Kinetics of Oxidation of Iodide (I–) and Hypoiodous ... - ACS Publications

Dec 30, 2016 - View: ACS ActiveView PDF | PDF | PDF w/ Links | Full Text HTML .... formation of brominated products: Comparison to peroxydisulfate (PD...
0 downloads 0 Views 912KB Size
Subscriber access provided by - Access paid by the | UCLA Library

Letter -

Oxidation Kinetics of Iodide (I) and Hypoiodous Acid (HOI) by Peroxymonosulfate (PMS) and Formation of Iodinated Products in the PMS/I/NOM System -

Juan Li, Jin Jiang, Yang Zhou, Su-yan Pang, Yuan Gao, Chengchun Jiang, Jun Ma, Yi-xin Jin, Yi Yang, Guan-qi liu, Li-Hong Wang, and Chao-ting Guan Environ. Sci. Technol. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.estlett.6b00471 • Publication Date (Web): 30 Dec 2016 Downloaded from http://pubs.acs.org on January 3, 2017

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

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

Page 1 of 28

Environmental Science & Technology Letters

Oxidation Kinetics of Iodide (I-) and

1 2

Hypoiodous Acid (HOI) by

3

Peroxymonosulfate (PMS) and Formation of

4

Iodinated Products in the PMS/I-/NOM

5

System Juan Li, † Jin Jiang, †, * Yang Zhou, †Su–Yan Pang, ‡ Yuan Gao, † Chengchun

6 7

Jiang, §Jun Ma,† Yixin Jin,† Yi Yang, † Guanqi Liu, † Lihong Wang,† and

8

Chaoting Guan†

9



State Key Laboratory of Urban Water Resource and Environment, School of

10

Municipal and Environmental Engineering, Harbin Institute of Technology, Harbin

11

150090, China

12



13

Heilongjiang Province, College of Chemical and Environmental Engineering, Harbin

14

University of Science and Technology, Harbin 150040, China

15

§

16

518055, China

Key laboratory of Green Chemical Engineering and Technology of College of

School of Civil and Environmental Engineering, Shenzhen Polytechnic, Shenzhen

17 1

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 2 of 28

18

* Corresponding

19

Phone: 86−451−86283010; Fax: 86-451−86283010; E-mail: [email protected],

20

[email protected]

21

Abstract

author: Prof. Jin Jiang

22

In this work, the transformation kinetics of iodide (I-) and hypoiodous acid (HOI) by

23

peroxymonosulfate (PMS) and potential formation of iodinated products of concerns

24

in the presence of natural organic matters (NOM) were investigated. As pH increased

25

from 5 to 10, the apparent second-order rate constants of PMS reaction with I- gradually

26

decreased from 1.01×103 to 3.86×102M-1s-1, while those for HOI increased

27

dramatically from 1.08 ×102 to 7.90×104M-1s-1. The obtained pH-dependent rate

28

profiles were well explained by the effects of pH-affected speciation of PMS and/or

29

HOI. Considerable amounts of total organic iodine (TOI) could be formed in the PMS/I-

30

/NOM system over a wide pH range. Under similar conditions, the TOI levels formed

31

in the PMS/I-/NOM system were generally higher than those formed in the case of

32

HOCl but much lower than those formed in the case of NH2Cl. Also, specific iodoform

33

(IF) and monoiodoacetic acid (MIAA) were detected in both simulated and authentic

34

waters during treatment with PMS. This work for the first time demonstrates the

35

potential formation of iodinated products of concerns during water treatment with PMS

36

and thus has important implications on its applications.

37 38

Introduction 2

ACS Paragon Plus Environment

Page 3 of 28

Environmental Science & Technology Letters

39

Iodide (I-) exists ubiquitously in natural environments (water, soil, and minerals)1, 2.

40

The concentrations of I- in surface waters are usually less than 100µg/L. Occasionally,

41

high levels of I- (up to few mg/L) are found in coastal water and hydraulic fracturing

42

contaminated water.3-5 I- can be easily oxidized to hypoiodous acid (HOI) by natural

43

microorganism6 and metal oxides,7-9 as well as by water treatment oxidants such as

44

chlorine (HOCl),10,

45

(KMnO4)17. The formed HOI usually undergoes three competition pathways: (i) further

46

oxidation to nontoxic IO3-, a safe sink of iodine; (ii) reaction with background natural

47

organic matters (NOM) to form toxic iodinated products; and (iii) disproportionation

48

into I- and iodate (IO3-). The relative contribution of reactions (i) - (iii) plays a decisive

49

role in the final fate of iodine.

11

chloramine (NH2Cl),12,

13

ozone(O3),14-16 and permanganate

50

Peroxymonosulfate (PMS), as a relatively stable and inexpensive oxidant, shows great

51

potentials for applications in water treatment and subsurface remediation.18-21 Also,

52

PMS is sometimes used as a broad-spectrum disinfectant in swimming pool and

53

aquaculture disinfection.22-25 For instance, Wang et al.18 reported that PMS displayed

54

high efficiency for the oxidative remediation of arsenite (As(III)) to form As(V), which

55

greatly decreased As(III) toxicity and mobility. Yang et.al.19 found that PMS could

56

effectively remove sulfur-containing odor compound mercaptan in wet scrubbing

57

process. Very recently, Chesney and co-workers26 demonstrated that PMS could rapidly

58

inactivate the disease-associated pathogenic prion protein contaminated land surfaces

59

by an oxidative modification to the amino acid residuals of the peptide fragments. 3

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 4 of 28

60

However, there is only limited information about halogen transformation during PMS

61

oxidation/disinfection. Earlier studies have demonstrated that PMS is able to oxidize

62

Cl-, Br-, and I- to form reactive halogen species HOX (i.e., HOCl, HOBr, and HOI)

63

with second order reaction rate constants kx- decreasing in the order of kI- (1400

64

~1730M-1s-1) > kBr- (0.7~1.0M-1s-1) > kCl- (0.0021~0.0018M-1s-1).27-29 The further

65

oxidation of HOCl and HOBr by PMS were negligible28, while the reaction of HOI

66

with PMS has not yet been investigated. The rapid transformation of I- to HOI by PMS

67

indicates the potential risk of the formation of iodinated products, whose toxicity are

68

generally several to hundreds of times more genotoxic and cytotoxic than their

69

chlorinated and brominated analogues.30-32 This issue, however, has not been addressed

70

so far.

