Can Carbamates Undergo Radical Oxidation in the ... - ACS Publications

Nov 24, 2017 - to investigate the radical oxidation of carbamate pesticides, with carbaryl and carbofuran taken as a case study, and their possible ef...
0 downloads 0 Views 771KB Size
Subscriber access provided by READING UNIV

Article

Can carbamates undergo radical oxidation in the soil environment? A case study on carbaryl and carbofuran Irmina #wiel#g-Piasecka, Maciej Witwicki, Maria Jerzykiewicz, and Julia Jezierska Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b03386 • Publication Date (Web): 24 Nov 2017 Downloaded from http://pubs.acs.org on November 27, 2017

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

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

Page 1 of 40

Environmental Science & Technology

1

Can carbamates undergo radical oxidation in the

2

soil environment? A case study on carbaryl and

3

carbofuran

4

Irmina Ćwieląg-Piaseckaa*, Maciej Witwickib*, Maria Jerzykiewiczb, Julia Jezierskab

5

a

6

and Life Sciences, Grunwaldzka 53 St., Wroclaw, Poland

7

b

8

TOC Art

Institute of Soil Science and Environmental Protection, Wroclaw University of Environmental

Faculty of Chemistry, Wroclaw University, 14 F. Joliot–Curie St., 50-383 Wroclaw, Poland

9

10

ABSTRACT

11

Radical oxidation of carbamate insecticides, namely carbaryl and carbofuran, was investigated

12

with spectroscopic (electron paramagnetic resonance [EPR] and UV-Vis) and theoretical (density

13

functional theory [DFT] and ab initio orbital-optimized spin-component scaled MP2

14

[OO-SCS-MP2]) methods. The two carbamates were subjected to reaction with •OH, persistent

15

DPPH• and galvinoxyl radical, as well as indigenous radicals of humic acids. The influence of

ACS Paragon Plus Environment

Environmental Science & Technology

16

fulvic acids on carbamate oxidation was also tested. The results obtained with EPR and UV-Vis

17

spectroscopy indicate that carbamates can undergo direct reactions with various radical species,

18

oxidizing themselves into radicals in the process. Hence, they are prone to participate in the

19

prolongation step of the radical chain reactions occurring in the soil environment. Theoretical

20

calculations revealed that from the thermodynamic point of view hydrogen atom transfer is the

21

preferred mechanism in the reactions of the two carbamates with the radicals. The activity of

22

carbofuran was determined experimentally (using pseudo-first-order kinetics) and theoretically

23

to be noticeably higher in comparison with carbaryl and comparable with gallic acid. The

24

findings of this study suggest that the radicals present in soil can play an important role in natural

25

remediation mechanisms of carbamates.

26

Keywords: scavenging kinetics, carbamates, galvinoxyl, DPPH, UV-Vis, EPR, DFT, OO-SCS-

27

MP2, ab initio

28

INTRODUCTION

29

Soil organic matter (SOM) plays a crucial role in immobilization of many xenobiotics in

30

soil.1,2 Depending on the soil physical and chemical properties and the chemical nature of organic

31

pollutants various binding mechanisms are responsible for this immobilization. The most

32

commonly reported mechanisms in the literature are ionic interactions, covalent and hydrogen

33

bonds, charge-transfer or electron donor-acceptor processes, Van der Waals forces, hydrophobic

34

bonding or partitioning between water and organic humic phases.3,4

35

Radicals, molecules containing a single, unpaired electron in an s- or p-type atomic orbital

36

or a molecular orbital constituting their linear combination, occur in soil in relatively high

37

concentrations as they can be generated in it biologically and abiologically.5–7 Moreover, various

ACS Paragon Plus Environment

Page 2 of 40

Page 3 of 40

Environmental Science & Technology

38

factors affect their generation or initiate formation of additional, environmentally persistent

39

radicals.8–16 Hence, much research has been conducted on the role of radicals in pesticide binding

40

mechanisms. For instance, formation of radicals was observed during reactions of humic

41

substances with various herbicides (e.g. s-triazines, substituted ureas), where charge transfer

42

complexes were formed.7 Conversely, chlorophenoxy herbicides quenched the radical content of

43

humic substances.3 Recently it was postulated that chemisorption and electron transfer from

44

organochlorine pesticide pentachlorophenol to transition metal ions and other electron sinks in

45

soil, lead to formation of new, environmentally persistent radicals.17 According to the literature a

46

number of substituted aromatic pesticides undergo oxidation reactions that involve the

47

generation of stable radicals in soil through removal of a proton or an electron from the

48

molecule.18 Nonetheless, the role of radical species in the binding mechanism of contaminants to

49

soil constituents is still under debate and further information about environmentally important

50

radical reactions involving various organic compounds can be found in the literature.3,7,19–21

51

Carbamate pesticides, represented here by carbaryl and carbofuran, belong to the group of

52

nonpolar, nonionic, toxic compounds widely applied as insecticides.22,23 In addition to being

53

agrochemical agents, carbamates are also used in the pharmaceutical and polymer industry.24–28

54

Hence, any information about their fate after they are released into the environment is of

55

significant importance. These N-substituted esters of carbamic acid are moderately persistent,

56

mobile and have strong affinity to soil organic matter.29 Their adsorption, mobility and

57

degradation in soil have been widely investigated.30–32 However, to the best of our knowledge,

58

there is still a limited amount of research on the possible oxidation of these two carbamates

59

involving various radical species and the role of SOM in this process. This gap in knowledge is

60

additionally exposed by the fact that the reactions of carbamates with •OH and •NO3 in the

ACS Paragon Plus Environment

Environmental Science & Technology

61

atmosphere have been investigated.33–35 Therefore, the main aim of this study was to investigate

62

the radical oxidation of carbamate pesticides, with carbaryl and carbofuran taken as a case study,

63

and their possible effect on soil organic matter. In order to achieve this goal, merged

64

spectroscopic (UV-Vis, electron paramagnetic resonance [EPR]) and theoretical (density

65

functional theory [DFT], ab initio) investigations were carried out.

66

MATERIALS AND METHODS

67

Chemicals and reagents. PBN (N-t-butyl-α-phenylnitrone) and DMPO (5,5-dimethyl-1-

68

pyrroline N-oxide) spin traps, GO• (galvinoxyl, 2,6-Di-tert-butyl-α-(3,5-di-tert-butyl-4-oxo-2,5-

69

cyclohexadien-1-ylidene)-p-tolyloxy), and DPPH• (2,2-diphenyl-1-picrylhydrazyl) radicals, and

70

investigated active substances of pesticides, that is carbaryl (1-naphthyl-N-methylcarbamate) and

71

carbofuran (2,3-dihydro-2,2-dimethyl-7-benzofuranol N-methylcarbamate), were purchased from

72

Sigma–Aldrich and Leonardite Humic Acid Standard from International Humic Substances

73

Society (IHSS). Ethanol, HCl, HF, NaOH, FeSO4 .7H2O and AgNO3 were obtained from POCH

74

(Avantor Performance Materials, Poland). All the reagents were of analytical grade.

75

Caution: carbaryl and carbofuran are highly toxic; proper care should be exercised while

76

handling these compounds.

77

Extraction, fractionation and purification of humic acids (HA) and fulvic acids (FA).

78

The source of HA and FA for this study was a topsoil horizon collected from the agricultural

79

Gleyic Phaeozems36 derived from silt loam located in the area of Domaniów, near Wroclaw,

80

Poland. Humic substances were extracted from the soil using the procedure recommended by the

81

International Humic Substances Society.37 The extracts were purified with a mixture of 0.1 M

82

HCl and 0.3 M HF in polypropylene tubes. They were left overnight and centrifuged. The

ACS Paragon Plus Environment

Page 4 of 40

Page 5 of 40

Environmental Science & Technology

83

precipitate was transferred to dialysis tubes (Spectra/Por 7 MWCO 10,000, Spectrum Europe

84

B.V., The Netherlands) and dialyzed with distilled water until a negative Cl− test with AgNO3.

85

Afterwards the humic acids were freeze dried. FA extracts were passed through an XAD-8

86

column followed by H+-saturated cation exchange resin. The eluate was freeze dried.

87

UV-Vis measurements. All the measurements were done under deaerated conditions at

88

298 K in ethanol solutions (2 mL in a 10 mm quartz cuvette) containing DPPH• (2.2 × 10-4 M) or

89

GO• (8.3 × 10-6 M) and a various amount of carbaryl (1.5 × 10−2 – 1.2 × 10−1 M) or carbofuran

90

(1.0 × 10−2 – 70 × 10−2 M). UV-Vis spectral changes associated with the undergoing reaction

91

were monitored at 516 and 429 nm for DPPH• (ε = 4.30 × 103 M–1 cm –1) and GO• (ε = 1.16 ×

92

105 M–1 cm –1), respectively, using a Varian Cary 50Conc UV-Visible Spectrophotometer. The

93

following equations of calibration curves were used to calculate DPPH• and GO• concentrations

94

in the reaction system:

95

A(516 nm) = 4.302 × 103 [DPPH•] + 8.14 × 40-3

96

A(429 nm) = 1.161 × 105 [GO•]

(R2 = 0.999)

(R2 = 0.996).

