Subscriber access provided by Colorado State University | Libraries

Article

Degradation of chlorotriazine pesticides by sulfate radicals and influence of organic matter Holger V. Lutze, Stephanie Bircher, Insa Rapp, Nils Kerlin, Rani Bakkour, Melanie Geisler, Clemens von Sonntag, and Torsten C Schmidt Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es503496u • Publication Date (Web): 27 Oct 2014 Downloaded from http://pubs.acs.org on October 28, 2014

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

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

Page 1 of 22

Environmental Science & Technology

1

Degradation of chlorotriazine

2

pesticides by sulfate radicals and the

3

influence of organic matter

4 5 6 7 8

Holger V. Lutze1,2*, Stephanie Bircher1, Insa Rapp1, Nils Kerlin1, Rani Bakkour1, Melanie Geisler1, Clemens von Sonntag1,3 and Torsten C. Schmidt1,2,4

9

Essen, Germany, (2) IWW Water Centre, Moritzstr. 26, D-45476 Mülheim an der Ruhr,

(1) University Duisburg-Essen, Instrumental Analytical Chemistry Universitätsstr. 5 D-45141

10

Germany, (3) Max-Planck-Institut für Bioanorganische Chemie, Stiftstrasse 34-36, P.O. Box

11

101365, D-45470 Mülheim an der Ruhr, Germany, (4) Centre for Water and Environmental

12

Research, Universitätsstraße 2, D-45117 Essen, Germany

13 14 15 16 17

e-mail: [email protected] RECEIVED DATE (to be automatically inserted after your manuscript is accepted if required according to the journal that you are submitting your paper to)

18

19

TOC art

20 1

ACS Paragon Plus Environment

Environmental Science & Technology

21

2

ACS Paragon Plus Environment

Page 2 of 22

Page 3 of 22

Environmental Science & Technology

22

Abstract: Atrazine, propazine, and terbuthylazine are chlorotriazine herbicides that have been

23

frequently used in agriculture and thus are potential drinking water contaminants. Hydroxyl

24

radicals produced by advanced oxidation processes can degrade these persistent compounds.

25

These herbicides are also very reactive with sulfate radicals (2.2–4.3 × 109 M-1 s-1). However,

26

the dealkylated products of chlorotriazine pesticides are less reactive towards sulfate radicals

27

(e.g., desethyl-desisopropyl-atrazine (DEDIA; 1.5 × 108 M-1 s-1). The high reactivity of the

28

herbicides is largely due to the ethyl- or isopropyl group. For example, desisopropyl-atrazine

29

(DIA) reacts quickly (k = 2 × 109 M-1 s-1), whereas desethyl-atrazine (DEA) reacts more

30

slowly (k = 9.6 × 108 M-1 s-1). The tert-butyl group does not have a strong effect on reaction

31

rate, as shown by the similar second order reaction rates between desethyl-terbuthylazine

32

(DET; k = 3.6 × 108 M-1 s-1) and DEDIA. Sulfate radicals degrade a significant proportion of

33

atrazine (63%) via dealkylation, in which deethylation significantly dominates over

34

deisopropylation (10:1). Sulfate- and hydroxyl radicals react at an equally fast rate with

35

atrazine (k (hydroxyl radical + atrazine) = 3 × 109 M-1 s-1). However, sulfate- and hydroxyl

36

radicals differ considerably in their reaction rates with humic acids (k (sulfate radical + humic

37

acids) = 6.6 × 103 L mg C-1 s-1; k (hydroxyl radical + humic acids) = 1.4 × 104 L mg C-1 s-1).

38

Thus, in the presence of humic acids, atrazine is degraded more efficiently by sulfate radicals

39

than by hydroxyl radicals.

40 41

KEYWORDS: CHLOROTRIAZINE PESTICIDES, DEALKYLATION, RATE CONSTANTS, SULFATE

42

RADICALS, HYDROXYL RADICALS, WATER TREATMENT

43 44

Introduction

45

Oxidation based on the sulfate radical (SO4●─) is of increasing interest as an oxidative water

46

treatment. SO4●─ is a strong one-electron oxidant, but it also readily reacts by addition to C–C

47

double bonds and by H-abstraction. Thus, it is capable of oxidizing a large number of

48

pollutants such as arsenic [1], trichloroethene [2], tert-butylmethylether [3], endosulfan [4],

49

and different derivates of phenol [5-7]. SO4●─ can be produced by photolysis [8, 9],

50

thermolysis [10], or reduction of peroxodisulfate (S2O82─) by transition metals in their low

51

oxidation states [5, 11-13]. Mechanisms of SO4●─ reactions differ somewhat from those of

52

hydroxyl radical (●OH) reactions. SO4●─ reacts more readily by electron transfer than ●OH, 3

ACS Paragon Plus Environment

Environmental Science & Technology

53

but slower by H-abstraction and addition [14, 15]. Thus, reactivity, product pattern, and

54

energy efficiency of SO4●─-based oxidation may also be different from ●OH-based oxidation.

55

For example, perfluorinated carboxylic acids are inert toward ●OH but can be degraded by

56

SO4●─ [16, 17]. This indicates that SO4●─-based oxidation may complement more common

57

(●OH-based) advanced oxidation processes.

58

The reaction of SO4●─ with atrazine has been described to be very fast (k (SO4●─ + atrazine) =

59

3 × 109 M-1 s-1 [18]), similar to the reaction rate of ●OH (k (● OH + atrazine) = 2.4–3.0 × 109

60

M-1 s-1 [19-21]). The herbicide is readily dealkylated by ● OH, however, the triazine ring

61

system also seems to be attacked, but products have not been identified [21]. In the European

62

Union, a drinking water standard of 0.1 µg L-1 has been defined for a particular pesticide and

63

0.5 µg L-1 for the sum of all pesticides. Chlorotriazine pesticides continue to be an issue for

64

the water suppliers and public health. Although the use of atrazine as a herbicide has been

65

banned in the European Union since March 2004 (commission decision 2004/248/EC &

66

commission decision 2004/247/EC), atrazine has frequently been found in European ground

67

waters even in the year 2010. The same study also reports the detection of terbuthylazine [22].