71

The objectives of this study were (i) to investigate the oxidation kinetics of I- and HOI

72

by PMS over a wide pH range (5~10), and (ii) to evaluate the formation of iodinated

73

products (including total organic iodine (TOI), iodoform (IF), and monoiodoacetic acid

74

(MIAA)) during water treatment with PMS, as compared to HOCl and NH2Cl.

75

Materials and Methods.

76

Materials. PMS (available as Oxone), NaClO (4% active chlorine), phenol, 2-

77

iodophenol, and 4-iodophenol were purchased from Sigma-Aldrich. KI, KIO3, and

78

ammonium chloride (NH4Cl) were purchased from Sinopharm Chemical Reagent Co.

79

Ltd., China. IF (99%) and MIAA (97%) were purchased form J & K Scientific Ltd.,

80

China. Suwannee river humic acid (SRHA, 2S101H) and Suwannee river fulvic acid 4

ACS Paragon Plus Environment

Page 5 of 28

Environmental Science & Technology Letters

81

(SRFA, 1S101F) were purchased from International Humic Substance Society (IHSS).

82

Another humic acid was purchased from Sigma-Aldrich (Sigma HA) and purified

83

following the procedure as previously described.33 All other reagents were of analytical

84

grade or better and used without further purification. All solutions were prepared using

85

deionized (DI) water (18.2 MΩ/cm) from a Milli-Q purification system (Millipore,

86

Billerica, MA). Stock solutions of oxidants (i.e., PMS, HOCl, and NH2Cl) were

87

prepared and standardized as described in SI Text S1.

88

Reaction kinetics. Reaction kinetics of PMS with I- were investigated by a stopped-

89

flow spectrophotometer under pseudo-first-order conditions with I- in excess in the pH

90

range of 5~10. Details can be found in SI Text S2.

91

Batch reactions for PMS with HOI were initiated by adding excess PMS into pH-

92

buffered solutions containing freshly prepared HOI (stoichiometric oxidation of I- by

93

OCl-) in the pH range of 5~8. Samples were periodically collected and quenched by

94

phenol in excess to trap the unreacted HOI,34 and thereafter As(III)

95

quench the residual PMS in the samples before analysis for iodophenols, I-, and IO3-.

96

(see SI Text S3 for the details). For pH 9~10, a sequential-mixing stopped-flow

97

technique (see SI Text S4 for the details) was used since the reaction was too fast to be

98

followed by manual operation.

18

was added to

99

Formation of iodinated products. The experiments for the formation of iodinated

100

products were conducted in 250mL amber glass bottles and the headspace free

101

conditions were kept. Pre-determined amounts of oxidant (PMS, HOCl, or NH2Cl) 5

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 6 of 28

102

were added into pH-buffered solutions containing I- and NOM representative to initiate

103

the reactions. Samples were withdrawn after 24h reaction for the determination of

104

iodine species (i.e., I-, HOI, IO3-, TOI, IF, and MIAA) (See SI Text S5 for the details).

105

Natural water taken from Songhua river, Harbin, China (DOC = 6.2mg·C/L, alkalinity

106

= 1.3mM as HCO3-, and pH =7.3) was stored at 4°C and used within two days after

107

vacuum-filtered through 0.45µM cellulose membrane filter. A similar procedure to that

108

used in simulated water was followed with the exception that I- at an environmentally

109

relevant level (i.e., 0.5µM) was spiked. Samples were withdrawn after 24h reaction for

110

the determination of IF and MIAA.

111

All experiments were conducted at 23  2°C in duplicates or triplicates and the

112

average data with their standard deviations were displayed. Phosphate buffer (2mM)

113

and borate buffer (2mM) were used for pH 5~7 and pH 8~10, respectively.

114

Analytical

methods.

A stopped-flow

spectrophotometer

(SX20,

Applied

115

Photophysics Ltd.) equipped with a photomultiplier tube (PMT) detector and a

116

sequential-mixing accessory was used to carry out the fast kinetics. Iodophenols were

117

analyzed by HPLC (details can be found in SI Text S6). I- and IO3- were determined by

118

ion chromatography (IC, Thermo Dionex ICS-3000) (see SI Text S6 for the details). IF

119

and MIAA were determined by liquid/liquid extraction with methyl-tert-butyl ether

120

(MtBE) without/with acidic methanol derivation followed by gas chromatography and

121

electron capture detection (GC/ECD, GC-6890 Agilent) according to EPA Method

122

551.1 and 552.2., respectively. TOI was measured by a Multi 2500 TOX analyzer (Jena) 6

ACS Paragon Plus Environment

Page 7 of 28

Environmental Science & Technology Letters

123

via an adsorption-pyrolysis-titration method (see SI Text S7 for the details).