(Eq. 1) (Eq. 2)

97

In order to assess activity of the studied carbamates in radical reactions possibly occurring

98

in soil systems, two techniques were employed. In the first the scavenging capacity was judged

99

against DPPH• by measuring the percentage of remaining DPPH• after a fixed reaction time:

100

remaining DPPH• (%) = At× A0–1 × 100%,

101

where A0 and At correspond to the DPPH• absorbance at 516 nm in the absence and presence of

102

pesticide after a fixed time, respectively. In the second, the rate of DPPH• and GO• scavenging

103

reaction was determined using pseudo-first-order kinetics (the concentration of carbaryl and

104

carbofuran was maintained at more than 45-fold excess of the radical concentration; further

ACS Paragon Plus Environment

(Eq. 3)

Environmental Science & Technology

105

details are given as Supporting Information). These two techniques are frequently employed in

106

estimation of antiradical properties of antioxidants.38–40

107

EPR measurements. All the EPR measurements were conducted at a frequency close to 9.7

108

GHz (X-band) at room temperature using a Bruker ELEXYS E500 spectrometer equipped with a

109

NMR gaussmeter and frequency counter. Microwave power was set to 13 mW, magnetic field

110

modulation to 100 kHz and modulation amplitude to 1 G. Direct EPR measurements of freeze-

111

dried samples of HA and HA saturated with carbaryl and carbofuran were done in quartz EPR

112

tubes. Leonardite, a naturally occurring partly oxidized lignite with high humic acid content of

113

known radical concentration was used as a standard in quantitative EPR measurements. All the

114

experimental spectra were accumulated five times and simulated using the WinEPR SimFonia

115

program (1.25 version) developed by Bruker.

116

Radicals frequently have short half-lives and therefore they are often impossible to detect in

117

direct EPR measurements. Indirect EPR detection and identification of such radicals is possible

118

using spin trapping techniques. A spin trap reacts with the radical to form a spin adduct that is

119

more stable than the initial radical and thus can be observed in the EPR experiment. In this study

120

DMPO and PBN were employed as spin traps for the radicals formed from carbaryl and

121

carbofuran. These spin traps are commonly used in studies of radical reactions involving

122

superoxide and hydroxyl radicals.41,42 PBN is considered to react more rapidly with carbon-

123

centered radicals than with the oxygen-centered ones, opposite to DMPO.43,44

124

The compositions and abbreviations of all the EPR investigated samples are compiled in

125

Table 1. All these experiments were repeated with addition of 1 mg of iron(II) sulfate

126

heptahydrate to the investigated solutions. This should enhance the production of •OH, but was

127

found to have no effect on the result of the experiments. All the EPR spectra were collected

ACS Paragon Plus Environment

Page 6 of 40

Page 7 of 40

Environmental Science & Technology

128

twice, that is 5 minutes and 1 h after the initiation of the reaction. The samples investigated in

129

spin trapping studies were measured in glass capillaries inserted into quartz EPR tubes.

130

Table 1. Composition of all samples from EPR experiments.

Sample

0a 0b 0f 1 2 1a 1a* 1b 1b* 2a 2a* 2b 2b* 1f 2f 1af 1a’f 1bf 1b’f 2af 2a’f 2bf 2b’f

Oxidizing agent, H2O2 (volume, concentration) 0.05 mL, 3% 0.05 mL, 3% 0.05 mL, 3% 0.05 mL, 3% 0.22 mL, 3% 0.05 mL, 3% 0.05 mL, 30% 0.05 mL, 3% 0.05 mL, 30% 0.22 mL, 3% 0.22 mL, 30% 0.22 mL, 3% 0.22 mL, 30% 0.05 mL, 3% 0.22 mL, 3% 0.05 mL, 3% 0.05 mL, 3% 0.05 mL, 3% 0.05 mL, 3% 0.22 mL, 3% 0.22 mL, 3% 0.22 mL, 3% 0.22 mL, 3%

Pesticide

Spin trap

0.5 mL of carbaryl (25 mM) 0.5 mL of carbofuran (25 mM) 0.5 mL of carbaryl (25 mM) 0.5 mL of carbaryl (25 mM) 0.5 mL of carbofuran (25 mM) 0.5 mL of carbofuran (25 mM) 0.08 mL of carbaryl (25 mM) 0.08 mL of carbaryl (25 mM) 0.08 mL of carbofuran (25 mM) 0.08 mL of carbofuran (25 mM) 0.5 mL of carbaryl (25 mM) 0.5 mL of carbaryl (2.5 mM) 0.5 mL of carbofuran (25 mM) 0.5 mL of carbofuran (2.5 mM) 0.04 mL of carbaryl (25 mM) 0.04 mL of carbaryl (2.5 mM) 0.04 mL of carbofuran (25 mM) 0.04 mL of carbofuran (2.5mM)

8.5 mg of PBN 0.5 mL of 0.2 M DMPO 8.5 mg of PBN 8.5 mg of PBN 8.5 mg of PBN 8.5 mg of PBN 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO 8.5 mg of PBN 0.5 mL of 0.2 M DMPO 8.5 mg of PBN 8.5 mg of PBN 8.5 mg of PBN 8.5 mg of PBN 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO 0.5 mL of 0.2 M DMPO

Fulvic acid (2.5 mg⋅mL-1 in ethanol) 0.5mL 0.5 mL 0.08 mL 0.5 mL 0.5 mL 0.5 mL 0.5 mL 0.04 mL 0.04 mL 0.04 mL 0.04 mL

Ethanol

1.0 mL 0.08 mL 0.5 mL 0.5 mL 0.5 mL 0.5 mL 0.5 mL -

131

132

Sorption experiment. Sorption experiment was carried out using the simplified

133

equilibration method.45 Stock solutions of carbaryl (4.97 × 10-5 M) and carbofuran (4.52 × 10-5

134

M) in 1.0 ×10-2 M CaCl2 were prepared. 10 mL of the stock solutions were added to 50 mg of

135

HA giving 1s and 2s mixtures, respectively. As a control sample 50 mg of HA with 10 mL of 1.0

136

× 10-2 M CaCl2 solution was used (0s). All the mixtures were shaken for 24 h, centrifuged and

137

remained humic acids precipitates were decanted. Preliminary tests showed that a sorption

138

equilibrium was reached for the investigated carbamates within 24 hours. The residues of the

ACS Paragon Plus Environment

Environmental Science & Technology

139

humic acids were washed with distilled water until the negative reaction for the chloride ions

140

(with AgNO3) and freeze-dried. Such prepared samples were then analyzed for the EPR radical

141

spin concentration as described above.

142

Theoretical calculations. Recent years have witnessed increasing interest in the application

143

of modern electronic structure theory in solving environmental problems.46–53 In conjugation

144

with the experiments, theoretical calculations were conducted with the ORCA 3.0.3 suite of

145

programs.54 In all the calculations the def2-TZVP basis set developed by Ahlrichs and coworkers

146

was used55,56 and the COSMO formalism57,58 was employed to cover the solvent effects (water).

147

The calculations for open-shell species were conducted in the unrestricted protocol. The

148

geometry optimizations were carried out at the DFT level using the gradient-corrected BP86

149

functional, which provides accurate molecular structures.59–62 Each of the stationary points was

150

fully characterized as a true minimum through a vibrational analysis. On these molecular

151

structures single point calculations were carried out at the DFT level with the hybrid B3LYP,63–65

152

hybrid meta-GGA TPSSh,66 and double hybrid B2PLYP67 approximations as well as at the ab

153

initio level with the OO-SCS-MP2 method68. The bond dissociation enthalpies (BDE), ionization

154

potentials (IP) and Gibbs free energies (∆G) were calculated at 298 K with all these theoretical

155

methods but using the zero point, thermal (vibrational, rotational and translational) and entropy

156

corrections computed at the BP86/def2-TZVP theory level. BDE and IP were predicted for the

157

most stable conformers; the reported ∆G values refer to the most stable conformers of the most

158

stable isomers. A full summary of the calculations is given as Supporting Information (Table

159

S.1).

160

RESULTS AND DISCUSSION

161

Spectroscopic measurements. The radical chain mechanism involves three steps: initiation,

ACS Paragon Plus Environment

Page 8 of 40

Page 9 of 40

Environmental Science & Technology

162

propagation and termination.20 The initiation step can be induced by a number of processes, but

163

in a soil system the hydroxyl radical (•OH) can be expected to be the first radical of the chain, as

164

it is a powerful oxidant known to react with many organic compounds at nearly diffusion-limited

165

rates.69–73 In environment •OH can be generated photochemically6,74,75, in the Fenton reaction76–

166

78

167

independent pathway of •OH formation might be possible in soil.80

168

or by metal-loaded humic acids79. Recently, it was also demonstrated that a sunlight

EPR experiments. In order to judge to what extent carbamates (carb-H) can undergo the

169

reaction with •OH in the propagation phase:

170

carb-H + •OH → carb• + H2O,

171

and to identify the resulting radical product (carb•), EPR measurements were performed. If carb•

172

were detected, either in a direct measurement or as a spin adduct, it would obviously indicate that

173

radical derivatives of carbaryl and carbofuran can be formed and that they are therefore likely to

174

further participate in various radical reactions in soil systems.

(Eq. 4)

175

In the reaction systems H2O2 was always the initial source of •OH radicals, but regardless

176

of the lack or presence of Fe(II) no radicals (in 0a and 0b mixtures) and no spin adducts (in 1a,

177

1b, 2a, 2b mixtures) originating from the carbamates were observed. When DMPO was used (in

178

2a, 2b and reference sample 2) two radical species were detected (Figure S.1 in Supporting

179

Information): the well-known DMPO spin adduct of •OH (giso = 2.0055, aiso(14N) = aiso(1H) =

180

14.5 G) and a nitroxide radical that is the product of DMPO oxidation (giso = 2.0055, aiso(14N) =

181

14.2 G).81 The PBN adduct of •OH was not observed as it is known to decompose rapidly to

182

benzaldehyde and tert-butylhydroxylamine (t1/2 < 1 minute).42

183

It is well known that the spin trapping can be hindered by many factors and these include

184

the rate of production of primary (•OH) and secondary (carb•) radicals.82,83 Therefore, to enhance

ACS Paragon Plus Environment

Environmental Science & Technology

185

the formation rate of carb• the concentration of H2O2 was increased (1a*, 1b*, 2a* and 2b*

186

samples). As a result PBN adducts of radicals formed from carbamates were EPR detected (giso =

187

2.0056, aiso(14N) = 14.75 G, aiso(1H) = 2.85 G and giso = 2.0055, aiso(14N) = 15.0 G, aiso(1H) = 3.65

188

G for carbofuran and carbaryl, respectively), but only in case of this spin trap (1a* and 1b*). The

189

exemplary EPR spectrum is shown in Figure 1A which clearly confirms that the carb• radicals

190

are formed in the reaction with •OH.