68

The mechanism of the ●OH reaction with atrazine has already been investigated in detail [19,

69

21], and a mechanistic study on the reaction of atrazine with SO4●─ was published recently

70

[23]. The present work provides a kinetic study on the reaction of different chlorotriazine

71

herbicides and their products with SO4●─ and also provides further advancements in the

72

mechanistic ideas on the reaction of atrazine with SO4●─ by providing product yields of main

73

products. Furthermore, the efficiency of SO4●─-based oxidation to degrade atrazine in the

74

presence of organic matter was investigated.

75 76

Materials and methods

77

All chemicals were commercially available and used as received: Acetaldehyde (p.a.) from

78

Sigma-Aldrich; acetone (p.a.) from KMF Laborchemie Handels GmbH; argon (5.0) from Air

79

liquid; atrazine (97.4%) from Sigma-Aldrich; 4-chlorobenzoic acid (pCBA) (99%) from

80

Aldrich; desethyl-atrazine (DEA) (99%) from Sigma-Aldrich; desethyl-desisopropyl-atrazine

81

(DEDIA) (97.8%) from Sigma-Aldrich; desisopropyl-atrazine (DIA) (95.4%) from Sigma-

82

Aldrich; 2,4-dinitrophenyl-hydrazine (50% in water, p.a.) (DNPH) from Merck; ethanol

83

(≥99.8% v/v) from Sigma-Aldrich; humic acids (Depur) (45–70%) from Carl Roth;

84

hydrochloric acid (37% in water, p.a.) from Merck; methanol (p.a.) from Sigma-Aldrich; 4

ACS Paragon Plus Environment

Page 4 of 22

Page 5 of 22

Environmental Science & Technology

85

perchloric acid (60.4% in water, p.a.) from J.T. Baker; simazine (99%) from Sigma-Aldrich;

86

sodium hydroxide (≥99.9%, p.a.) from Sigma-Aldrich; sodium peroxodisulfate (p.a.) from

87

Sigma-Aldrich; sulfuric acid (98%, p.a.) from Merck; tert-butyl alcohol (≥99%) from Merck;

88

terbuthylazine (98.8%) from Sigma-Aldrich; and desethyl-terbuthylazine (DET) (97.4%) from

89

Sigma-Aldrich.

90

SO4●─ rate constants for atrazine and the other chlorotriazines (terbuthylazine, propazine,

91

DEA, DIA, DEDIA, and DET) have been determined by competition kinetics (Refer to the

92

Supporting Information; SI) [24] using the following competitors: pCBA, k (SO4●─ + pCBA)

93

= 3.6 × 108 M-1 s-1 [25], acetanilide, k (SO4●─ + acetanilide) = 3.6 × 109 M-1 s-1 [25], and

94

atrazine, k (SO4●─ + atrazine) = 3 × 109 M-1 s-1 [18]. The caption for Table 1 indicates the

95

choice of competitors for determining the different rate constants. SO4●─ were generated by

96

thermolysis of peroxodisulfate (S2O82–) at 60°C. The second order reaction rates obtained at

97

this temperature correspond to the temperature at which the reference compound was

98

determined (refer to the SI). The reaction solutions were adjusted to pH 7 with sodium

99

hydroxide. Unless otherwise specified, buffers were not added because they can also react

100

with SO4●─ (e.g., k (HPO42– + SO4●─) = 1.2 × 106 M-1 s-1 [15]). This reaction leads to the

101

formation of HPO4●─, which could also react with the chlorotriazines but at a different rate.

102

Similar problems may arise with other buffers.

103

In a SO4●─-based process, solutions slowly acidify because in the reaction of SO4●─ plus

104

triazines, protons could be released (see below). This can be problematic since protonation of

105

the substrates may change both the reaction kinetics and the mechanism. Thus, we maintained

106

the pH at least 1 pH unit above the highest pKa value of the corresponding substrates.

107

Because SO4●─ normally reacts slower with protonated species than with deprotonated species

108

(compare reaction rates compiled in Neta et al. [15]), the maximal 10% contribution of these

109

compounds in our study can be neglected. In the present experiments, the reaction time was

110

sufficiently short so that the pH did not drop below this value. After heating up the sample,

111

S2O82– was added to initiate the reaction (S2O82─ dose was 1 mM). After different reaction

112

times, aliquots were withdrawn and chilled on ice. Methanol was immediately added (1 M in

113

the sample, ≈4% v/v) for scavenging the SO4●─ that could be formed at low rates during

114

storage time. It was observed that small amounts of chloride (2–30 µM) were present in all

115

experiments. Since concentrations of Cl─ in the µM range can be introduced into samples by

116

several means such use of standard electrodes for determination of pH (KCl as background 5

ACS Paragon Plus Environment

Environmental Science & Technology

117

electrolyte) and impurities in chemicals, it is very difficult to avoid the presence of Cl─. SO4●─

118

reacts moderately fast with Cl─ (see below), and in that reaction ●OH are formed at neutral pH

119

[26]. For minimizing a bias by ● OH in the competition experiments, tert-butanol was added.