124

Results and Discussion

125

Reaction Kinetics of PMS with I-. The reaction of PMS with I- was determined to be

126

first-order with respect to each reactant (see SI Figure S1). The obtained apparent

127

second-order rate constants ( kI- , PMS ) as a function of pH were shown in Figure 1a. As

128

can be seen, PMS exhibited considerable reactivity towards I- with kI- , PMS decreasing

129

from 1013(±89) to 386 (±13) M-1s-1 as pH increased from 5 to 10. This pH dependence

130

of kI- , PMS can be well explained by the reactions between pH-affected species of PMS

131

(HSO5- and SO52; pKa=9.30) and I- (solid line in Figure 1a). The species-specific rate

132

constants calculated by nonlinear least-squares regression of experimental data were

133

k I- +HSO - = 1112 (±29) M-1s-1 and k I- +SO 2- =218 (±73) M-1s-1, respectively, and they 5 5

134

well matched the results given by Secco27 and Lente28. Contributions of individual

135

reactions of HSO5- and SO52- with I- to the overall reaction rate were calculated and

136

shown in Figure 1a (dashed and dot-dashed lines). Moreover, a comparison of pH-

137

dependent apparent second-order rate constants ( kI- ) for the reactions between I- and

138

various oxidants [PMS, HOCl, KMnO4, O3, and NH2Cl] was made. As shown in Figure

139

1b, PMS displayed a mild reactivity with I- compared to other oxidants.

140

Figure 1

141

Reaction Kinetics of PMS with HOI. The kinetics of PMS reaction with HOI were

142

further investigated in the pH range of 5~10. Disappearance of HOI in the presence of

143

excess PMS followed the pseudo-first-order rate law (see SI Figure S3), confirming 7

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 8 of 28

144

that the reaction was first-order with respect to HOI. Pseudo-first-order rate constants

145

(kobs, s-1) determined at various concentrations of PMS at a constant pH showed a

146

linearity (SI Figure S4, for example), demonstrating that the reaction was also first-

147

order with respect to PMS. Measured apparent second-order rate constants ( kHOI,PMS ,

148

M-1s-1) as a function of pH were summarized in Figure 2a.

149

Figure 2

150

The kHOI,PMS values showed a strong pH dependence with increasing more than 2

151

orders of magnitude from pH 5 to 10. This pH dependence can be quantitatively

152

described by the parallel reactions between individual acid-base species of HOI

153

(pKa=10.43) and PMS (pKa=9.30), as shown by reactions (1~8) (where reactions (3~6)

154

are rate-determining35):

155 156

HSO 5  SO52  H +

pK a  9.30

(1)

HOI  OI  H +

pK a  10.43

(2)

HSO5 +HOI  IO2  SO42  2H +

kHOI-1,1

(3)

HSO5 +OI  IO2  SO42  H+

kHOI-1,2

(4)

SO52 +HOI  IO2  SO24  H +

kHOI-2,1

(5)

SO52 +OI  IO2  SO42

kHOI-2,2

(6)

HSO5 +IO2  IO3  SO24 +H+

fast

(7)

SO52 +IO2  IO3  SO42

fast

(8)

Accordingly, kHOI,PMS is given by

kHOI, PMS = j1,2 kHOI-i,j i  j i=1,2

(9) 8

ACS Paragon Plus Environment

Page 9 of 28

Environmental Science & Technology Letters

i

 j represent the respective fractions of PMS and HOI as the species

157

where

158

i and j at a given pH, and kHOI-i,j represents the species-specific second-order rate

159

constant for each i and j pair. The kHOI-i,j values determined by nonlinear least-

160

squares regression of experimental data ( kHOI,PMS ) were kHOI-1,1 = 112( ±18), kHOI-1,2

161

= 1.7(±0.16) ×106, kHOI-2,1 = ~0, and kHOI-2,2 = 1.5(±0.74) ×105M1s-1, respectively.

162

Accordingly, the contribution of each reaction (i.e., reactions 3~6) to the overall

163

reaction rate was calculated (Figure 2a, dashed lines). As can be seen, the reaction

164

between HSO5- and HOI dominates at lower pH (pH7).

and

166

A comparison of the pH-dependent apparent second-order rate constants ( kHOI ) for

167

the reactions between HOI and selective oxidants [PMS, HOCl, KMnO4, O3, and

168

NH2Cl] was also made. As shown in Figure 2b, PMS displayed high reactivity towards

169

HOI compared to HOCl, NH2Cl, and KMnO4.

170

Interestingly, it was found that the oxidation rates of HOI to IO3- by PMS (Figure 2a)

171

at pH>8 were even faster than those of I- to HOI by PMS (Figure 1a). The exact reasons

172

for the unexpectedly high reactivity of PMS with HOI are unclear so far, which

173

deserves further investigations. Previous studies have reported that PMS can undergo

174

self-decomposition especially at alkaline pH, where reactive oxygen species (ROS)

175

such as singlet oxygen (1O2) is generated.

176

scavengers such as furfuryl alcohol (for 1O2) and methanol (for sulfate radical, SO4.-)

177

on the oxidation kinetics of HOI by PMS was observed (data not shown). Either

36, 37

However, no influence of specific

9

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 10 of 28

178

negligible decomposition of PMS was observed in control experiments without HOI

179

within the time scales investigated.

180

Evolution of IO3- during oxidation of HOI by PMS in the pH range of 5~10 was also

181

investigated and the results were shown in SI Figure S5. As can be seen, a good mass

182

balance (i.e., [HOI] + [IO3-]) was maintained during the kinetic runs, indicating that

183

IO3- was the unique product for HOI oxidation by PMS.

184

Stoichiometry. Stoichiometries for the reaction of PMS with I- were further evaluated.

185

As shown in SI Table S1, the amounts of PMS consumed (Δ[PMS]) were

186

approximately 3 times of the amounts of IO3- formed (Δ[IO3-]) (Δ[PMS]/Δ[IO3-]=3),

187

which was consistent with the theoretical value according to eq 10.

188

3HSO5- + I-  IO3  3SO24  3H+

(10)

189

Formation of iodinated products in the presence of NOM.