191

More complex systems (1af, 1bf, 2af, 2bf) containing FA in addition to the carbamates,

192

spin traps and H2O2 were studied to estimate the effect of SOM on the formation of carbamate

193

derived radicals. But first direct EPR measurements were performed on ethanolic solution of FA

194

(0f), as it is well-known that anionic semiquinones (oxygen-centered radicals) derived from

195

fulvic acids (FA) can be observed in direct EPR experiments.84–87 However, this requires alkaline

196

conditions and the presence of oxidising agent (for example ambient oxygen). Since in our

197

experiments the reaction conditions were not alkaline, then the FA semiquinones were not

198

observed.

199

Then the reference EPR measurements were also performed on the samples containing FA

200

and spin traps but without carbamates (1f and 2f), revealing formation of PBN spin adduct with

201

radicals from FA (giso = 2.0055, aiso(14N) = 14.8 G and aiso(1H) = 3.2 G). The corresponding

202

DMPO adduct was not observed. The values of EPR parameters indicated that the carbon-

203

centered radicals were derived from fulvic acid.88,89 It is reasonable as the reactions with •OH are

204

known to be only slightly selective and PBN spin adducts with carbon-centered radicals in

205

similar phenolic systems were observed previously.89,90

ACS Paragon Plus Environment

Page 10 of 40

Page 11 of 40

Environmental Science & Technology

206

207

Figure 1. EPR spectra of: A) PBN spin adduct of radical generated from carbaryl (black line,

208

1a* sample) and simulation of its experimental spectrum (red, dashed line; giso = 2.0055, aiso(14N)

209

= 15.0 G, aiso(1H) = 3.65 G); B) PBN spin adduct of fulvic acid radical (blue line, 1f sample) and

210

PBN spin adduct of fulvic acid radical in the presence of carbaryl (1a’f sample; black line; see

211

text for details).

212

In contrast, the EPR measurements of the mixtures containing carbamates (1af, 1bf, 2af,

213

2bf) revealed that neither their radical spin adducts nor spin adducts with radicals from FA were

214

detected. In the next stage of the studies a tenfold lower concentration of carbaryl and carbofuran

215

was added to the reaction mixtures containing FA, H2O2 and spin traps (1a’f, 1b’f, 2a’f, 2b’f).

216

As a result the EPR spectra of the PBN spin adduct with FA radicals were recorded (1a’f and

217

1b’f) but, as is apparent from Figure 1B, their intensity was significantly lower in comparison to

218

the spectra recorded for the carbamate-free mixture (1f). This fact implies that the carbamates

219

can take part in the prolongation step of the radical chain process, probably by scavenging of

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 40

220

•OH according to the Eq. 4, what prevents further •OH reaction with FA and is accompanied by

221

formation of carb• as products. The formation of carb• is credible as the PBN-carb• adducts in

222

the 1a* and 1b* samples (see in the text above and Figure 1A) are clearly identified. However,

223

in the reaction mixtures containing FA the possibility for reaction of carbamates with radicals

224

originating from FA cannot be excluded:

225

FA• + carb-H → FA-H + carb•,

226

which in turn would prevent FA• from forming the spin adducts. And finally carbamates may

227

also react with the •PBN-FA spin adducts:

228 229 230

(Eq. 5)

carb-H + •PBN-FA → carb• + H-PBN-FA.

(Eq. 6)

In order for the process given by the Eq. 5 or 6 to occur the carbamates would have to be able to react with radicals that are weaker oxidants than •OH.

231

UV-Vis measurements. To determine the carbamates susceptibility to undergo oxidation by

232

less reactive radicals, leading to the formation of carb•, the easily detectable persistent radical

233

DPPH•, as universal standard in similar studies38–40, was used:

234

carb-H + DPPH• → carb• + H-DPPH.

(Eq. 7)

235

The change in concentration of DPPH• is equivalent to the concentration of formed carb•

236

radicals and can be accurately determined. A similar approach is frequently applied to ascertain

237

antiradical properties of antioxidants. The determination of antiradical properties is based on the

238

logical assumption that if a compound is able to react with DPPH•, it is certainly capable to react

239

with the much more aggressive •OH. Therefore, if the reaction between the carbamates and

240

DPPH• is observed, it will additionally corroborate the carbamates ability to react with •OH.

241

Moreover, there is an analogy between the methodology adopted here and the investigation of

242

atmospheric reactions of carbamates with •OH. In both cases the reactions were monitored

ACS Paragon Plus Environment

Page 13 of 40

Environmental Science & Technology

243

through the changes in concentration of one of the substrates, not by observing the radical

244

products.33,35

245

In order to confirm that carbaryl and carbofuran participate in the prolongation step three

246

mixtures were prepared and monitored over time. Two of them contained one of the carbamates

247

(30 mM) and DPPH• (0.22 mM) the third only DPPH• (0.22 mM). These results are shown in

248

Figure 2. The DPPH• absorption at 516 nm underwent a significant decrease in the case of

249

carbaryl and carbofuran confirming that its concentration in the reaction mixtures was lessened.

250

What stands out in Figure 2 is that DPPH• was reduced by carbofuran significantly faster than by

251

carbaryl.

252

For instance, after 30 seconds more than 30% of DPPH• was scavenged by carbofuran and

253

5% by carbaryl, whereas after 50 minutes about 64% was scavenged by carbofuran and 34% by

254

carbaryl. Interestingly, Figure 2 implies that the rate of carb• formation is higher for carbofuran.

255

To verify this the reaction of DPPH• with each carbamate was monitored using pseudo-first-

256

order kinetics. The calculated (as explained in Supporting Information) pseudo-first-order rate

257

constants, kobs, increased with the concentration of carbaryl and carbofuran, exhibiting first-order

258

dependence (Figure S.2). The second-order rate constants k2 for the reaction given in Eq. 7 were

259

determined from the slopes of the linear plots of kobs vs. concentration of the carbamates. The

260

value of k2 for carbofuran was three and a half times higher than for carbaryl, namely 0.060 M-1

261

s-1 (R2 = 0.988) for carbofuran and 0.017 M-1 s-1 (R2 = 0.980) for carbaryl. This fully explains the

262

faster decay of DPPH• shown for carbofuran in Figure 2. Moreover, the significant difference

263

between the values of k2 strongly suggests that the chemical structure of carbamates has a

264

decisive effect on the formation rate of carb•. This correlation is discussed below in the

265

theoretical section.

ACS Paragon Plus Environment

Environmental Science & Technology

266

267

Figure 2. Changes in DPPH• absorption in the 400 – 700 nm range due to its reaction with

268

carbaryl and carbofuran. The percentages of remaining DPPH• are given in brackets; they were

269

calculated according to Eq. 3.

270

Another important issue is the influence of SOM antioxidant activity on the observed radical

271

oxidation of carbamates. SOM is known to act as an antioxidant due to the large number of

272

phenolic moieties in its structure.91,92 Therefore, SOM can be expected to slow down or even

273

prevent the formation of radical derivatives of carbamates in the reactions with •OH or other

274

radicals. The fact that the second-order rate constants for scavenging of DPPH• by phenolic

ACS Paragon Plus Environment

Page 14 of 40

Page 15 of 40

Environmental Science & Technology

275

antioxidants are clearly higher than values determined here for carbamates seems to corroborate

276

such a hypothesis.39,40 On the other hand, the EPR spin-trapping experiments performed on

277

model system comprising FA, a constituent of SOM, and carbamates suggested that the carb•

278

radicals can be formed in the presence of FA. If SOM is able to prevent the carbamates from

279

participating in the prolongation step of radical oxidation in the soil environment, it must

280

scavenge the radicals, for instance •OH, which in reaction with carbamates can lead to carb• (Eq.

281

4). If SOM is to be favored in such a reaction:

282

SOM-H + •OH → SOM• + H2O,

(Eq. 8)

283

over the carbamates, the ability of the resulting phenoxy or semiquinone radicals (SOM•) to react

284

with carbaryl and carbofuran should be verified.

285

Given the complicated and unresolved structure of natural phenolic systems such as SOM or

286

tannins, model compounds of low molecular weight and with well-defined structure are

287

frequently used to mimic the properties of these natural systems.14–16,93–95

288

To establish qualitatively and quantitatively whether the reaction:

289

SOM• + carb–H → SOM-H + carb•

290

can occur in the soil environment, we decided to use a model oxygen-centered radical, i.e.

291

galvinoxyl (GO•). GO• was chosen because its molecular and electronic structure resemble the

292

radical species spread in SOM which (i) are characterized by higher spin delocalization onto

293

aromatic carbon atoms than their counterparts derived from simple compounds such as catechol

294

or 3,4-dihydroxybenzoic acid and (ii) have radical centers that are expected, especially in the

295

case of indigenous ones, to be isolated from external effects by the macromolecular matrix.7,96

296

Correspondingly, in GO• the spin density is efficiently delocalized and access to the oxygen

297

atoms is hindered by tert-butyl moieties (Figure S.3).

