/D0EW00231C

1436 P450 in drug metabolism studies, C.R. Acad. Sci., Ser. IIC-Chem., 2000, 3, 751–755.

Environmental Science: Water Research & Technology Accepted Manuscript


1437 171. M. C. Dodd, A. D. Shah, U. von Gunten and C. Huang, Interactions of
1438 fluoroquinolone antibacterial agents with aqueous chlorine: reaction kinetics,
1439 mechanisms, and transformation pathways, Environ. Sci. Technol., 2005, 39, 7065–
1440 7076.
1441 172. W. H. D. Jr, L. A. Hull and D. H. Rosenblatt, Oxidations of amines. IV.
1442 Oxidative fragmentation, J. Org. Chem., 1967, 32, 3783–3787.
1443 173. G. Mascolo, A. Lopez, R. Passino, G. Ricco and G. Tiravanti, Degradation of
1444 sulphur containing s-triazines during water chlorination, Water Res., 1994, 28, 2499–
1445 2506.
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

1446 174. H. He, P. Zhou, K. K. Shimabuku, X. Fang, S. Li, Y. Lee and M. C. Dodd,
1447 Degradation and deactivation of bacterial antibiotic resistance genes during exposure
1448 to free chlorine, monochloramine, chlorine dioxide, ozone, ultraviolet light, and
1449 hydroxyl radical, Environ. Sci. Technol., 2019, 53, 2013–2026.
1450 175. M. J. Napolitano, D. J. Stewart and D. W. Margerum, Chlorine dioxide oxidation
1451 of guanosine 5’-monophosphate, Chem. Res. Toxicol., 2006, 19, 1451–1458.
1452 176. G. Hey, R. Grabic, A. Ledin, J. la Cour Jansen and H. R. Andersen, Oxidation of
1453 pharmaceuticals by chlorine dioxide in biologically treated wastewater, Chem. Eng.
1454 J., 2012, 185–186, 236–242.
1455 177. Y. Lee and U. von Gunten, Oxidative transformation of micropollutants during
1456 municipal wastewater treatment: Comparison of kinetic aspects of selective (chlorine,
1457 chlorine dioxide, ferrateVI, and ozone) and non-selective oxidants (hydroxyl radical),
1458 Water Res., 2010, 44, 555–566.
1459 178. D. J. de Ridder, A. R. D. Verliefde, S. G. J. Heijman, J. Q. J. C. Verberk, L. C.
1460 Rietveld, L. T. J. van der Aa, G. L. Amy and J. C. van Dijk, Influence of natural
1461 organic matter on equilibrium adsorption of neutral and charged pharmaceuticals onto
1462 activated carbon, Water Sci. Technol., 2011, 63, 416–423.
1463 179. W. Gan, T. Bond, X. Yang and P. Westerhoff, Role of chlorine dioxide in
1464 N-nitrosodimethylamine formation from oxidation of model amines, Environ. Sci.
1465 Technol., 2015, 49, 11429–11437.
1466 180. X. Yang, W. Guo, X. Zhang, F. Chen, T. Ye and W. Liu, Formation of
1467 disinfection by-products after pre-oxidation with chlorine dioxide or ferrate, Water
1468 Res., 2013, 47, 5856–5864.
71
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.
Page 73 of 74

1470
1469

72
Environmental Science: Water Research & Technology

DOI: 10.1039/D0EW00231C
View Article Online

Environmental Science: Water Research & Technology Accepted Manuscript


Environmental Science: Water Research & Technology Page 74 of 74

View Article Online


Graphic abstract DOI: 10.1039/D0EW00231C

Environmental Science: Water Research & Technology Accepted Manuscript


inorganic
e organic
quinones NOM
aldehydes
O
Cl O contaminants
emerging
organic acids
...... ......

chlorite
Published on 15 July 2020. Downloaded on 7/25/2020 2:02:34 AM.

The reaction kinetics and mechanisms of ClO2 reactions with inorganic and organic
compounds are overviewed.

You might also like