190

(i) PMS. Figure 3a comparatively displayed the TOI formation from three NOM

191

representatives (Sigma HA, SRHA, and SRFA) in the presence of PMS and I- at pH 5,

192

7, and 9. For all the tested NOM representatives, the TOI levels showed obvious

193

dependence on pH and decreased with the increase of pH. This result can be well

194

explained by the pH-affected oxidation rates of HOI formed in situ by PMS. At

195

investigated pH (5, 7, and 9), I- was oxidized to HOI by PMS with comparable rate

196

constants (Figure 1a), while the oxidation rates of HOI to IO3- by PMS increased more

197

than 2 orders of magnitude from pH 5 to 9 (Figure 2a). Accordingly, a higher HOI

198

exposure was available at lower pH, leading to the observed higher TOI level. 10

ACS Paragon Plus Environment

Page 11 of 28

Environmental Science & Technology Letters

199

Meanwhile, considerable amounts of IF (4.5 ~ 25nM) and MIAA (27 ~ 65nM) were

200

also detected during treatment with PMS in the presence of I- and NOM representative

201

(Figure 3b). The total iodine that incorporated in IF and MIAA (i.e., 3×[IF]+[MIAA])

202

accounted for a small fraction (PMS>HOCl at 11

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 12 of 28

220

pH 5 and 7, while it changed to NH2Cl>HOCl>PMS at pH 9 for each NOM

221

representative. The trend obtained at pH 5 and 7 was somewhat unexpected, since PMS

222

exhibited a much higher reactivity towards HOI than HOCl therein (Figure 2b). This

223

finding probably resulted from the competitive reaction of HOCl in high concentrations

224

vs HOI in low concentrations towards reactive sites in NOM, although HOCl might be

225

less reactive toward NOM than HOI.38, 39 The oxidant demands in the case of HOCl

226

were much higher than those obtained in the case of PMS for each NOM representative

227

(SI Figure S7). Allard and co-workers40 also reported a similar competition effect of

228

HOCl for reactive sites of NOM, which decreased the TOI formation from chlorination

229

of UV-irradiated iopamidol in the presence of NOM. It seemed likely that this

230

competition effect of HOCl was weakened at pH 9 due to its deprotonation into OCl-

231

(pKa=7.54). Another possible explanation might involve the exchange of iodine from

232

the already formed iodinated products by chlorine from HOCl (i.e., the transformation

233

of TOI to TOCl) via the ipso free-radical substitution process.41-43 Recently, Wendel et

234

al.44 proposed that this chlorine-iodine exchange process resulted in the formation of

235

chlorinated products during chlorination of iopamidol. Zhu and Zhang45 reported that

236

the transformation of TOI to TOCl in the presence of chlorine residue via chlorine-

237

iodine exchange consumed ~15% of the TOI initially formed.

238

The formation of IF and MIAA during treatment with PMS vs HOCl/NH2Cl was

239

further investigated (Figure 4b). As shown, IF was predominantly formed in the case

240

of NH2Cl, while MIAA was preferentially formed in the case of HOCl. Both IF and 12

ACS Paragon Plus Environment

Page 13 of 28

Environmental Science & Technology Letters

241

MIAA were moderately formed during treatment with PMS. The higher IF

242

concentrations obtained in the case of NH2Cl was in agreement with the higher HOI

243

exposures therein that enabled triple iodination to IF.46

244

Meanwhile, differences in TOI levels were also observed among three NOM

245

representatives for each oxidant under similar conditions. The aromatic moiety

246

contents in NOM (i.e., SUVA, specific UV absorbance at 254nm, expressed in L·mgC-

247

1

248

disinfection byproducts during treatment with HOCl or NH2Cl.47-50 In light of this, the

249

relationship between the levels of iodinated products (i.e., TOI/IF/MIAA) and SUVA

250

values of selected NOM representatives (SI Table S2) for each oxidant was further

251

examined, but no obvious correlation was found (data not shown). This observation

252

was probably resulted from the different reaction pathways between NOM and HOI vs

253

HOCl/NH2Cl and/or the transformation of iodinated products to their chlorinated

254

analogues in the presence of residual HOCl as discussed above. Zhu and Zhang45

255

reported that the majority of TOCl (77.5 ~ 84.7%) were resulted from the reactions

256

between HOCl and NOMslow (i.e., slow reaction sites in NOM), whereas only 2.2 ~ 22.8%

257

of the TOI came from the reaction of HOI and NOMslow during chlorination of NOM

258

in the presence of I-.

·m-1) always have good correlations with the formation potentials of chlorinated

259

In addition, as shown in Figure 4a, considerable amounts of HOI were detected in the

260

case of NH2Cl. This was due to the much slower oxidation rates of HOI by NH2Cl vs

261

HOCl and PMS. Interestingly, the occurrence of I- and IO3- were also observed in the 13

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 14 of 28

262

case of NH2Cl, which was somewhat unexpected according to the kinetic rate constants

263

of I- and HOI with NH2Cl (Figure 1b and Figure 2b). This finding was explained by

264

the contributions of side reactions including (i) the disproportionation of HOI51-53; (ii)

265

the reduction of HOI back to I- by NOM (i.e., oxidation pathway for the reaction of

266

HOI with NOM other than substitution pathway)34, 54; and (iii) further oxidation of HOI

267

by HOCl in situ formed from the slow hydrolysis of NH2Cl55-57.

268

Formation of iodinated products from authentic water. Figure 5 comparatively

269

showed the formation of IF and MIAA from natural water containing 0.5µM I- treated

270

by HOCl, NH2Cl, and PMS (15μM). As can be seen, IF at concentration of ~1.5nM

271

was formed during treatment with PMS, which was much lower than those formed in

272

the cases of HOCl and NH2Cl (i.e., 4.0nM and 10.5nM, respectively). MIAA formed

273

during treatment with PMS (~7.6nM) was comparable to that formed in the case of

274

NH2Cl (~10nM) but much lower than that formed in the case of HOCl (~21nM).