ACS Paragon Plus Environment

(Eq. 9)

Environmental Science & Technology

Page 16 of 40

298

The rate of the model reaction:

299

GO• + carb–H → GO-H + carb•,

300

was again investigated employing pseudo-first-order kinetics. The plots kobs vs. concentration of

301

carbaryl and carbofuran (Figure S.2) showed linear dependence with kobs undergoing

302

enhancement with the concentration of the carbamates. From the slopes the second-order rate

303

constants k2, were determined as 0.0079 M-1 s-1 (R2 = 0.993) and 0.021 M-1 s-1 (R2 = 0.986) for

304

carbaryl and carbofuran, respectively. As in the case of the reaction with DPPH•, the value of k2

305

is significantly higher for carbofuran. However, it should be noticed that the k2 constants for the

306

reactions of the two carbamates with GO• are noticeably lower in comparison with their

307

counterparts determined for the reaction with DPPH•.

(Eq. 10)

308

The environmental application of the findings presented above is of great interest. The

309

results strongly suggest that even if the antiradical action of SOM leads to the formation of

310

SOM•, this resulting radical can undergo a subsequent reaction with carbamates and therefore

311

trigger their participation in the prolongation step of radical processes taking place in the soil

312

system.

313

Sorption experiment. In order to confirm this conclusion an attempt to estimate the effect

314

which the carbamates can exert on the radicals naturally present in humic acids (indigenous

315

radicals) was made. EPR measurements of the spin (radical) concentration in pure humic acid

316

and the same humic acid treated with CaCl2, carbaryl and carbofuran were carried out after

317

freeze-drying the samples. Regardless of the chemical factor to which HA was exposed, the g

318

parameter remained unchanged and equal to 2.0032, which is a value typical for the indigenous

319

radicals.7,97 Addition of the insecticides solutions in CaCl2 to the pure HA resulted in a decrease

ACS Paragon Plus Environment

Page 17 of 40

Environmental Science & Technology

320

of spin concentration from 2.68 × 1017 spins per gram for untreated HA and 4.44 × 1017 spins per

321

gram for HA treated only with CaCl2 (in 0s) to 1.20 × 1017 (in 1s) and 1.55 × 1017 spins per gram

322

(in 2s) for HA with carbaryl and carbofuran, respectively. This experiment convincingly

323

demonstrated that the two tested carbamates are able to react even with the indigenous radicals.

324

The increase in the spin concentration for HA saturated only with CaCl2 follows the trend

325

reported previously for other diamagnetic metal ion, e. g. Zn(II) and Cd(II), interacting with

326

humic acids.14,98,99 The quenching of indigenous radicals by another group of pesticides, that is

327

water-dissolved chlorophenoxyalkanoic acids, was reported previously, but the quenching

328

mechanism was different in that case.3,100 The indigenous radicals were suggested not to react

329

directly with chlorophenoxyalkanoic acids but with radical intermediates of their degradation.

330

Theoretical calculations. In order to obtain a deeper insight into the reactions in which the

331

radical derivatives of carbamates can be formed, the thermodynamics of these processes was

332

evaluated using theoretical methods. Two different mechanisms were considered. One of them,

333

the one proposed above, assumes that a radical (R•) removes a hydrogen atom from the

334

carbamate (carb-H) that itself transforms into a radical (for this H-atom transfer the details are

335

given as Supporting Information):

336

R• + carb–H → R–H + carb•.

(Eq. 11)

337

The one-electron transfer was the second mechanism taken here under theoretical

338

investigation. In this mechanism the carbamate was assumed to contribute an electron to a

339

radical, becoming itself a radical cation in the process:

340

R• + carb–H → R– + carb–H•+.

341

The two mechanisms were previously considered in the context of antiradical activity of

(Eq. 12)

ACS Paragon Plus Environment

Environmental Science & Technology

342

Page 18 of 40

phenolic compounds.101,102

343

In the H-atom transfer the X–H bond dissociation enthalpy (BDE) should be considered a

344

crucial parameter, as the H atoms more weakly bonded would be more active in the reaction

345

given in Eq. 11. BDE was calculated at 298 K as the enthalpy difference for the reaction:

346

carb–H → carb• + H•.

(Eq. 13)

347

In order to determine the susceptibility of different hydrogen atoms to the abstraction, the

348

BDE values for the N–H and all C–H bonds were calculated for the two carbamates at DFT

349

(BP86, B3LYP, TPSSh, B2PLYP) and ab initio (OO-SCS-MP2) levels (Table S.2), which have

350

been reported to perform well for radical systems.62,103–107 For the sake of discussion clarity all

351

the C and N atoms forming bonds with hydrogen were labelled as shown in Figure 3. Before

352

performing a detailed analysis, it is reasonable to compare the performance of the computational

353

models. Table S.2 shows that all the employed methods gave a very consistent outcome, making

354

the final conclusions method-independent; in the further discussion the values from the B3LYP

355

calculations are used.

356

For the two carbamates the BDE values predicted for the hydrogen atoms bonded to

357

aromatic carbons are in range 112.1 – 114.3 kcal/mol and are significantly higher than those for

358

the Nb–H and other C–H bonds. For carbaryl the second largest BDE was found for the Nb–H

359

bond (105.5 kcal/mol) and the lowest for the Ca–H bond of the methyl group (93.1 kcal/mol),

360

showing that the latter hydrogen atoms should be the most active in the H-abstraction mechanism

361

in the case of carbaryl. These results for Ca–H and Nb–H are in line with those reported

362

previously for the N-methyl methylcarbamate by N. Borduas et al.34 (93.2 and 106.8 kcal/mol for

363

C–H and N–H, respectively) and for N-n-propyl methylcarbamate by R. J. Berry et al.108 (96.1

ACS Paragon Plus Environment

Page 19 of 40

364

Environmental Science & Technology

and 111.4 kcal/mol for C–H and N–H, respectively).

365

In the case of carbofuran the BDE values for the Ca–H and Nb–H bonds are 93.0 and 105.4

366

kcal/mol, respectively, so they very closely resemble their carbaryl counterpart and hence they

367

do not correlate with the higher rates of reactions between carbofuran and the radicals. It differs

368

in structure from carbaryl by the 2,2-dimethyl-3H-furanyl ring which is expected to bring about

369

dissimilar reactivity, which is well manifested in BDE. Although this value for the Cg–H bond is

370

101.6 kcal/mol, the homolytic breaking of the Cf–H bond requires only 84.2 kcal/mol. This is

371

significantly less than the energy necessary to enforce the analogical process for O–H in phenol

372

according to the theoretical predictions of Leopoldini et al.101 (97.1 kcal/mol). Moreover, this

373

value for carbofuran is only slightly higher than the BDE reported for gallic acid (82.3

374

kcal/mol),101 which is known to be a very strong antioxidant able to react with a vast array of

375

radicals. This surprisingly low BDE can be explained by the fact that the homolytic breaking of

376

Cf–H in carbofuran leads to the formation of an aromatic radical, which is stabilized by a

377

significant delocalization of spin density (Figure S.4). It should be noted that the hydrogen

378

abstraction from aromatic carbons results in formation of a σ-type aryl radical with strongly

379

localized spin density.109,110

380

ACS Paragon Plus Environment

Environmental Science & Technology

381

382

Figure 3. Labeling of atoms forming bonds with hydrogens in carbaryl (a – i) and carbofuran

383

molecule (a -g’) along with dissociation enthalpies calculated for these bonds at the B3LYP level

384

and at 298 K. Values for phenols are taken from ref.101.

385

In order to compare the capacity of carbaryl and carbofuran to donate an electron to a

386

radical (one-electron transfer), adiabatic ionization potentials (IP) were computed as the enthalpy

387

difference between the carbamate (carb–H) and its radical cation (carb–H•+), as it was done

388

previously for phenolic compounds.101 All the calculated values of IP are listed in Table S.3. It is

389

important to note that the ones obtained using OO-SCS-MP2 are significantly higher in

390

comparison with the results of DFT methods. Recently it has been shown that MP2 and its

391

variants suffer severely from a poor Hartree-Fock reference wave function when they are

392

employed to predict ionization energies unless an orbital optimization (OO) procedure is

393

included in the computation.111 In the latter case, even the accuracy of CCSD(T) can be

ACS Paragon Plus Environment

Page 20 of 40

Page 21 of 40

Environmental Science & Technology

394

surpassed. Taking into account that the DFT performance in the IP calculations has been

395

demonstrated as unsatisfactory,111 the values obtained with the OO-SCS-MP2 method ought to

396

be considered as the most accurate and they show that the IP of carbofuran is about 12 kcal/mol

397

lower than that of carbaryl (160.4 and 172.5 kcal/mol, respectively).

398

In order to characterize the processes that can trigger the radical oxidation of carbaryl or

399

carbofuran in greater depth, the Gibbs free energies (∆G) were calculated for the reactions given

400

in Eq. 11 (H-atom abstraction) and Eq. 12 (one-electron transfer) at 298 K. Three different

401

radical substrates were considered: DPPH•, GO• and •OH. The results are summarized in Table

402

S.4.