120

By that means, formation of ●OH was suppressed since ●OH reacts about 1000 times faster

121

with tert-butanol than SO4●─ (k (SO4●─) = 4.0–9.5 × 105 M-1 s-1 [27, 28], k (●OH) = 6 × 108 M-

122

1

123

and 10 mM), while 5 µM of the triazines and competitors were added. In the case of atrazine

124

(2–3 × 109 M-1 s-1 = k (●OH) [19, 21] = k(SO4●─) [18]), ●OH are largely scavenged by tert-

125

butanol >95% (with 1 and 10 mM tert-butanol), whereas SO4●─ are scavenged by only 3–6%

126

(at 1 mM tert-butanol) and 20–40% (at 10 mM tert-butanol) (calculated by competition

127

kinetics (refer to the SI) and the reaction rates given above). In the case of the product studies,

128

tert-butanol was not added, so as to prevent interferences in the product pattern. However, the

129

substrate (atrazine) was dosed in higher concentrations (20–25 µM) compared with the

130

competition kinetic experiments. Since atrazine reacts faster with SO4●─ than Cl─, the reaction

131

of SO4●─ plus Cl─ is suppressed (average value obtained from rate constants provided by [30]

132

is k (Cl─ + SO4●─) = 3.3 × 108 M-1 s-1). Given a maximal Cl─ concentration of 30 µM (see

133

above), no more than 14% of the SO4●─ reacted with Cl─ yielding ●OH (calculated by

134

competition kinetics). As will be explained below, this did not influence the outcome of

135

product formation in the primary reaction of SO4●─ plus atrazine.

136

Photochemical experiments were conducted using a merry-go-round photo reactor according

137

to Wegelin et al. [31] equipped with quartz tubes for irradiating the sample (radiation source

138

GPH303T5L/4, 15 W, Heraeus Noble Light; refer to the SI). Fluence rates were determined

139

by atrazine actinometry in solutions containing humic acids according to Canonica et al. [32]

140

and by uridine actinometry in pure water according to von Sonntag and Schuchmann [33].

141

All compounds were separated by HPLC on a C-18 reversed phase column and detected by

142

UV absorption. The following HPLC-systems were used: 1) Shimadzu: liquid chromatograph

143

LC 20-AT, UV/Vis detector SPD 20A, auto sampler SIL 20A, degasser DGU-20A5,

144

communication bus module CBM 20A, and column oven CTO-10AS vp. 2) Shimadzu: liquid

145

chromatograph LC-10AT vp, diode array detector SPD M10A vp, auto sampler SIL 10 AD

146

vp, degasser DGU 14A, system controller SCL-10A vp, and column oven CTO-10AS vp. 3)

147

Agilent 1100 Series: quaternary pump G1311A, UV-Vis detector VWD G1314A, autosampler

148

ALS G1313A, degasser G1379A, and column oven Colom G1316A.

s-1 [29]). For the individual experiments, two concentrations of tert-butanol were dosed (1

6

ACS Paragon Plus Environment

Page 6 of 22

Page 7 of 22

Environmental Science & Technology

149

For separation of the analytes reversed phase C-18 columns were used (5 µm particle size,

150

250 × 4.6 mm from Knauer or Bischoff). For the mobile phase, various isocratic mixtures and

151

gradients of water and methanol as well as water and acetonitrile were used. All compounds

152

were determined by UV absorption at their maxima (chlorotriazine herbicides: 210–225 nm,

153

pCBA: 235–239 nm, and acetanilide: 224 nm).

154

Acetone and acetaldehyde were derivatized with 2,4-dinitrophenylhydrazine [34] and the

155

hydrazones quantified by HPLC-UV. The experiments were carried out in air-saturated

156

solution.

157

Cl─ was analyzed by ion chromatography (Metrohm 883 basic) equipped with a conductivity

158

detector coupled with ion suppression (anion separation column with quaternary ammonium

159

groups: Metrosep A Supp 4 - 250/4.0 mm, particle size 9 µm; eluent: a solution of HCO3─

160

(1.7-1.9 mM), CO32─ (1.8 mM) was mixed with acetonitrile (70% aqueous buffer : 30 %

161

acetonitrile); flow: 1 mL min-1).

162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177

7

ACS Paragon Plus Environment

Environmental Science & Technology

178

Results and discussion

179

Kinetics. The rate constants for the reaction of SO4●─ and ●OH with the investigated triazine

180

herbicides and their dealkylation products are compiled in Table 1.

181 182

Table 1: Rate constants for the reaction of SO4●─ and ●OH with s-triazines and competitors. n

183

= number of replicates. SO4●─ / M-1 s-1 (4.2 ± 0.20) × 109 a) n = 4 (4.7 ± 0.10) × 109 c) n = 4 Average: 4.5 ± 0.30 × 109 3 × 109 [18] 2.6 × 109 [23] 1.4 × 109 [35] (2.0 ± 0.57) × 109 a) n = 4



(9.6 ± 0.17) × 108 a) n = 4

1.2 × 109 [20]

(1.5 ± 0.08) × 108 a) n = 4

isopropyl- >> tert-butyl group attached to the triazine ring.

201

This indicates that in case of triazine herbicides, N-ethyl and to a somewhat lesser extent the

202

N-isopropyl group is a key factor in the reactivity of SO4●─ towards s-triazine-herbicides. In

203

that it finds its analogy in ●OH reactions, albeit SO4●─ display a more distinctive prevalence to

204

the ethyl function.

205

However, SO4●─ react remarkably faster with DEDIA than ●OH.

). Yet, the reaction of SO4●─ plus DEA displayed a distinctive decrease in rate, albeit it was N-isopropyl group, because the reactivity of DET and DEDIA was

206 207

Product formation. Similar to the studies of Kahn et al. [23], the reaction of SO4●─ with

208

atrazine led to dealkylation and thereby produced DEA and DIA as well the corresponding

209

products acetaldehyde and acetone (Figure 1).

210

9

ACS Paragon Plus Environment

Environmental Science & Technology

Atrazinkonzentration / µM DEA DIA Acetaldehyd Aceton

25

Concentration / µM

Page 10 of 22

20 15 10 5 0 0

5

10

15

20

25

30

35

Time / min 211 212 213 214

Figure 1: Degradation of atrazine by SO4●─ and product formation in air-saturated aqueous solution. Initial pH = 7. [S2O82–] = 1 mM. T = 60°C. Atrazine (squares), acetaldehyde (open circles), DEA (filled circles), DIA (filled triangles), and acetone (open triangles).

215 ●

216

The formation of these products has also been observed as a result of

217

degradation of atrazine [21]. There is, however, a remarkable difference. In the case of ●OH,

218

the DEA to DIA ratio can be calculated to be ≈3, while with SO4●─ it was ≈10 (Table 2).