275

Figure 5

276

To the best of our knowledge, this work for the first time examines the oxidation

277

kinetics of I- and HOI by PMS and demonstrates the potential formation of iodinated

278

products in the PMS/I-/NOM system. The formation of IF and MIAA from natural

279

water containing I- treated by PMS has been confirmed. These results have important

280

implications for the future applications of PMS-based oxidation/disinfection processes.

281

Acknowledgment

282

This work was financially supported by the National Natural Science Foundation of 14

ACS Paragon Plus Environment

Page 15 of 28

Environmental Science & Technology Letters

283

China (51578203&51378316), the National Key Research and Development Program

284

(2016YFC0401107), the Chinese Postdoctoral Science Foundation (2015T80366), the

285

Funds of the State Key Laboratory of Urban Water Resource and Environment (HIT,

286

2016DX13), the Foundation for the Author of National Excellent Doctoral Dissertation

287

of China (201346), Heilongjiang Province Natural Science Foundation (QC2014C055),

288

and the Fundamental Research Funds for the Central Universities of China.

289

Supporting Information

290

The additional texts, figures, and tables addressing supporting data. This material is

291

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

292

References

293

1. Richardson, S. D.; Fasano, F.; Ellington, J. J.; Crumley, F. G.; Buettner, K. M.;

294

Evans, J. J.; Blount, B. C.; Silva, L. K.; Waite, T. J.; Luther, G. W.; et al., Occurrence

295

and mammalian cell toxicity of iodinated disinfection byproducts in drinking water.

296

Environ. Sci. Technol. 2008, 42, 8330-8338.

297 298

2. Fuge, R.; Johnson, C., The geochemistry of iodine- a review. Environ. Geochem. Health. 1986, 8, 31-54.

299

3. Lee, H.; Kim, H.; Lee, H.; Lee, C., Reaction of aqueous iodide at high

300

concentration with O3 and O3/H2O2 in the presence of natural organic matter:

301

Implications for drinking water treatment. Environ. Chem. Lett. 2015, 13, 453-458.

302

4. Parker, K. M.; Zeng, T.; Harkness, J.; Vengosh, A.; Mitch, W. A., Enhanced 15

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 16 of 28

303

formation of disinfection byproducts in shale gas wastewater-impacted drinking water

304

supplies. Environ. Sci. Technol. 2014, 48, 11161-11169.

305

5. Harkness, J. S.; Dwyer, G. S.; Warner, N. R.; Parker, K. M.; Mitch, W. A.;

306

Vengosh, A., Iodide, bromide, and ammonium in hydraulic fracturing and oil and gas

307

wastewaters: Environmental implications. Environ. Sci. Technol. 2015, 49, 1955-1963.

308

6. Itoh, N.; Tsujita, M.; Ando, T.; Hisatomi, G.; Higashi, T., Formation and emission

309

of monohalomethanes from marine algae. Phytochemistry 1997, 45, 67-73.

310

7. Gallard, H.; Allard, S.; Nicolau, R.; von Gunten, U.; Croué, J. P., Formation of

311

iodinated organic compounds by oxidation of iodide-containing waters with

312

manganese dioxide. Environ. Sci. Technol. 2009, 43, 7003-7009.

313

8. Lin, Y.; Washburn, M. P.; Valentine, R. L., Reduction of lead oxide (PbO2) by

314

iodide and formation of iodoform in the PbO2/I-/NOM system. Environ. Sci. Technol.

315

2002, 42, 2919-2924.

316

9. Allard, S.; von Gunten, U.; Sahli, E.; Nicolau, R.; Gallard, H., Oxidation of iodide

317

and iodine on birnessite (δ-MnO2) in the pH range 4–8. Water Res. 2009, 43, 3417-

318

3426.

319

10. Nagy, J. C.; Kumar, K.; Margerum, D. W.; Carrie A. Delcomyn, Non-metal redox

320

kinetics: Oxidation of iodide by hypochlorous acid and by nitrogen trichloride

321

measured by the pulsed-accelerated-flow method. Inorg. Chem. 1988, 27, 2773-2780.

322

11. Criquet, J.; Allard, S.; Salhi, E.; Joll, C. A.; Heitz, A.; von Gunten, U., Iodate and

323

iodo-trihalomethane formation during chlorination of iodide-containing waters: Role 16

ACS Paragon Plus Environment

Page 17 of 28

Environmental Science & Technology Letters

324

of bromide. Environ. Sci. Technol. 2012, 46, 7350-7357.

325

12. Krishan, K.; Richard, A. D.; Margerum, D. W., Atom-transfer redox kinetics:

326

General-acid-assisted oxidation of iodide by chloramines and hypochlorite. Inorg.

327

Chem. 1986, 25, 4344-4350.

328

13. Yang, Y.; Komaki, Y.; Kimura, S. Y.; Hu, H.; Wagner, E. D.; Mariñas, B. J.; Plewa,

329

M. J., Toxic impact of bromide and iodide on drinking water disinfected with chlorine

330

or chloramines. Environ. Sci. Technol. 2014, 48, 12362-12369.

331

14. Liu, Q.; Schurter, L. M.; Muller, C. E.; Aloisio, S.; Francisco, J. S.; Margerum, D.

332

W., Kinetics and mechanisms of aqueous ozone reactions with bromide, sulfite,

333

hydrogen sulfite, iodide, and nitrite ions. Inorg. Chem. 2001, 40, 4436-4442.

334

15. Allard, S.; Nottle, C. E.; Chan, A.; Joll, C.; von Gunten, U., Ozonation of iodide-

335

containing waters: Selective oxidation of iodide to iodate with simultaneous

336

minimization of bromate and I-THMs. Water Res. 2013, 47, 1953-1960.