403

Regardless of the radical substrate in the one-electron transfer the ∆G values are positive,

404

but these results need to be interpreted with caution. As mentioned above DFT methods show

405

general large errors in predicted ionization energies,111 and the calculation of ∆G for the one-

406

electron transfer (Eq. 12) involves two of them, that is for R–/R• and carb-H/carb-H•+. As can be

407

seen from Table S.4. ∆G values obtained using all functionals are alarmingly lower in

408

comparison with OO-SCS-MP2 approach, and should be considered as errors in the DFT-based

409

ionization energy prediction. The OO-SCS-MP2 method shows that ∆G for the one-electron

410

transfer between the two carbamates and all the chosen radicals is high. To illustrate, ∆G is

411

139.0, 94.5 and 93.1 kcal/mol for such a reaction between carbofuran and •OH, DPPH• and GO•,

412

respectively. In the case of carbaryl these values become even higher. All in all, the OO-SCS-

413

MP2 calculations distinctively suggest that for the two studied carbamates the occurrence of one-

414

electron transfer in the soil environment is thermodynamically improbable.

415

In contrast to one-electron transfer, the ∆G values predicted for H-atom abstraction are

416

moderately positive for the model radicals DPPH• and GO• and negative for •OH. For example

ACS Paragon Plus Environment

Environmental Science & Technology

417

∆G predicted for carbofuran at the OO-SCS-MP2 level is -32.2, 14.9 and 16.9 kcal/mol for •OH,

418

DPPH• and GO•, respectively; and for carbaryl -22.2, 24.9 and 26.8 kcal/mol, respectively.

419

Although in the case of DPPH• and GO• the reaction is not spontaneous, it was proven by the

420

UV-Vis spectroscopy experiments that the input of energy at 298 K is sufficient to sustain the

421

generation of carb•. It is also notable that the predicted ∆G values for the H-atom abstraction

422

closely correlate with different activity of the two carbamates. In comparison with carbaryl, the

423

∆G values predicted for more reactive carbofuran are noticeably less positive in the case of

424

DPPH• and GO• and more negative in the case of •OH.

425

Implications. The more general significance of this study is that apart from previously

426

confirmed photodegradation,32,112–115 hydrolysis116–118 or biodegradation119–121 the radical

427

oxidation of carbamates should be taken into account when their fate in the soil environment is

428

considered. This process might be particularly important in the presence of environmentally

429

persistent radicals. First, they can induce the formation of •OH (and other reactive oxygen

430

species),122 which subsequently can react with carbamates. Second, it should be assumed that

431

carbamates are likely capable of reacting directly with various environmentally persistent

432

radicals as they were able to react with persistent DPPH• and GO•. And this fact may also open

433

the discussion on more practical application. In the last twenty years much research has been

434

undertaken on the development of soil remediation techniques based on advanced oxidation

435

processes (AOP), in which •OH initiates a series of oxidation reactions.123 Fenton processes are

436

the most commonly used AOP in the remediation of pesticide-contaminated soils, but their direct

437

application is very aggressive to soil and can be devastating to the native microbes.123 The AOP

438

procedures are more applicable in the case of persistent pesticides, while the carbamates are

439

considered to be moderately long-living in soil. However, due to their high toxicity an immediate

ACS Paragon Plus Environment

Page 22 of 40

Page 23 of 40

Environmental Science & Technology

440

action may be required if they are released to the environment in an uncontrolled way. Our

441

studies suggest that in such a situation there is a possibility that persistent radicals, that are less

442

harmful to the environment than AOP, can be considered as agents remediating carbamate-

443

contaminated soils.

444

ASSOCIATED CONTENT

445

Supporting information

446

The Supporting Information is available free of charge on the ACS Publications website. It

447

covers additional text concerning pseudo-first-order kinetics and the H-atom transfer mechanism

448

as well as four figures illustrating: the EPR spectrum of DMPO spin adducts, plots showing the

449

linear relation between kobs and [carb-H], molecular structure and singly occupied molecular

450

orbital (SOMO) of galvinoxyl radical (GO•) and the most stable isomer of the radical derived

451

from carbofuran via H-atom abstraction. Four tables contain a summary of theoretical

452

calculations, bond dissociation enthalpies, adiabatic ionization potentials, and the values of

453

Gibbs free energy for the H-atom and one-electron transfer.

454

AUTHOR INFORMATION

455

Corresponding Authors

456

*Author 1: Phone (0048) 713205635; [email protected]

457

Author 2: Phone (0048) 713757215; [email protected]

458

Notes

459

The authors declare no competing financial interest.

460

ACKNOWLEDGMENTS

461

This work was financially supported by the National Science Centre (NCN), Project No.

ACS Paragon Plus Environment

Environmental Science & Technology

462

2012/05/D/ST10/02223. All computations were performed using computers of the Wroclaw

463

Center for Networking and Supercomputing (Grant No. 47).

464

REFERENCES

465

(1)

pesticides with soil organic matter. Sci. Total Environ. 1992, 123–124, 205–217.

466

467

(2)

(3)

(4)

(5)

(6)

Ruggiero,P.; Dec, J.; Bollag, J. M. Soil as a Catalytic System. In Soil Biochemistry; Stotzky, G.; Bollag, J. M., Ed.; Marcel Dekker: New York, 1996; pp 79–122.

476

477

Steelink, C.; Tollin, G. Free radicals in soils. In Soil Biochemistry; McLaren, A.; Peterson, G. H., Ed.; Marcel Dekker: New York, 1967; pp 147–173.

474

475

Gevao, B.; Semple, K. T.; Jones, K. C. Bound pesticide residues in soils: a review. Environ. Pollut. 2000, 108 (1), 3–14.

472

473

Senesi, N. Binding mechanisms of pesticides to soil humic substances. Sci. Total Environ. 1992, 123–124 (Supplement C), 63–76.

470

471

Spark, K. M.; Swift, R. S. Effect of soil composition and dissolved organic matter on pesticide sorption. Sci. Total Environ. 2002, 298, 147–161.

468

469

Bollag, J. M.; Myers, C. J.; Minard, R. D. Biological and chemical interactions of

(7)

Senesi, N. Application of Electron Spin Resonance (ESR) Spectroscopy in Soil Chemistry

478

BT - Advances in Soil Science; Stewart, B. A., Ed.; Springer New York: New York, NY,

479

1990; pp 77–130.

480

(8)

Jia, H.; Nulaji, G.; Gao, H.; Wang, F.; Zhu, Y.; Wang, C. Formation and Stabilization of

481

Environmentally Persistent Free Radicals Induced by the Interaction of Anthracene with

482

Fe(III)-Modified Clays. Environ. Sci. Technol. 2016, 50 (12), 6310–6319.

ACS Paragon Plus Environment

Page 24 of 40

Page 25 of 40

483

Environmental Science & Technology

(9)

Nwosu, U. G.; Khachatryan, L.; Youm, S. G.; Roy, A.; dela Cruz, A. L. N.; Nesterov, E.

484

E.; Dellinger, B.; Cook, R. L. Model system study of environmentally persistent free

485

radicals formation in a semiconducting polymer modified copper clay system at ambient

486

temperature. RSC Adv. 2016, 6 (49), 43453–43462.

487

(10)

Nwosu, U. G.; Roy, A.; dela Cruz, A. L. N.; Dellinger, B.; Cook, R. Formation of

488

environmentally persistent free radical (EPFR) in iron(III) cation-exchanged smectite clay.

489

Environ. Sci. Process. Impacts 2016, 18 (1), 42–50.

490

(11)

dela Cruz, A. L.; Cook, R. L.; Lomnicki, S. M.; Dellinger, B. Effect of low temperature

491

thermal treatment on soils contaminated with pentachlorophenol and environmentally

492

persistent free radicals. Env. Sci Technol 2012, 46 (11), 5971–5978.

493

(12)

their model compounds. Chemosphere 2013, 92 (4), 445–450.

494

495

Jerzykiewicz, M. The effect of Hg(II) ions on the free radicals of humic substances and

(13)

Dela Cruz, A. L. N.; Gehling, W.; Lomnicki, S.; Cook, R.; Dellinger, B. Detection of

496

environmentally persistent free radicals at a superfund wood treating site. Environ. Sci.

497

Technol. 2011, 45 (15), 6356–6365.

498

(14)

Christoforidis, K. C.; Un, S.; Deligiannakis, Y. Effect of Metal Ions on the Indigenous

499

Radicals of Humic Acids: High Field Electron Paramagnetic Resonance Study. Environ.

500

Sci. Technol. 2010, 44 (18), 7011–7016.

501

(15)

Witwicki, M.; Jerzykiewicz, M.; Jaszewski, A. R.; Jezierska, J.; Ozarowski, A. Influence

502

of Pb(II) Ions on the EPR Properties of the Semiquinone Radicals of Humic Acids and

503

Model Compounds: High Field EPR and Relativistic DFT Studies. J. Phys. Chem. A 2009,

504

113 (51), 14115–14122.

ACS Paragon Plus Environment

Environmental Science & Technology

505

(16)

Giannakopoulos, E.; Christoforidis, K. C.; Tsipis, A.; Jerzykiewicz, M.; Deligiannakis, Y.

506

Influence of Pb(II) on the Radical Properties of Humic Substances and Model

507

Compounds. J. Phys. Chem. A 2005, 109 (10), 2223–2232.

508

(17)

dela Cruz, A. L. N.; Gehling, W.; Lomnicki, S.; Cook, R.; Dellinger, B. Detection of

509

Environmentally Persistent Free Radicals at a Superfund Wood Treating Site. Environ.

510

Sci. Technol. 2011, 45 (15), 6356–6365.

511

(18)

DRAGUN, J.; HELLING, C. S. PHYSICOCHEMICAL AND STRUCTURAL

512

RELATIONSHIPS OF ORGANIC CHEMICALS UNDERGOING SOIL- AND CLAY-

513

CATALYZED FREE-RADICAL OXIDATION1. Soil Sci. 1985, 139 (2).

514

(19)

interacting with chemical organic contaminants. J. Geochemical Explor. 2013, 129, 6–13.