219

Acetone at low turnovers was too uncertain to be reported due to a fluctuating background of

220

acetone in the samples. At higher turnovers, acetone was mainly formed as a secondary

221

product from SO4●─-induced dealkylation of the major product DEA (Figure 1). In contrast,

222

acetone was formed substantially in primary reactions of ●OH with atrazine [21] (Table 2).

223

Due to the possible presence of chloride in the µM range, one has to consider a contribution

224

of ●OH (see above). As explained above ●OH could contribute to atrazine degradation by

225

14%. When we consider that the reaction of ●OH plus atrazine yields 51% DEA and 20%

226

DIA, then a 14% contribution of ●OH in atrazine degradation yields a maximum of 7% DEA

227

and 3% DIA. A similar situation applies for yields of acetaldehyde and acetone. These minor

228

influenc was neglected.

229 10

ACS Paragon Plus Environment

OH-induced

Page 11 of 22

230 231 232 233

Environmental Science & Technology

Table 2: Product yields (%) of atrazine degradation in the SO4●─ and ●OH reactions. The primary yields were derived from a plot of atrazine degradation versus product formation by a linear regression (n = 4 replicates). *Detected but not quantified due to fluctuating background contamination. Product

SO4●─

Acetaldehyde Acetone DEA DIA Mass balance

57 ± 11% * 57 ± 4% 6 ± 0.8% 63%

OH [21] 53% 20% 51% 20% 71%

(DEA+DIA)

(DEA+DIA)

≈10

≈3

Ratio DEA:DIA



234 235

Desisopropylation occured to a minor extent (≈6%) in primary reactions of atrazine, reflecting

236

the low concentration of DIA during atrazine degradation (Figure 1). The cleavage of the N-

237

ethyl group was largely favored over the cleavage of the N-isopropyl group.

238

Product yields shown in Table 2 indicate that ≈40% of the SO4●─ reactions gave rise to

239

unidentified products. Such an incomplete mass balance has also been reported for ●OH

240

reactions with atrazine [21] and in purine free-radical chemistry [14]. For the purines, N-

241

centered and O-centered radicals that do not react with O2 play a major role, and in the

242

present study it is envisaged that these produced a large number of dimers that escaped

243

detection. It is conceivable that the missing fraction in our system was due to the production

244

of such dimers. With three ring nitrogens and two exocyclic nitrogens, the number of

245

potential dimers is very large; hence, the yield of each individual dimer may be very low and

246

result in concentrations below the detection limit. A compilation of other possible degradation

247

products in the missing fraction can be found in Kahn et al. [23].

248 249

Reaction mechanism. In the reaction of SO4●─ with electron-rich aromatic compounds,

250

electron transfer is a preferred reaction pathway. In the case of methoxy derivatives of

251

benzene, the corresponding radical cations are generated as primary short-lived intermediates

252

[36]. With aromatic carboxylic acids, the formation of phenyl radicals was observed by ESR,

253

pointing to a decarboxylation of the originally formed radical cations [36]. For benzene

254

lacking electron-donating groups, hydroxycyclohexadienyl radicals have been observed [37].

255

A rapid reaction of water with the originally formed radical cations may account for this. The 11

ACS Paragon Plus Environment

Environmental Science & Technology

256

SO4●─ reacts with these benzene derivates as fast as it reacts with the triazine herbicides (≥1 ×

257

109 M-1 s-1 [25]), and for atrazine it was suggested that SO4●─ also reacts via electron transfer

258

[18, 23]. Based on this suggestion, a mechanism can be proposed resulting in dealkylation

259

along with formation of carbonylic compounds and free amines. Furthermore, the below

260

mechanistic prospects provide a possible explanation for the prevalence of deethylation in the

261

reaction of SO4●─ plus atrazine. In the first step, electron transfer yields a radical cation

262

(reaction 1).

263

264 265 266

Kahn et al. [23] supposed one possible pathway to be the substitution of the chlorine atom by

267

a hydroxyl group. However, products detected in the present study indicate that another

268

reaction might occur. In the present study it is proposed that the primary radical cation will be

269

in equilibrium with the corresponding N-centered radicals (reactions 2 and 4). The N-centered

270

radicals are expected to undergo a (water-catalyzed) 1,2-H shift analogous to the well-

271

documented 1,2-H shift of alkoxyl radicals [38], thereby forming C-centered radicals

12

ACS Paragon Plus Environment

Page 12 of 22

Page 13 of 22

Environmental Science & Technology

272

(reactions 3 and 5). The N-centered radicals do not react with O2, but the C-centered ones

273

react rapidly (reaction 6, k = 3 × 109 M-1 s-1) [21].

274 275

The resulting peroxyl radical subsequently loses O2●─ (reactions 7 and 8) or HO2● (reaction

276

10) [21]. The subsequent acid and base catalyzed hydrolysis of the imine yields the carbonylic

277

compound and the free amine (reaction 10) [19, 21].

278

The N–H bond at the N-ethyl group is more acidic than that at the N-isopropyl group and

279

deprotonation will occur more preferably at the H–N-ethyl site (reaction 4). Hence, the

280

formation of the C-centered radical and the subsequent reactions (see above) are favored at

281

the ethyl side group. That is a conceivable explanation for the prevalence of deethylation in

282

the reaction of SO4●─ with atrazine.