337

16. Magi, L.; Schweitzer, F.; Pallares, C.; Cherif, S. M.; Mirabel, P.; George, C.,

338

Investigation of the uptake rate of ozone and methyl hydroperoxide by water surfaces.

339

J. Phys. Chem. A 1997, 4943-4949.

340

17. Zhao, X.; Salhi, E.; Liu, H.; Ma, J.; von Gunten, U., Kinetic and mechanistic

341

aspects of the reactions of iodide and hypoiodous acid with permanganate: Oxidation

342

and disproportionation. Environ. Sci. Technol. 2016, 50, 4358-4365.

343

18. Wang, Z.; Bush, R. T.; Sullivan, L. A.; Chen, C.; Liu, J., Selective oxidation of

344

arsenite by peroxymonosulfate with high utilization efficiency of oxidant. Environ. Sci. 17

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 18 of 28

345

Technol. 2014, 48, 3978-3985.

346

19. Yang, S.; Li, Y.; Wang, L.; Feng, L., Use of peroxymonosulfate in wet scrubbing

347

process for efficient odor control. Sep. Purif. Technol. 2015, 158, 80-86.

348

20. Yu, M.; Teel, A. L.; Watts, R. J., Activation of peroxymonosulfate by subsurface

349

minerals. J. Contam. Hydrol. 2016, 191, 33-43.

350

21. Zhou, Y.; Jiang, J.; Gao, Y.; Ma, J.; Pang, S.; Li, J.; Lu, X.; Yuan, L., Activation

351

of peroxymonosulfate by benzoquinone: A novel nonradical oxidation process.

352

Environ. Sci. Technol. 2015, 49, 12941-12950.

353

22. Su, X.; D'Souza, D. H., Inactivation of human norovirus surrogates by

354

benzalkonium chloride, potassium peroxymonosulfate, tannic acid, and gallic acid.

355

Foodborne Pathog. Dis. 2012, 9, 829-834.

356

23. Danner, G. R.; Merrill, P., Disinfectants, disinfection, and biosecurity in

357

aquaculture. In Aquaculture biosecurity prevention, control, and eradication of aquatic

358

animal disease; Scarfe, A. D.; Lee, C.; O'Bryen, P. J.; Blackwell Publishing

359

Professional: Boston, 2006; pp 91-128.

360

24. Maninous, M. E.; Simth, S. A.; Kuhn, D. D., Effect of common aquaculture

361

chemicals against Edwardsiella ictaluri and E. tarda. J. Aquat. Anim. Health 2010, 22,

362

224-228.

363

25. Mainous, M. E.; Kuhn, D. D.; Smith, S. A., Efficacy of common aquaculture

364

compounds for disinfection of aeromonas hydrophila, A. salmonicida subsp.

365

salmonicida, and A. salmonicida subsp. achromogenes at various temp. North Am. J. 18

ACS Paragon Plus Environment

Page 19 of 28

Environmental Science & Technology Letters

366

Aquacult. 2011, 73, 456-461.

367

26. Chesney, A. R.; Booth, C. J.; Lietz, C. B.; Li, L.; Pedersen, J. A.,

368

Peroxymonosulfate rapidly inactivates the disease-associated prion protein. Environ.

369

Sci. Technol. 2016, 50, 7095-7105.

370

27. Secco, F.; Venturini, M., Mechanisms of peroxide reactions: Kinetics of reduction

371

of peroxomonosulphuric and peroxomonophosphoric acids by iodide ion. J. Chem.

372

Soc., Dalton Trans. 1976, 1410-1414.

373

28. Lente, G.; Kalmár, J.; Baranyai, Z.; Kun, A.; Kék, I.; Bajusz, D.; Takács, M.; Veres,

374

L.; Fábián, I., One-versus two-electron oxidation with peroxomonosulfate ion:

375

Reactions with iron(II), vanadium(IV), halide ions, and photoreactionwith cerium(III).

376

Inorg. Chem. 2009, 48, 1763-1773.

377

29. Donald, H. F.; Charles. J. B.; Stephen R. C.; John, O. E., The kinetics of the

378

oxidation of halide ions by monosubstituted peroxides. J. Am. Chem. Soc. 1960, 82,

379

778-782.

380

30. Hladik, M. L.; Hubbard, L. E.; Kolpin, D. W.; Focazio, M. J., Dairy-impacted

381

wastewater is a source of iodinated disinfection byproducts in the environment.

382

Environ. Sci. Technol. Lett. 2016, 3, 190-193.

383

31. Wendel, F. M.; Ternes, T. A.; Richardson, S. D.; Duirk, S. E.; Pals, J. A.; Wagner,

384

E. D.; Plewa, M. J., Comparative toxicity of high-molecular weight iopamidol

385

disinfection byproducts. Environ. Sci. Technol. Lett. 2016, 3, 81-84.

386

32. Richardson, S.; Plewa, M.; Wagner, E.; Schoeny, R.; Demarini, D., Occurrence, 19

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 20 of 28

387

genotoxicity, and carcinogenicity of regulated and emerging disinfection by-products

388

in drinking water: A review and roadmap for research. Mutat. Res., Rev. Mutat. Res.

389

2007, 636, 178-242.

390

33. Hong, S.; Elimelech, M., Chemical and physical aspects of natural organic matter

391

(NOM) fouling of nanofiltration membranes. J. Membr. Sci. 1997, 132, 159-181.

392

34. Bichsel, Y.; von Gunten, U., Formation of iodo-trihalomethanes during

393

disinfection and oxidation of iodide-containing waters. Environ. Sci. Technol. 2000,

394

34, 2784-2791.

395

35. Bichsel, Y.; von Gunten, U., Oxidation of iodide and hypoiodous acid in the

396

disinfection of natural waters. Environ. Sci. Technol. 1999, 33, 4040-4045.