515

516

(20)

(21)

(22)

Rajagopal, B. S.; Sethunathan, N. Influence of nitrogen fertilisers on the persistence of carbaryl and carbofuran in flooded soils. Pestic. Sci. 1984, 15 (6), 591–599.

521

522

Free Radicals in Biology and Environment; Minisci, F., Ed.; NATO ASI Series, Springer Science and Business Media: Dordrecht, 1997.

519

520

Fossey, J.; Lefort, D.; Sorba, J. Free Radicals in Organic Chemistry; Wiley:New York, 1995.

517

518

Aleksandrova, O. N. Spin labeling ESR investigation of the molecular environment of soil

(23)

Fang, F.; Kanan, S.; Patterson, H. H.; Cronan, C. S. A spectrofluorimetric study of the

523

binding of carbofuran, carbaryl, and aldicarb with dissolved organic matter. Anal. Chim.

524

Acta 1998, 373 (2–3), 139–151.

525

(24)

Finkin-Groner, E.; Moradov, D.; Shifrin, H.; Bejar, C.; Nudelman, A.; Weinstock, M.

ACS Paragon Plus Environment

Page 26 of 40

Page 27 of 40

Environmental Science & Technology

526

Indoline-3-propionate and 3-aminopropyl carbamates reduce lung injury and pro-

527

inflammatory cytokines induced in mice by LPS. Br. J. Pharmacol. 2015, 172 (4), 1101–

528

1113.

529

(25)

Niphakis, M. J.; Cognetta, A. B.; Chang, J. W.; Buczynski, M. W.; Parsons, L. H.; Byrne,

530

F.; Burston, J. J.; Chapman, V.; Cravatt, B. F. Evaluation of NHS Carbamates as a Potent

531

and Selective Class of Endocannabinoid Hydrolase Inhibitors. ACS Chem. Neurosci. 2013,

532

4 (9), 1322–1332.

533

(26)

Chaturvedi, D. Role of Organic Carbamates in Anticancer Drug Design. In Chemistry and

534

Pharmacology of Naturally Occurring Bioactive Compounds; CRC Press, 2013; pp 117–

535

140.

536

(27)

Implications for the Polyurethane Industry, 1st ed.; Elsevier Inc., 2014; Vol. 66.

537

538

Heyn, R. H.; Jacobs, I.; Carr, R. H. Synthesis of Aromatic Carbamates from CO2.

(28)

Winkler, M.; Meier, M. A. R. Olefin cross-metathesis as a valuable tool for the

539

preparation of renewable polyesters and polyamides from unsaturated fatty acid esters and

540

carbamates. Green Chem. 2014, 16 (6), 3335–3340.

541

(29)

Ahmad, R.; Kookana, R. S.; Alston, A. M.; Skjemstad, J. O. The Nature of Soil Organic

542

Matter Affects Sorption of Pesticides. 1. Relationships with Carbon Chemistry as

543

Determined by 13C CPMAS NMR Spectroscopy. Environ. Sci. Technol. 2001, 35 (5),

544

878–884.

545 546

(30)

Thapar, S.; Bhushan, R.; Mathur, R. P. Degradation of organophosphorus and carbamate pesticides in soils—HPLC determination. Biomed. Chromatogr. 1995, 9 (1), 18–22.

ACS Paragon Plus Environment

Environmental Science & Technology

547

(31)

Nkedi-Kizza, P.; Brown, K. D. Sorption, Degradation, and Mineralization of Carbaryl in

548

Soils, for Single-Pesticide and Multiple-Pesticide Systems. J. Environ. Qual. 1998, 27,

549

1318–1324.

550

(32)

Bachman, J.; Patterson, H. H. Photodecomposition of the carbamate pesticide carbofuran:

551

Kinetics and the influence of dissolved organic matter. Environ. Sci. Technol. 1999, 33

552

(6), 874–881.

553

(33)

Kwok, E. S. C.; Aschmann, S. M.; Atkinson, R. Rate Constants for the Gas-Phase

554

Reactions of the OH Radical with Selected Carbamates and Lactates. Environ. Sci.

555

Technol. 1996, 30 (1), 329–334.

556

(34)

Borduas, N.; Abbatt, J. P. D.; Murphy, J. G.; So, S.; da Silva, G. Gas-Phase Mechanisms

557

of the Reactions of Reduced Organic Nitrogen Compounds with OH Radicals. Environ.

558

Sci. Technol. 2016, 50 (21), 11723–11734.

559

(35)

thiocarbamates. Environ. Sci. Technol. 1992, 26 (9), 1798–1807.

560

561

(36)

IUSS Working Group WRB. World reference base for soil resources 2014. International soil classification system for naming soils and creating legends for soil maps; 2014.

562

563

Kwok, E. S. C.; Atkinson, R.; Arey, J. Gas-phase atmospheric chemistry of selected

(37)

Swift, R. S. Organic Matter Characterization. In Methods of Soil Analysis Part 3—

564

Chemical Methods; SSSA Book Series SV - 5.3; Soil Science Society of America,

565

American Society of Agronomy: Madison, WI, 1996; pp 1011–1069.

566 567

(38)

Brand-Williams, W.; Cuvelier, M. E.; Berset, C. Use of a free radical method to evaluate antioxidant activity. LWT - Food Sci. Technol. 1995, 28 (1), 25–30.

ACS Paragon Plus Environment

Page 28 of 40

Page 29 of 40

568

Environmental Science & Technology

(39)

Mishra, K.; Ojha, H.; Chaudhury, N. K. Estimation of antiradical properties of

569

antioxidants using DPPH

570

2012, 130 (4), 1036–1043.

571

(40)

- assay: A critical review and results. Food Chem.

Inami, K.; Iizuka, Y.; Furukawa, M.; Nakanishi, I.; Ohkubo, K.; Fukuhara, K.; Fukuzumi,

572

S.; Mochizuki, M. Chlorine atom substitution influences radical scavenging activity of 6-

573

chromanol. Bioorganic Med. Chem. 2012, 20 (13), 4049–4055.

574

(41)

Jerzykiewicz, M.; Ćwieląg-Piasecka, I.; Witwicki, M.; Jezierski, A. EPR spin trapping and

575

DFT studies on structure of active antioxidants in biogycerol. Chem. Phys. Lett. 2010, 497

576

(1–3), 135–141.

577

(42)

Reinke, L. A.; Moore, D. R.; Sang, H.; Janzen, E. G.; Kotake, Y. Aromatic hydroxylation

578

in PBN spin trapping by hydroxyl radicals and cytochrome P-450. Free Radic. Biol. Med.

579

2000, 28 (3), 345–350.

580

(43)

Thomas, C. E.; Ohlweiler, D. F.; Carr, A. A.; Nieduzak, T. R.; Hay, D. A.; Adams, G.;

581

Vaz, R.; Bernotas, R. C. Characterization of the radical trapping activity of a novel series

582

of cyclic nitrone spin traps. J. Biol. Chem. 1996, 271 (6), 3097–3104.

583

(44)

Andersen, M. L.; Skibsted, L. H. Electron Spin Resonance Spin Trapping Identification of

584

Radicals Formed during Aerobic Forced Aging of Beer. J. Agric. Food Chem. 1998, 46

585

(4), 1272–1275.

586

(45)

Guidel. Test. Chem. 2000, No. January, 1–44.

587

588

OECD. OECD 106 Adsorption - Desorption Using a Batch Equilibrium Method. OECD

(46)

Kersten, M.; Tunega, D.; Georgieva, I.; Vlasova, N.; Branscheid, R. Adsorption of the

ACS Paragon Plus Environment

Environmental Science & Technology

589

Herbicide 4-Chloro-2-methylphenoxyacetic Acid (MCPA) by Goethite. Environ. Sci.

590

Technol. 2014, 48 (20), 11803–11810.

591

(47)

Belzunces, B.; Hoyau, S.; Benoit, M.; Tarrat, N.; Bessac, F. Theoretical study of the

592

atrazine pesticide interaction with pyrophyllite and Ca2+-montmorillonite clay surfaces. J.

593

Comput. Chem. 2017, 38 (3), 133–143.

594

(48)

Schampera, B.; Šolc, R.; Tunega, D.; Dultz, S. Experimental and molecular dynamics

595

study on anion diffusion in organically modified bentonite. Appl. Clay Sci. 2016, 120, 91–

596

100.

597

(49)

Aquino, A. J. A.; Tunega, D.; Schaumann, G. E.; Haberhauer, G.; Gerzabek, M. H.;

598

Lischka, H. Proton transfer processes in polar regions of humic substances initiated by

599

aqueous aluminum cation bridges: A computational study. Geoderma 2014, 213, 115–123.

600

(50)

Biphenyls. Environ. Sci. Technol. 2016, 50 (12), 6293–6298.

601

602

Krzemińska, A.; Paneth, P. DFT Studies of SN2 Dechlorination of Polychlorinated

(51)

Le Pape, P.; Blanchard, M.; Brest, J.; Boulliard, J.-C.; Ikogou, M.; Stetten, L.; Wang, S.;

603

Landrot, G.; Morin, G. Arsenic Incorporation in Pyrite at Ambient Temperature at Both

604

Tetrahedral S–I and Octahedral FeII Sites: Evidence from EXAFS–DFT Analysis.

605

Environ. Sci. Technol. 2017, 51 (1), 150–158.

606

(52)

Jing, Y.; Chaplin, B. P. Mechanistic Study of the Validity of Using Hydroxyl Radical

607

Probes To Characterize Electrochemical Advanced Oxidation Processes. Environ. Sci.

608

Technol. 2017, 51 (4), 2355–2365.