283

The main reaction pathway of ●OH is abstraction of hydrogen located at the α-position of the

284

alkyl groups (preferably at the ethyl group), thus resulting in a C-centered radical that is also

285

formed in reactions 3 and 5 [21]. Kahn et al. [23] suggested that SO4●─ could also react via H-

286

abstraction, in principle. However, reaction rates of H-abstraction by SO4●─ typically have a

287

range of 107–108 M-1 s-1. Thus, the high reaction rate of >109 M-1 s-1 for atrazine, DIA,

288

terbuthylazine, and propazine would be unusual. Furthermore, H-abstraction cannot explain 13

ACS Paragon Plus Environment

Environmental Science & Technology

289

the strong prevalence for deethylation of atrazine. With respect to H-abstraction, the side

290

groups of atrazine (ethylamine- and isopropylamine group) are comparable to ethanol and 2-

291

propanol that react with SO4●─ via H-abstraction at the α-C-atom [27]. Eibenberger at al. [27]

292

determined the reaction rates of SO4●─ plus 2-propanol (k = 3.2 × 107 M-1 s-1) to be faster than

293

SO4●─ plus ethanol (k = 1.6 × 107 M-1 s-1). This indicates that if the SO4●─ reacts with atrazine

294

by H-abstraction, deisopropylation would be the dominant pathway instead of the observed

295

strong prevalence for deethylation.

296

Atra-imine from reaction 8 was observed in the present HPLC-DAD measurements. When the

297

analysis occurred ≤2 hours after the experiment, atra-imine was eluted as a dominant product

298

peak at a retention time between atrazine and DEA (Figure S2 (SI). For a quantitative view on

299

the evolution of atra-imine, its response factor was determined (refer to the SI).

300

The evolution of atra-imine in atrazine degradation is shown in Figure 2. Even though atra-

301

imine was considered to be the precursor of DEA, the latter compound appeared in all

302

samples. This can be explained by partial hydrolysis of the imine before HPLC determination

303

began (an approximate 2–3 h delay between experiment and analysis). The full yield of DEA

304

was also determined after complete turnover of the atra-imine (i.e., after ≈72 h). The sum of

305

atra-imine and DEA determined in the measurements started after 2–3 h (see above)

306

resembled the DEA yield after complete conversion (measurement 72 h after the experiment,

307

Figure 2 inset).

308

14

ACS Paragon Plus Environment

Page 14 of 22

Page 15 of 22

Environmental Science & Technology

25

12 y = 1.04x - 0.5

DEA (72 h)

10

Concentration / µM

20

8 6 4 2

15

0 0 10

2

4 6 8 10 DEA + Atra-imine

12

5

0 0

10

20

30

40

50

60

Time / min

309 310 311 312 313 314

Figure 2: Degradation of atrazine (open circles) and formation of products: atra-imine (gray circles), DEA (black triangles), and DEDIA (open triangles). Formation of DEA after 72 h (gray squares). Thermal activation of S2O82─, T = 60°C, initial pH 7. [atrazine]0 = 25 µM. [S2O82─]0 = 1 mM. Inset: correlation of the sum of atra-imine and DEA of the first measurement versus DEA formation after 72 h.

315 316

The hydrolysis of isopropyl imine was much faster than atra-imine, and immediately after the

317

reaction, DIA was present in full yield. The reason for the difference in hydrolysis rates of the

318

two imines is not fully understood. Full conversion of the imine to DEA was assured when

319

determining the total yields of desethyl products (Table 2).

320

The above mechanistic considerations can explain why in SO4●─-induced dealkylation of

321

atrazine, formation of DEA plus acetaldehyde was favored over the formation of DIA plus

322

acetone. Yet, in the ●OH-induced reaction, such a preference was also observed, albeit to a

323

smaller extent [21]. This requires that ●OH preferably abstract hydrogen at the α–C of the

324

ethyl side group than the isopropyl group.

325 326

Reactions in the presence of humic acids. Figure 3 shows the degradation of atrazine in the

327

presence of humic acids under conditions of direct photolysis, UV/H2O2, and UV/S2O82–.

328

Degradation by the UV/S2O82– treatment was more efficient than by UV/H2O2, and 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 22

329

considerably more efficient than direct photolysis. The reason for this is the higher rate

330

constant of the reaction ●OH plus dissolved organic matter (DOM) compared with the

331

reaction SO4●─ plus DOM (see below).

concentration / µM

0.5 0.4 0.3 0.2 0.1 0.0 0

2

4

6

8

10

12

14

16

18

Time / min 332 333 334 335 336 337 338 339 340 341

Figure 3: Photochemical degradation of atrazine (0.5 µM) in the presence of humic acids (15 mg C L-1). [S2O82–] = [H2O2] = 1 mM. [phosphate] = 2.5 mM (a contribution of phosphate radicals in this system can be ruled out (refer to the SI). pH = 7.2. Radiation source: low pressure mercury lamp (15 W), Fluence rate (254 nm) = 58 µEinstein m-2 s-1 in pure water and 8.7 µEinstein m-2 s-1 in presence of humic acids (A254 = 1.65 cm-1). Direct photolysis (crosses). UV/H2O2 (closed circles). UV/S2O82– (open circles). Curved lines represent calculated best fit of atrazine degradation for k (SO4●─ plus DOC) and k (●OH plus DOC) according to equation 1 (refer to the SI). Symbols represent measured data. Observed degradation kinetics of atrazine: UV/S2O82–: 2.0 × 10-3 s-1; UV/H2O2: 8.8 × 10-4 s-1; UV: 3.6 × 10-4 s-1.

342 343

According to equation 1, the second order reaction rate of the reaction SO4●─ / ●OH plus

344

humic acids can be calculated using the first order degradation rate of atrazine given the

345

following information: fluence rate, first order degradation of atrazine by direct photolysis,

346

reaction rates with the corresponding radical plus atrazine, and the concentration of humic

347

acids (expressed as mg C L-1).

348 16

ACS Paragon Plus Environment

Page 17 of 22

Environmental Science & Technology

349

Equation 1:

350

k´(atrazine) =

351

Where k´(atrazine ) is the first order degradation rate of atrazine / s-1, ε peroxide is the molar

352

absorption of H2O2 or S2O82─ / m2 mol-1, φ peroxide is the quantum yield of H2O2 or S2O82─ / mol

353

Einstein-1, H is the fluence rate / Einstein m-2 s-1, [ peroxide] is the concentration of H2O2 or

354

S2O82─ / M, k ( atrazine + radical ) is the second order rate constant of atrazine plus ● OH or

355

SO4●─ / M-1 s-1, k ( photolysis of atrazine ) is the rate of atrazine photolysis / s-1, [DOC ] is the

356

concentration of DOC / mg C L-1, k ( DOC + radical ) is the rate constant of the reaction ●OH

357

or SO4●─ plus DOC / L mg-1 C s-1.