397

36. Qi, C.; Liu, X.; Ma, J.; Lin, C.; Li, X.; Zhang, H., Activation of peroxymonosulfate

398

by base: Implications for the degradation of organic pollutants. Chemosphere 2016,

399

151, 280-288.

400

37. Evans, D. F.; Upton, M. W., Studies on singlet oxygen in aqueous solution. Part 3.

401

The decomposition of peroxy-acids. J. Chem. Soc., Dalton Trans. 1985, 1151-1153.

402

38. Vikesland, P. J.; Fiss, E. M.; Wigginton, K. R.; McNeill, K.; Arnold, W. A.,

403

Halogenation of bisphenol-A, triclosan, and phenols in chlorinated waters containing

404

iodide. Environ. Sci. Technol. 2013, 47, 6764-6772.

405

39. Lee, Y.; Yoon, J.; von Gunten, U., Kinetics of the oxidation of phenols and

406

phenolic endocrine disruptors during water treatment with ferrate (Fe(VI)). Environ.

407

Sci. Technol. 2005, 39, 8978-8984. 20

ACS Paragon Plus Environment

Page 21 of 28

Environmental Science & Technology Letters

408

40. Allard, S.; Criquet, J.; Prunier, A.; Falantin, C.; Le Person, A.; Yat-Man Tang, J.;

409

Croué, J., Photodecomposition of iodinated contrast media and subsequent formation

410

of toxic iodinated moieties during final disinfection with chlorinated oxidants. Water

411

Res. 2016, 103, 453-461.

412

41. Gao, Y.; Pang, S.; Jiang, J.; Ma, J.; Zhou, Y.; Li, J.; Wang, L.; Lu, X.; Yuan, L.,

413

Transformation of flame retardant tetrabromobisphenol A by aqueous chlorine and the

414

effect of humic acid. Environ. Sci. Technol. 2016, 50, 9608-9618.

415

42. Traynham, J. G., Ipso substitution in free-radical aromatic substitution reactions.

416

Chem. Rev. 1979, 79, 323-330.

417

43. Sasson, Y., Formation of carbon–halogen bonds (Cl, Br, I). Chem. Halides,

418

Pseudo-Halides Azides 2004, 535-628.

419

44. Wendel, F. M.; Lütke Eversloh, C.; Machek, E. J.; Duirk, S. E.; Plewa, M. J.;

420

Richardson, S. D.; Ternes, T. A., Transformation of iopamidol during chlorination.

421

Environ. Sci. Technol. 2014, 48, 12689-12697.

422

45. Zhu, X.; Zhang, X., Modeling the formation of TOCl, TOBr, and TOI during

423

chlor(am)ination of drinking water. Water Res. 2016, 96, 166-176.

424

46. Hua, G.; Reckhow, D. A., Comparison of disinfection byproduct formation from

425

chlorine and alternative disinfectants. Water Res. 2007, 41, 1667-1678.

426

47. Rostad, C. E.; Martin, B. S.; Barber, L. B.; Leenheer, J. A.; Daniel, S. R., Effect

427

of a constructed wetland on disinfection byproducts: Removal processes and

428

production of precursors. Environ. Sci. Technol. 2000, 34, 2703-2710. 21

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 22 of 28

429

48. Kristiana, I.; Gallard, H.; Joll, C.; Croué, J., The formation of halogen-specific

430

TOX from chlorination and chloramination of natural organic matter isolates. Water

431

Res. 2009, 43, 4177-4186.

432

49. Yang, X.; Shang, C.; Lee, W.; Westerhoff, P.; Fan, C., Correlations between

433

organic matter properties and DBP formation during chloramination. Water Res. 2008,

434

42, 2329-2339.

435

50. Reckhow, D. A.; Singer, P. C.; Malcolm, R. L., Chlorination of humic materials:

436

Byproduct formation and chemical interpretations. Environ. Sci. Technol. 1990, 24,

437

1655-1664.

438

51. Bichsel, Y.; von Gunten, U., Hypoiodous acid: Kinetics of the buffer-catalyzed

439

disproportionation. Water Res. 2000, 34, 3197-3203.

440

52. Truesdale, V. W.; Canosa-Mas, C., Kinetics of disproportionation of hypoiodous

441

acid in phosphate and borate buffer at pH < 8.5 modelled using iodide feedback. J.

442

Chem. Soc., Faraday Trans. 1995, 91, 2269-2273.

443

53. Krisztina, S.; Körtvélyesi, T., Kinetics and mechanism of the hydrolytic

444

disproportionation of iodine. Int. J. Chem. Kinet. 2004, 36, 596-602.

445

54. Guan, C.; Jiang, J.; Luo, C.; Ma, J.; Pang, S.; Jiang, C.; Jin, Y.; Li, J.,

446

Transformation of iodide by carbon nanotube activated peroxydisulfate and formation

447

of iodoorganic compounds in the presence of natural organic matter. Environ. Sci.

448

Technol. 2016. DOI: 10.1021/acs.est.6b04158

449

55. Vikesland, P. J.; Ozekin, K.; Valentine, R. L., Monochloramine decay in model 22

ACS Paragon Plus Environment

Page 23 of 28

Environmental Science & Technology Letters

450

and distribution system waters. Water Res. 2001, 35, 1766-1776.

451

56. Doederer, K.; Gernjak, W.; Weinberg, H. S.; Farré, M. J., Factors affecting the

452

formation of disinfection by-products during chlorination and chloramination of

453

secondary effluent for the production of high quality recycled water. Water Res. 2014,

454

48, 218-228.

455

57. Diehl, A.; Speitel, G.; Symons, J.; Krasner, S.; Hwang, S.; Barrett, S., DBP

456

formation during chloramination. J. Am. Water Works Assoc. 2000, 92, 76-90.