609

(53)

Assaf, N. W.; Altarawneh, M.; Oluwoye, I.; Radny, M.; Lomnicki, S. M.; Dlugogorski, B.

ACS Paragon Plus Environment

Page 30 of 40

Page 31 of 40

Environmental Science & Technology

610

Z. Formation of Environmentally Persistent Free Radicals on α-Al2O3. Environ. Sci.

611

Technol. 2016, 50 (20), 11094–11102.

612

(54)

(1), 73–78.

613

614

(55)

Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97 (4), 2571–2577.

615

616

Neese, F. The ORCA program system. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2

(56)

Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and

617

quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys.

618

Chem. Chem. Phys. 2005, 7 (18), 3297.

619

(57)

Klamt, A.; Schuurmann, G. COSMO: a new approach to dielectric screening in solvents

620

with explicit expressions for the screening energy and its gradient. J. Chem. Soc. Perkin

621

Trans. 2 1993, No. 5, 799–805.

622

(58)

Sinnecker, S.; Rajendran, A.; Klamt, A.; Diedenhofen, M.; Neese, F. Calculation of

623

Solvent Shifts on Electronic g-Tensors with the Conductor-Like Screening Model

624

(COSMO) and Its Self-Consistent Generalization to Real Solvents (Direct COSMO-RS).

625

J. Phys. Chem. A 2006, 110 (6), 2235–2245.

626

(59)

behavior. Phys. Rev. A 1988, 38 (6), 3098–3100.

627

628

(60)

Perdew, J. P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33 (12), 8822–8824.

629

630

Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic

(61)

Neese, F.; Ames, W.; Christian, G.; Kampa, M.; Liakos, D. G.; Pantazis, D. A.; Roemelt,

ACS Paragon Plus Environment

Environmental Science & Technology

631

M.; Surawatanawong, P.; Shengfa, Y. E. Dealing with Complexity in Open-Shell

632

Transition Metal Chemistry from a Theoretical Perspective: Reaction Pathways, Bonding,

633

Spectroscopy, And Magnetic Properties. In Theoretical and Computational Inorganic

634

Chemistry; van Eldik, R., Harvey, J. B. T.-A. in I. C., Eds.; Academic Press, 2010; Vol.

635

62, pp 301–349.

636

(62)

Properties: g Matrix. ChemPhysChem 2015, 16 (9), 1912–1925.

637

638

(63)

(64)

Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37 (2), 785–789.

641

642

Becke, A. D. Density‐functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98 (7), 5648–5652.

639

640

Witwicki, M. Theoretical Characterisation of Phosphinyl Radicals and Their Magnetic

(65)

Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of

643

Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force

644

Fields. J. Phys. Chem. 1994, 98 (45), 11623–11627.

645

(66)

Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional

646

Ladder: Nonempirical Meta\char21{}Generalized Gradient Approximation Designed for

647

Molecules and Solids. Phys. Rev. Lett. 2003, 91 (14), 146401.

648

(67)

correlation. J. Chem. Phys. 2006, 124 (3), 34108.

649

650 651

Grimme, S. Semiempirical hybrid density functional with perturbative second-order

(68)

Neese, F.; Schwabe, T.; Kossmann, S.; Schirmer, B.; Grimme, S. Assessment of OrbitalOptimized, Spin-Component Scaled Second-Order Many-Body Perturbation Theory for

ACS Paragon Plus Environment

Page 32 of 40

Page 33 of 40

Environmental Science & Technology

Thermochemistry and Kinetics. J. Chem. Theory Comput. 2009, 5 (11), 3060–3073.

652

653

(69)

Goldstone, J. V; Pullin, M. J.; Bertilsson, S.; Voelker, B. M. Reactions of Hydroxyl

654

Radical with Humic Substances:  Bleaching, Mineralization, and Production of

655

Bioavailable Carbon Substrates. Environ. Sci. Technol. 2002, 36 (3), 364–372.

656

(70)

Brezonik, P. L.; Fulkerson-Brekken, J. Nitrate-Induced Photolysis in Natural Waters: 

657

Controls on Concentrations of Hydroxyl Radical Photo-Intermediates by Natural

658

Scavenging Agents. Environ. Sci. Technol. 1998, 32 (19), 3004–3010.

659

(71)

Westerhoff, P.; Aiken, G.; Amy, G.; Debroux, J. Relationships between the structure of

660

natural organic matter and its reactivity towards molecular ozone and hydroxyl radicals.

661

Water Res 1999, 33 (10), 2265–2276.

662

(72)

chemicals in water. Environ. Sci. Technol. 1987, 21 (5), 443–450.

663

664

(73)

Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (•OH). Chem. Rev. 2015, 115 (24), 13051–13092.

665

666

Zepp, R. G.; Hoigne, J.; Bader, H. Nitrate-induced photooxidation of trace organic

(74)

Lee, E.; Glover, C. M.; Rosario-Ortiz, F. L. Photochemical Formation of Hydroxyl

667

Radical from Effluent Organic Matter: Role of Composition. Environ. Sci. Technol. 2013,

668

47 (21), 12073–12080.

669

(75)

Effluent Organic Matter. Environ. Sci. Technol. 2012, 46 (7), 3788–3794.

670

671 672

Dong, M. M.; Rosario-Ortiz, F. L. Photochemical Formation of Hydroxyl Radical from

(76)

Vermilyea, A. W.; Voelker, B. M. Photo-Fenton Reaction at Near Neutral pH. Environ. Sci. Technol. 2009, 43 (18), 6927–6933.

ACS Paragon Plus Environment

Environmental Science & Technology

673

(77)

White, E. M.; Vaughan, P. P.; Zepp, R. G. Role of the photo-Fenton reaction in the

674

production of hydroxyl radicals and photobleaching of colored dissolved organic matter in

675

a coastal river of the southeastern United States. Aquat. Sci. 2003, 65 (4), 402–414.

676

(78)

Zepp, R. G.; Faust, B. C.; Hoigne, J. Hydroxyl radical formation in aqueous reactions (pH

677

3-8) of iron(II) with hydrogen peroxide: the photo-Fenton reaction. Environ. Sci. Technol.

678

1992, 26 (2), 313–319.

679

(79)

Loaded Humic Acids. Environ. Sci. Technol. 1999, 33 (11), 1814–1818.

680

681

Paciolla, M. D.; Davies, G.; Jansen, S. A. Generation of Hydroxyl Radicals from Metal-

(80)

Page, S. E.; Kling, G. W.; Sander, M.; Harrold, K. H.; Logan, J. R.; McNeill, K.; Cory, R.

682

M. Dark Formation of Hydroxyl Radical in Arctic Soil and Surface Waters. Environ. Sci.

683

Technol. 2013, 47 (22), 12860–12867.

684

(81)

Hideg, É.; Spetea, C.; Vass, I. Singlet oxygen and free radical production during acceptor-

685

and donor-side-induced photoinhibition: Studies with spin trapping EPR spectroscopy.

686

Biochim. Biophys. Acta - Bioenerg. 1994, 1186 (3), 143–152.

687

(82)

Swartz, H. M.; Khan, N.; Khramtsov, V. V. Use of electron paramagnetic resonance

688

spectroscopy to evaluate the redox state in vivo. Antioxidants Redox Signal. 2009, 9 (10),

689

1757–1771.

690

(83)

Radicals in Biology; Pryor, W., Ed.; Academic Press: New York, 1980; pp 115–154.

691

692 693

(77) Janzen, E. G. A Critical Review of Spin Trapping in Biological Systems. In Free

(84)

SENESI, N.; SCHNITZER, M. EFFECTS OF pH, REACTION TIME, CHEMICAL REDUCTION AND IRRADIATION ON ESR SPECTRA OF FULVIC ACID. Soil Sci.

ACS Paragon Plus Environment

Page 34 of 40

Page 35 of 40

Environmental Science & Technology

1977, 123 (4).

694

695

(85)

Senesi, N. Molecular and quantitative aspects of the chemistry of fulvic acid and its

696

interactions with metal ions and organic chemicals: Part I. The electron spin resonance

697

approach. Anal. Chim. Acta 1990, 232 (Supplement C), 51–75.

698

(86)

Paul, A.; Stösser, R.; Zehl, A.; Zwirnmann, E.; Vogt, R. D.; Steinberg, C. E. W. Nature

699

and Abundance of Organic Radicals in Natural Organic Matter:  Effect of pH and

700

Irradiation. Environ. Sci. Technol. 2006, 40 (19), 5897–5903.

701

(87)

soil, HA, FA and humin substances. J. Braz. Chem. Soc. 2004, 15 (1), 34–37.

702

703

(88)

Buettner, G. R. Spin Trapping - Electron-Spin-Resonance Parameters of Spin Adducts. Free Radic. Bio. Med. 1987, 3 (4), 259–303.

704

705

Saab, S. C.; Martin-Neto, L. Studies of semiquinone free radicals by ESR in the whole

(89)

Jerzykiewicz, M.; Ćwieląg-Piasecka, I.; Witwicki, M.; Jezierski, A. α-Tocopherol impact

706

on oxy-radical induced free radical decomposition of DMSO: Spin trapping EPR and

707

theoretical studies. Chem. Phys. 2011, 383 (1–3), 27–34.

708

(90)

Navarrete, M.; Rangel, C.; Corchado, J. C.; Espinosa-García, J. Trapping of the OH

709

Radical by α-Tocopherol:  A Theoretical Study. J. Phys. Chem. A 2005, 109 (21), 4777–

710

4784.

711

(91)

Humic Substances. Environ. Sci. Technol. 2012, 46 (9), 4916–4925.