358

The k (DOC + radical) is the only unknown and can either be derived by calculating the first

359

order degradation rate of atrazine (from experimental data) and solving the above equation for

360

k (DOC + radical), or by adjusting k (DOC + radical) and arriving at the best fit for the

361

experimentally determined degradation of atrazine. The latter method was used in the present

362

study for different DOC and peroxide concentrations (refer to the SI). Consequently, the

363

following rate constants were derived from the three individual experiments (concentration of

364

humic acids were 7.5 and 15 mg C L-1): k (SO4●─ + humic acids) = 6.6 ± 0.4 × 103 L mg C -1

365

s-1, and k (● OH + humic acids) = 1.4 ± 0.2 × 104 L mg C -1 s-1.

366

The value determined for ● OH was in accordance with a value determined for Suwannee

367

River humic acids (k = 1.9 × 104 L mg C-1 s-1 [39]; note that an average value for the reaction

368

of ●OH with NOM of 2.5× 104 L mg C-1 s-1 was reported by Schwarzenbach et a. [40]). This

369

confirms our experimental concept. Beside the lower reaction rate of SO4●─ toward DOC, the

370

higher molar absorption coefficient of S2O82─ and the higher quantum yield for radical

371

formation (refer to the SI) further increased the efficiency of atrazine degradation in

372

UV/S2O82─.

373

Beside DOC, HCO3─ is an important water matrix constituent, which might limit the

374

oxidation capacity of a radical-based oxidation system by scavenging (reaction rates of ●OH

375

and SO4●─ are given below).

376

Figure 4 compares the first order degradation rates of atrazine in UV/H2O2 with UV/S2O82─ in

377

the presence of humic acids at different concentrations of HCO3─ (2.5–10 mM). Atrazine was

2.303× ε peroxide × φ peroxide × H × [ peroxide ]× (k (atrazine + radical ) − k ( photolysis of atrazine))

[DOC]× k (DOC + radical )

17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 22

378

degraded twice as fast in UV/S2O82─ than in UV/H2O2, regardless of the presence of HCO3─.

379

This independence from alkalinity is further illustrated in Figure 4 b), showing that the ratios

380

of degradation rates in both processes do not correlate with the HCO3─ concentration. This is

381

in accordance with the similar reaction rates of the reactions ●OH and SO4●─ plus HCO3─

382

reported in the literature, cancelling the influence of HCO3─ ((k (●OH + HCO3─) = 0.85–1 ×

383

107 M-1 s-1 [29], recommended value 1 × 107 M-1 s-1[40], k (SO4●─ + HCO3─) = 2.8–9.1 × 106

384

M-1 s-1 [41, 42]).

385 0

3

a)

-0.5

b) 2.5 kobs(S2O82-)/kobs(H2O2)

ln(c/c0) (UV/S2O82-

-1 -1.5 -2 -2.5 -3 -3.5

y = 1.92x - 0.08

-4

2 1.5 1 0.5

-4.5 -5

0 -2.2

-1.7

-1.2

-0.7

-0.2

0

ln(c/c0) (UV/H2O2)

2

4 [HCO3─]

6

8

10

/ mM

386 387 388 389 390 391 392

Figure 4: a) Observed degradation rates of atrazine in UV/H2O2 and UV/S2O82─ in the presence of humic acids and different concentrations of HCO3─. b) Ratio of first order degradation rates for atrazine versus HCO3─ concentration (derived from a)). 15 mg C L-1 (circles). 7.5 mg C L-1 (triangles). [HCO3─] = 0, 2.5, 5, and 10 mM. [phosphate] = 2.5 mM in the case of experiments with 15 mg C L-1, and 1.25 mM in case of experiments with 7.5 mg C L-1. pH = 7.2, T = 25°C.

393 394

The lower reactivity of the organic matter in SO4●─ reactions compared with ●OH reactions

395

can be explained by the tendency that H-abstraction reactions in SO4●─ are comparatively

396

slow. The reactions of several alcohols, alkanes, and ethers with SO4●─ proceed via H-

397

abstraction, with reaction rates in the range of 107–108 M-1 s-1 [43, 44]. In contrast, H-

398

abstraction is about one to two orders of magnitude faster for ●OH reactions (compare with

399



400

react faster with ●OH but slower with SO4●─. This leads to a weaker scavenging of radicals,

OH rate constants in Buxton et al. [29]). Thus, saturated moieties of the organic matter might

18

ACS Paragon Plus Environment

Page 19 of 22

Environmental Science & Technology

401

which could improve the degradation efficiency of pollutants. However, it has to be

402

mentioned that the higher selectivity of SO4●─ could be counterbalanced in the case of target

403

compounds reacting faster with ●OH than with SO4●─.

404 405

The present study shows that the mode of chlorotriazine degradation in SO4●─-based

406

oxidation is somewhat similar to ●OH (dealkylation), despite different predominant reaction

407

mechanisms (H-abstraction versus electron transfer). Furthermore, SO4●─ might even degrade

408

DEDIA, a final product in common advanced oxidation processes, because the reaction is

409

about one order of magnitude faster than with ●OH. The higher selectivity of SO4●─ in the

410

presence of organic matter and HCO3─ further indicates that the degradation of DEDIA by

411

SO4●─ might be possible. However, the energy demand has to be carefully estimated under

412

water purification conditions to allow meaningful comparisons with other oxidative treatment

413

options for water treatment, such as ozonation.

414 415

Acknowledgements

416

All authors are deeply thankful for the expert advice of Prof. Clemens von Sonntag. His great

417

knowledge, vast experience, and brilliant creativity left its mark in the present and in

418

countless other studies. He is a noble person and remains our ideal as both a scientist and a

419

human being. H.V.L. thanks the German Water Chemistry Society for generous financial

420

support.