457

23

ACS Paragon Plus Environment

Environmental Science & Technology Letters

-

10

9

10

7

10

5

10

3

10

1

-1 -1

kI (M s )

-1 -1

kI ,PMS ( M s )

-

600 measured data model fit

300

(b) O3

10

900

PMS KMnO4

10 5

6

7

8

9

10

HOCl

NH2Cl

-1

0

458

11

(a)

1200

Page 24 of 28

5

6

pH

7

8

9

10

pH

459

FIGURE 1. (a) pH-dependent second-order rate constants ( kI- , PMS , M−1s−1) for

460

the reaction of PMS with I-. Symbols represent measured data and the lines

461

indicate the model prediction and contribution of individual reactions of HSO 5-

462

with I- (dashed) and SO52- with I- (dot dashed) to the overall reaction as a function

463

of pH. (b) Comparison of the pH-dependent apparent second-order rate constants

464

for the reactions of selective oxidants with I- ( kI-, M-1s-1).The kI- values for HOCl

465

were obtained from ref 10, for O3 obtained from ref 16, for KMnO4 from ref 17,

466

and for NH2Cl from ref 12.

467 468

24

ACS Paragon Plus Environment

Page 25 of 28

Environmental Science & Technology Letters

5

7

10

2

HO

I-2 ,2







10

3

10

469

5

-1 -1

10

 

1

k HO

,2 I-1

5

6

kHOI-1,1

7

(b) O3

10

model fit

k

-1 -1

kHOI,PMS( M s )

measured data 4

10

(a)

kHOI ( M s )

10

8

9

3

10

PMS

1

10

10

-1

10

-3

10

HOCl

KMnO4 NH2Cl

5

6

pH

7

8

9

10

pH

470

FIGURE 2. (a) pH-dependent second-order rate constants (kHOI,PMS, M−1s−1) for

471

the reaction of PMS with HOI. The symbols represent measured data and the

472

lines indicate the model prediction with eq 9 (solid) and contribution of individual

473

reactions of HSO5- with HOI (kHOI-1,1α1β1, dot dashed), HSO5- with OI- (kHOI-1,2α1β2,

474

dot-dot dashed), and SO52- with OI- (kHOI-2,2α2β2, short dashed) to the overall

475

reaction as a function of pH. (b) Comparison of the pH-dependent second-order

476

rate constants for the reactions of selective oxidants with HOI (kHOI, M-1s-1). The

477

kHOI values for HOCl, O3, and NH2Cl were obtained from ref 35 (the third-order

478

reaction was not considered in the case of HOCl, and the maximum species-

479

specific rate constants estimated in ref 35 were used to calculate the kHOI values in

480

the case of NH2Cl), and for KMnO4 from ref 17.

481 482 483 484 485 486 487 25

ACS Paragon Plus Environment

Environmental Science & Technology Letters

TOI

IO3

-

(a)

10 8 6 4 2 0

pH 5

IF

Iodinated products (nM)

Iodinated products (M)

12

Page 26 of 28

MIAA

(b)

80

60

40

20

0

7

9

NOM Sigma HA

5

7

9

SRHA

5

7 SRFA

9

pH 5

7

9

NOM Sigma HA

5

7

9

SRHA

5

7

9

SRFA

488 489

FIGURE 3. Formation of iodinated products from I- oxidation by PMS in the

490

presence of NOM. (a) TOI and IO3-; (b) IF and MIAA. Experimental conditions:

491

[I-] = 10μM, [PMS] = 100μM, [Sigma HA] = [SRHA] = [SRFA] = 4mg·C/L, and

492

reaction time t = 24h.

493 494 495 496 497 498 499 500 501 502

26

ACS Paragon Plus Environment

Page 27 of 28

Environmental Science & Technology Letters

Iodinated products (M)

12

TOI

HOI

-

-

I

IO3

(a)

10 8 6 4 2 0

pH 5 7 9 5 7 9 5 7 9 Oxidant Oxi-1 Oxi-2 Oxi-3 NOM Sigma HA

503

5 7 9 5 7 9 5 79

5 7 9 5 79 5 7 9

Oxi-1 Oxi-2 Oxi-3 SRHA

Oxi-1 Oxi-2 Oxi-3 SRFA

500 IF

MIAA

(b)

Iodinated products(nM)

400

300

200

100

0 pH 5 7 9 5 7 9 5 7 9 Oxidant Oxi-1 Oxi-2 Oxi-3 NOM Sigma HA

504

5 7 9 5 7 9 5 79

5 7 9 5 79 5 7 9

Oxi-1 Oxi-2 Oxi-3 SRHA

Oxi-1 Oxi-2 Oxi-3 SRFA

505

FIGURE 4. Comparison of iodinated products formed from I- oxidation by NH2Cl

506

(Oxi-1), HOCl (Oxi-2), or PMS (Oxi-3) in the presence of NOM. (a) TOI and IO3-;

507

(b) IF and MIAA. Experimental conditions: [I-] = 10μM, [NH2Cl] = [HOCl] =

508

[PMS] = 100μM, [sigma HA] = [SRHA] = [SRFA] = 4mg·C/L, and reaction time

509

t= 24 h.

510 511 512 27

ACS Paragon Plus Environment

Environmental Science & Technology Letters

Page 28 of 28

25 MIAA

IF

Iodinated products (nM)

20 15

10 5 0 NH2Cl

HOCl

513

PMS

514

FIGURE 5. Formation of IF and MIAA from I- spiked natural water treated by

515

HOCl, NH2Cl, and PMS. Experimental conditions: [I-] = 0.5μM, [HOCl] = [NH2Cl]

516

= [PMS] = 15μM, DOC = 6.2mg·C/L, alkalinity = 1.3mM as HCO3-, pH=7.3, and

517

reaction time t= 24h.

518 519 520 521 522 523

TOC Art

NOM

HOI

TOI

IO3-

IPMS

28

ACS Paragon Plus Environment