712

713 714

Aeschbacher, M.; Graf, C.; Schwarzenbach, R. P.; Sander, M. Antioxidant Properties of

(92)

Lindsey, M. E.; Tarr, M. A. Inhibition of Hydroxyl Radical Reaction with Aromatics by Dissolved Natural Organic Matter. Environ. Sci. Technol. 2000, 34 (3), 444–449.

ACS Paragon Plus Environment

Environmental Science & Technology

715

(93)

Christoforidis, K. C.; Un, S.; Deligiannakis, Y. High-Field 285 GHz Electron

716

Paramagnetic Resonance Study of Indigenous Radicals of Humic Acids. J. Phys. Chem. A

717

2007, 111 (46), 11860–11866.

718

(94)

Witwicki, M.; Jerzykiewicz, M.; Ozarowski, A. Understanding natural semiquinone

719

radicals - Multifrequency EPR and relativistic DFT studies of the structure of Hg(II)

720

complexes. Chemosphere 2015, 119, 479–484.

721

(95)

semiquinone complexes from natural phenols. Chemosphere 2015, 138, 233–238.

722

723

Jerzykiewicz, M.; Witwicki, M.; Jezierska, J. PH-dependent formation of Hg(II)-

(96)

Witwicki, M.; Jezierska, J. Protonated o-semiquinone radical as a mimetic of the humic

724

acids native radicals: A DFT approach to the molecular structure and EPR properties.

725

Geochim. Cosmochim. Acta 2012, 86, 384–391.

726

(97)

Watanabe, A.; McPhail, D. B.; Maie, N.; Kawasaki, S.; Anderson, H. A.; Cheshire, M. V.

727

Electron spin resonance characteristics of humic acids from a wide range of soil types.

728

Org. Geochem. 2005, 36 (7), 981–990.

729

(98)

Jezierski, A.; Czechowski, F.; Jerzykiewicz, M.; Chen, Y.; Drozd, J. Electron

730

paramagnetic resonance (EPR) studies on stable and transient radicals in humic acids from

731

compost, soil, peat and brown coal. Spectrochim. Acta - Part A Mol. Biomol. Spectrosc.

732

2000, 56 (2), 379–385.

733

(99)

Jezierski, A.; Czechowski, F.; Jerzykiewicz, M.; Drozd, J. EPR investigations of structure

734

of humic acids from compost, soil, peat and soft brown coal upon oxidation and metal

735

uptake. Appl. Magn. Reson. 2000, 18 (1), 127–136.

ACS Paragon Plus Environment

Page 36 of 40

Page 37 of 40

736

Environmental Science & Technology

(100) Senesi, N. Organic Pollutant Migration in Soils as Affected by Soil Organic Matter.

737

Molecular and Mechanistic Aspects BT - Migration and Fate of Pollutants in Soils and

738

Subsoils; Petruzzelli, D., Helfferich, F. G., Eds.; Springer Berlin Heidelberg: Berlin,

739

Heidelberg, 1993; pp 47–74.

740

(101) Leopoldini, M.; Marino, T.; Russo, N.; Toscano, M. Antioxidant Properties of Phenolic

741

Compounds:  H-Atom versus Electron Transfer Mechanism. J. Phys. Chem. A 2004, 108

742

(22), 4916–4922.

743

(102) Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.-L.; Olivier, Y.; Trouillas, P.

744

Free Radical Scavenging by Natural Polyphenols: Atom versus Electron Transfer. J. Phys.

745

Chem. A 2013, 117 (10), 2082–2092.

746

(103) Severino, J. F.; Goodman, B. A.; Kay, C. W. M.; Stolze, K.; Tunega, D.; Reichenauer, T.

747

G.; Pirker, K. F. Free radicals generated during oxidation of green tea polyphenols:

748

Electron paramagnetic resonance spectroscopy combined with density functional theory

749

calculations. Free Radic. Biol. Med. 2009, 46 (8), 1076–1088.

750

(104) Šolc, R.; Gerzabek, M. H.; Lischka, H.; Tunega, D. Radical sites in humic acids: A

751

theoretical study on protocatechuic and gallic acids. Comput. Theor. Chem. 2014, 1032,

752

42–49.

753

(105) Witwicki, M.; Jezierska, J. Toward an Understanding of the Ambiguous Electron

754

Paramagnetic Resonance Spectra of the Iminoxy Radical from o-Fluorobenzaldehyde

755

Oxime: Density Functional Theory and ab Initio Studies. J. Phys. Chem. A 2015, 119 (34),

756

9109–9120.

757

(106) Błoch-Mechkour, A.; Bally, T. Electromers of the benzene dimer radical cation. Phys.

ACS Paragon Plus Environment

Environmental Science & Technology

758

759

Chem. Chem. Phys. 2015, 17 (16), 10624–10629. (107) Kossmann, S.; Neese, F. Correlated ab Initio Spin Densities for Larger Molecules:

760

Orbital-Optimized Spin-Component-Scaled MP2 Method. J. Phys. Chem. A 2010, 114

761

(43), 11768–11781.

762

(108) Berrya, R. J.; A.L. Wilsona, M. S. A Computational study of bond dissociation enthaphies

763

and hydrogen abstarction energy barriers in model urethanes. J. Mol. Struct. 2000, 496,

764

121–129.

765

(109) Mattar, S. M. Accurate Calculation of the Phenyl Radical’s Magnetic Inequivalency,

766

Relative Orientations of Its Spin Hamiltonian Tensors, and Its Electronic Spectrum. J.

767

Phys. Chem. A 2007, 111 (2), 251–260.

768

(110) Martinez, O.; Crabtree, K. N.; Gottlieb, C. A.; Stanton, J. F.; McCarthy, M. C. An

769

Accurate Molecular Structure of Phenyl, the Simplest Aryl Radical. Angew. Chemie Int.

770

Ed. 2015, 54 (6), 1808–1811.

771

(111) Isegawa, M.; Neese, F.; Pantazis, D. A. Ionization Energies and Aqueous Redox

772

Potentials of Organic Molecules: Comparison of DFT, Correlated ab Initio Theory and

773

Pair Natural Orbital Approaches. J. Chem. Theory Comput. 2016, 12 (5), 2272–2284.

774

(112) Tanaka, K.; Robledo, S. M.; Hisanaga, T.; Ali, R.; Ramli, Z.; Bakar, W. A. Photocatalytic

775

degradation of 3, 4-xylyl N-methylcarbamate (MPMC) and other carbamate pesticides in

776

aqueous TiO 2 suspensions. J. Mol. Catal. A Chem. 1999, 144 (3), 425–430.

777 778

(113) Siampiringue, M.; Wong-Wah-Chung, P.; Sarakha, M. Impact of the soil structure and organic matter contents on the photodegradation of the insecticide carbaryl. J. Soils

ACS Paragon Plus Environment

Page 38 of 40

Page 39 of 40

779

780 781

782 783

784

Environmental Science & Technology

Sediments 2015, 15 (2), 401–409. (114) Climent, M. J.; Miranda, M. a. Gas chromatographic-mass spectrometric study of photodegradation of carbamate pesticides. J. Chromatogr. A 1996, 738 (2), 225–231. (115) Tanaka, K.; Reddy, K. S. N. Photodegradation of phenoxyacetic acid and carbamate pesticides on TiO2. Appl. Catal. B Environ. 2002, 39 (4), 305–310. (116) Sogorb, M. A.; Vilanova, E. Enzymes involved in the detoxification of organophosphorus,

785

carbamate and pyrethroid insecticides through hydrolysis. Toxicol. Lett. 2002, 128 (1–3),

786

215–228.

787

(117) Zamy, C.; Mazellier, P.; Legube, B. Analytical and kinetic study of the aqueous

788

hydrolysis of four organophosphorus and two carbamate pesticides. Int. J. Environ. Anal.

789

Chem. 2004, 84 (14–15), 1059–1068.

790

(118) Falah, I. I.; Hammers, W. E. The kinetics of the hydrolysis of organophosphorus and

791

carbamate pesticides in water and aqueous base buffers. Toxicol. Environ. Chem. 1994, 42

792

(1–2), 9–18.

793

(119) Pimmata, P.; Reungsang, A.; Plangklang, P. Comparative bioremediation of carbofuran

794

contaminated soil bynatural attenuation, bioaugmentation and biostimulation. Int.

795

Biodeterior. Biodegrad. 2013, 85, 196–204.

796

(120) Tien, C. J.; Lin, M. C.; Chiu, W. H.; Chen, C. S. Biodegradation of carbamate pesticides

797

by natural river biofilms in different seasons and their effects on biofilm community

798

structure. Environ. Pollut. 2013, 179, 95–104.

799

(121) Chanika, E.; Georgiadou, D.; Soueref, E.; Karas, P.; Karanasios, E.; Tsiropoulos, N. G.;

ACS Paragon Plus Environment

Environmental Science & Technology

800

Tzortzakakis, E. A.; Karpouzas, D. G. Isolation of soil bacteria able to hydrolyze both

801

organophosphate and carbamate pesticides. Bioresour. Technol. 2011, 102 (3), 3184–

802

3192.

803

(122) Khachatryan, L.; Vejerano, E.; Lomnicki, S.; Dellinger, B. Environmentally persistent free

804

radicals (EPFRs). 1. Generation of reactive oxygen species in aqueous solutions. Environ.

805

Sci. Technol. 2011, 45 (19), 8559–8566.

806

(123) Cheng, M.; Zeng, G.; Huang, D.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y. Hydroxyl radicals

807

based advanced oxidation processes (AOPs) for remediation of soils contaminated with

808

organic compounds: A review. Chem. Eng. J. 2016, 284, 582–598.

809

ACS Paragon Plus Environment

Page 40 of 40