421 422

References

423 424 425 426 427 428 429 430 431 432 433 434

1.

2.

3.

4.

Neppolian, B.; Celik, E.; Choi, H., Photochemical oxidation of arsenic(III) to arsenic(V) using peroxydisulfate ions as an oxidizing agent. Environ. Sci. Technol., 2008. 42(16): 6179-6184. Liang, C.; Lee, I.L.; Hsu, I.Y.; Liang, C.P.; Lin, Y.L., Persulfate oxidation of trichloroethylene with and without iron activation in porous media. Chemosphere, 2008. 70(3): 426-435. George, C.; El Rassy, H.; Chovelon, J.M., Reactivity of selected volatile organic compounds (VOCs) toward the sulfate radical (SO4●─). Int. J. Chem. Kinet., 2001. 33(9): 539-547. Shah, N.S.; He, X.; Khan, H.M.; Khan, J.A.; O'Shea, K.E.; Boccelli, D.L.; Dionysiou, D.D., Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: A comparative study. J. Hazard. Mater., 2013. 263: 584-592.

19

ACS Paragon Plus Environment

Environmental Science & Technology

435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481

5.

6. 7.

8.

9.

10. 11.

12. 13. 14. 15. 16.

17.

18.

19.

20.

Anipsitakis, G.P.; Dionysiou, D.D.; Gonzalez, M.A., Cobalt-mediated activation of peroxymonosulfate and sulfate radical attack on phenolic compounds. Implications of chloride ions. Environ. Sci. Technol., 2006. 40(3): 1000-1007. Lin, Y.T.; Liang, C.; Chen, J.H., Feasibility study of ultraviolet activated persulfate oxidation of phenol. Chemosphere, 2011. 82(8): 1168-1172. Fang, G.D.; Dionysiou, D.D.; Zhou, D.M.; Wang, Y.; Zhu, X.D.; Fan, J.X.; Cang, L.; Wang, Y.J., Transformation of polychlorinated biphenyls by persulfate at ambient temperature. Chemosphere, 2013. 90(5): 1573-1580. Herrmann, H., On the photolysis of simple anions and neutral molecules as sources of O─/OH, SOx─ and Cl in aqueous solution. Phys. Chem. Chem. Phys., 2007. 9(30): 3935-3964. Mark, G.; Schuchmann, M.N.; Schuchmann, H.P.; von Sonntag, C., The photolysis of potassium peroxodisulphate in aqueous solution in the presence of tert-butanol: a simple actinometer for 254 nm radiation. J. Photochem. Photobiol., A, 1990. 55(2): 157-168. Liang, C.; Su, H.W., Identification of sulfate and hydroxyl radicals in thermally activated persulfate. Ind. Eng. Chem. Res., 2009. 48(11): 5558-5562. Anipsitakis, G.P.; Dionysiou, D.D., Degradation of organic contaminants in water with sulfate radicals generated by the conjunction of peroxymonosulfate with cobalt. Environ. Sci. Technol., 2003. 37(20): 4790-4797. Anipsitakis, G.P.; Stathatos, E.; Dionysiou, D.D., Heterogeneous activation of oxone using Co3O4. J. Phys. Chem. B, 2005. 109(27): 13052-13055. Anipsitakis, G.P.; Dionysiou, D.D., Radical generation by the interaction of transition metals with common oxidants. Environ. Sci. Technol., 2004. 38(13): 3705-3712. von Sonntag, C., Free-radical-induced DNA damage and its repair - A chemical perspective. Springer Verlag Berlin Heidelberg, ISBN: 3-540-26120-6, 2005. Neta, P.; Huie, R.E.; Ross, A.B., Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data., 1988. 17(3): 1027-1040. Hori, H.; Hayakawa, E.; Einaga, H.; Kutsuna, S.; Koike, K.; Ibusuki, T.; Kiatagawa, H.; Arakawa, R., Decomposition of environmentally persistent perfluorooctanoic acid in water by photochemical approaches. Environ. Sci. Technol., 2004. 38(22): 61186124. Schröder, H.F.; Meesters, R.J.W., Stability of fluorinated surfactants in advanced oxidation processes - A follow up of degradation products using flow injection-mass spectrometry, liquid chromatography-mass spectrometry and liquid chromatographymultiple stage mass spectrometry. J. Chromatogr. A, 2005. 1082(1 SPEC. ISS.): 110119. Manoj, P.; Prasanthkumar, K.P.; Manoj, V.M.; Aravind, U.K.; Manojkumar, T.K.; Aravindakumar, C.T., Oxidation of substituted triazines by sulfate radical anion (SO4●─) in aqueous medium: A laser flash photolysis and steady state radiolysis study. J. Phys. Org. Chem., 2007. 20(2): 122-129. Acero, J.L.; Stemmler, K.; von Gunten, U., Degradation kinetics of atrazine and its degradation products with ozone and OH radicals: A predictive tool for drinking water treatment. Environ. Sci. Technol., 2000. 34(4): 591-597. De Laat, J.; Chramosta, N.; Dore, M.; Suty, H.; Pouillot, M., Rate constants for reaction of hydroxyl radicals with some degeradation by-products of atrazine by O3 or O3/H2O2. Environ. Technol., 1994. 15(5): 419-428. 20

ACS Paragon Plus Environment

Page 20 of 22

Page 21 of 22

482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528

Environmental Science & Technology

21. 22.

23.

24. 25.

26. 27.

28.

29.

30.

31.

32.

33.

34.

35.

36.

Tauber, A.; von Sonntag, C., Products and kinetics of the OH-radical-induced dealkylation of atrazine. Acta Hydroch. Hydrob., 2000. 28(1): 15-23. Loos, R.; Locoro, G.; Comero, S.; Contini, S.; Schwesig, D.; Werres, F.; Balsaa, P.; Gans, O.; Weiss, S.; Blaha, L.; Bolchi, M.; Gawlik, B.M., Pan-European survey on the occurrence of selected polar organic persistent pollutants in ground water. Water Res., 2010. 44(14): 4115-4126. Khan, J.A.; He, X.; Shah, N.S.; Khan, H.M.; Hapeshi, E.; Fatta-Kassinos, D.; Dionysiou, D.D., Kinetic and mechanism investigation on the photochemical degradation of atrazine with activated H2O2, S2O82- and HSO5. Chem. Eng. J., 2014. 252: 393-403. Peter, A.; von Gunten, U., Oxidation kinetics of selected taste and odor compounds during ozonation of drinking water. Environ. Sci. Technol., 2007. 41(2): 626-631. Neta, P.; Madhavan, V.; Zemel, H.; Fessenden, R.W., Rate constants and mechanism of reaction of sulfate radical anion with aromatic compounds. J. Am. Chem. Soc., 1977. 99(1): 163-164. McElroy, W.J., A laser photolysis study of the reaction of SO4- with Cl- and the subsequent decay of Cl2- in aqueous solution. J. Phys. Chem., 1990. 94(6): 2435-2441. Eibenberger, H.; Steenken, S.; O'Neill, P.; Schulte-Frohlinde, D., Pulse radiolysis and electron spin resonance studies concerning the reaction of SO4-·with alcohols and ethers in aqueous solution. J. Phys. Chem., 1978. 82(6): 749-750. Hayon, E.; Treinin, A.; Wilf, J., Electronic spectra, photochemistry, and autoxidation mechanism of the sulfite-bisulfite-pyrosulfite systems. SO2-, SO3-, SO4-, and SO5radicals. J. Am. Chem. Soc., 1972. 94(1): 47-57. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B., Critical review of rate constants for reactions of hydrated electrones, hydrogen-atoms and hydroxyl radicals (•OH/O•-) in aqueous solution. J. Phys. Chem. Ref. Data, 1988. 17(2): 513-886. NIST. National institute of standards and technology - data base of reaction rate constants. [accessed 2013]; Available from: http://kinetics.nist.gov/solution/SearchForm. Wegelin, M.; Canonica, S.; Mechsner, K.; Fleischmann, T.; Pesaro, F.; Metzler, A., Solar water disinfection: Scope of the process and analysis of radiation experiments. J..Water SRT - Aqua, 1994. 43(4): 154-169. Canonica, S.; Meunier, L.; von Gunten, U., Phototransformation of selected pharmaceuticals during UV treatment of drinking water. Water Res., 2008. 42(1-2): 121-128. von Sonntag, C.; Schuchmann, H.P., UV disinfection of drinking water and byproduct formation - some basic considerations. J. Water SRT - Aqua, 1992. 41(2): 6774. Vogel, M.; Büldt, A.; Karst, U., Hydrazine reagents as derivatizing agents in environmental analysis - A critical review. Fresenius J. Anal. Chem., 2000. 366(8): 781-791. Azenha, M.E.D.G.; Burrows, H.D.; Canle, L.M.; Coimbra, R.; Fernández, M.I.; García, M.V.; Rodrigues, A.E.; Santaballa, J.A.; Steenken, S., On the kinetics and energetics of one-electron oxidation of 1,3,5-triazines. Chem. Commun., 2003. 9(1): 112-113. O'Neill, P.; Steenken, S.; Schulte-Frohlinde, D., Formation of radical cations; of methoxylated benzenes by reaction with OH radicals, Tl2+, Ag2+, and SO4●─ in 21

ACS Paragon Plus Environment

Environmental Science & Technology

529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551

37.

38.

39.

40. 41. 42. 43. 44.

aqueous solution. An optical and conductometric pulse radiolysis and in situ radiolysis electron spin resonance study. J. Phys. Chem., 1975. 79(25): 2773-2779. Norman, R.O.C.; Storey, P.M.; West, P.R., Electron spin resonance studies. Part XXV. Reactions of the sulphate radical anion with organic compounds. J. Chem. Soc. B: Phys. Org., 1970: 1087-1095. Konya, K.G.; Paul, T.; Lin, S.; Lusztyk, J.; Ingold, K.U., Laser flash photolysis studies on the first superoxide thermal source. First direct measurements of the rates of solvent-assisted 1,2-hydrogen atom shifts and a proposed new mechanism for this unusual rearrangement. J. Am. Chem. Soc., 2000. 122(31): 7518-7527. Goldstone, J.V.; Pullin, M.J.; Bertilsson, S.; Voelker, B.M., Reactions of hydroxyl radical with humic substances: Bleaching, mineralization, and production of bioavailable carbon substrates. Environ. Sci. Technol., 2002. 36(3): 364-372. Schwarzenbach, R.P.; Gschwend, P.M.; Imboden, D.M., eds. Environmental organic chemistry. John Wiley & Sons, Inc. ISBN 0-47 1-35750-2. 2003. Huie, R.E.; Clifton, C.L., Temperature dependence of the rate constants for reactions of the sulfate radical, SO4-, with anions. J. Phys. Chem., 1990. 94(23): 8561-8567. Dogliotti, L., Flash photolysis of persulfate ions in aqueous solutions. Study of the sulfate and ozonide radical ions. J. Phys. Chem., 1967. 71(8): 2511-2516. Clifton, C.L.; Huie, R.E., Rate constants for hydrogen abstraction reactions of the sulfate radical, SO4-. Alcohols. Int. J. Chem. Kinet., 1989. 21(8): 677-687. Huie, R.E.; Clifton, C.L., Rate constants for hydrogen abstraction reactions of the sulfate radical, SO4-. Alkanes and ethers. Int. J. Chem. Kinet., 1989. 21(8): 611-619.

22

ACS Paragon Plus Environment

Page 22 of 22

Degradation of chlorotriazine pesticides by sulfate radicals and the influence of organic matter.

Atrazine, propazine, and terbuthylazine are chlorotriazine herbicides that have been frequently used in agriculture and thus are potential drinking wa...
564KB Sizes 1 Downloads 6 Views