Selective Transformation of β-Lactam Antibiotics by ... - ACS Publications

Jan 2, 2018 - While the β-lactam antibiotics are known to be susceptible to oxidative degradation by sulfate radical (SO4•–), here we report that...
2 downloads 10 Views 565KB Size
Subscriber access provided by UNIV OF DURHAM

Article

Selective Transformation of #-Lactam Antibiotics by Peroxymonosulfate: Reaction Kinetics and Non-Radical Mechanism Jiabin Chen, Cong Fang, Wenjun Xia, Tianyin Huang, and Ching-Hua Huang Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 02 Jan 2018 Downloaded from http://pubs.acs.org on January 2, 2018

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 35

Environmental Science & Technology

1

Selective Transformation of β-Lactam Antibiotics by

2

Peroxymonosulfate: Reaction Kinetics and Non-Radical Mechanism

3

Jiabin Chen1, Cong Fang1, Wenjun Xia1, Tianyin Huang1,*, Ching-Hua Huang2,*

4 5

1

School of Environmental Science and Engineering, Suzhou University of Science

6 7

and Technology, Suzhou 215001, P. R. China 2

School of Civil and Environmental Engineering, Georgia Institute of Technology,

8

Atlanta, Georgia 30332, United States

9 10

*

Corresponding Authors Phone: 404-894-7694; Fax: 404-358-7087;

11

E-mail: [email protected] (Ching-Hua Huang).

12

Phone: +86 0512 68096895; Fax: +86 0512 68096895.

13

Email: [email protected] (Tianyin Huang).

14 15

Manuscript submitted to

16

Environmental Science & Technology

17

October 30, 2017

18

(Manuscript word count: 5495 + 1500)

19 20 21 22 1

ACS Paragon Plus Environment

Environmental Science & Technology

23

ABSTRACT

24

While the β-lactam antibiotics are known to be susceptible to oxidative degradation by

25

sulfate radical (SO4·-), here we report that peroxymonosulfate (PMS) exhibits specific high

26

reactivity towards β-lactam antibiotics without SO4·- generation for the first time. Apparent

27

second-order reaction constants (k2, app) were determined for the reaction of PMS with three

28

penicillins, five cephalosporins, two carbapenems, and several structurally-related chemicals.

29

The pH-dependency of k2, app could be well modeled based on species-specific reactions. On

30

the basis of reaction kinetics, stoichiometry and structure-activity assessment, the thioether

31

sulfur, on the six- or five-membered rings (penicillins and cephalosporins) and the side chain

32

(carbapenems), was the main reaction site for PMS oxidation. Cephalosporins were more

33

reactive towards PMS than penicillins and carbapenems, and the presence of phenylglycine

34

side chain significantly enhanced cephalosporins’ reactivity towards PMS. Product analysis

35

indicated oxidation of β-lactam antibiotics to two stereoisomeric sulfoxides. A radical

36

scavenging study and electron paramagnetic resonance (EPR) technique confirmed lack of

37

involvement of radical species (e.g., SO4·-). Thus, the PMS-induced oxidation of β-lactam

38

antibiotics was proposed to proceed through a non-radical mechanism involving direct

39

two-electron transfer along with the heterolytic cleavage of the PMS peroxide bond. The new

40

findings of this study are important for elimination of β-lactam antibiotic contamination,

41

because PMS exhibits specific high reactivity and suffers less interference from the water

42

matrix than the radical process.

43 44 2

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

Environmental Science & Technology

45

INTRODUCTION

46

Persulfates (PS), including peroxymonosulfate (HSO5-, PMS) and peroxydisulfate

47

(S2O82-, PDS), have been increasingly considered as alternative oxidants for water treatment

48

and soil remediation.1, 2 They are relatively stable and strong oxidants (E0 = 1.96 V for PDS,

49

and 1.75 V for PMS),3 and the cleavage of their peroxide bonds can generate more reactive

50

radical species, i.e., sulfate radical (SO4·-).4 SO4·- is a strong oxidant with a high redox

51

potential (2.5-3.1 V) and thus can efficiently destruct a wide range of refractory organic

52

contaminants.5 Various strategies have been developed to activate PS for SO4·- generation. For

53

example, heating6, 7 and UV irradiation8, 9 can induce the homolytic cleavage of peroxide bond

54

in PS via intramolecular electron transfer. Transition metal ions activate PS to generate SO4·-

55

through intermolecular electron transfer.10, 11 All these activation strategies require external

56

energy or intensive chemical consumption.

57

Some organic compounds are considered as potential activators for PS. Ahmad et al.

58

reported that soil organic matter (SOM) significantly activated PDS at basic pH, with

59

phenoxide moieties as the potential activator.12 Phenoxide was subsequently confirmed to

60

activate PDS to generate SO4·- via reduction.13 Organic quinones14 and some other phenolic

61

compounds 15 were regarded as potential activators for PDS to produce SO4·-. They were also

62

effective to activate PMS but via a non-radical mechanism.16, 17 Indeed, some recent studies

63

have shown that PS could accept electrons from contaminants with the help of electron

64

mediator without SO4·- formation. Copper oxide (CuO),18 surface loaded-noble metal

65

nanoparticles (e.g., Pd),4 and carbonaceous materials (e.g., carbon nanotubes and graphitized

66

nanodiamonds)3 were reported as effective electron shuttles for PMS activation to avoid the 3

ACS Paragon Plus Environment

Environmental Science & Technology

67

generation of non-selective reactive radicals. This non-radical approach exhibited a lower

68

oxidation potential but higher selectivity towards contaminants, thus providing a promising

69

way for contaminant removal in the water treatment.18 In fact, the direct electron transfer from

70

contaminants, e.g., cationic dyes, to PS was possible, even without the electron shuttles.19

71

Such direct oxidation by PS could maximize the utilization efficiency of PS, and also

72

minimize the adverse impact of water matrix.1 In this work, PMS was also found to

73

effectively oxidize β-lactam antibiotics without electron mediator, catalyst or external energy.

74

The β-lactam antibiotics are among the most frequently used antibiotics worldwide,

75

accounting for 50-70% of the global antibiotic use.20 The high consumption inevitably leads

76

to discharge of these antibiotics into the environment, thus becoming a worldwide concern.

77

Indeed, several commonly used β-lactam antibiotics, e.g., cefalexin (CFX) and amoxicillin

78

(AMX), have been frequently detected in wastewater,21 surface water,22 and coastal water at

79

the concentrations of ng/L - µg/L.23, 24 Bacterial resistance to β-lactam antibiotics has been

80

observed in the environment, which poses potential threats to living organism and human

81

health.25 Various techniques have been explored to eliminate β-lactam antibiotics, including

82

membrane separation,26 activated carbon adsorption,27 photo-degradation,28 and biological

83

degradation.29,

84

degradation owing to the presence of some electron-rich moieties, e.g., primary amine and

85

thioether sulfur.31 Indeed, certain β-lactam antibiotics showed considerable reactivity with

86

various oxidants, e.g., KMnO4,32 ClO2,33 ferrate (VI),34 ozone,35 and peracetic acid.36 SO4·-

87

generated from UV/PDS was very reactive towards β-lactam antibiotics.37 However, the

88

non-selective SO4·- was potentially scavenged by the water matrices (e.g., Cl- and HCO3-),

30

The β-lactam antibiotics are expected to be susceptible to oxidative

4

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35

Environmental Science & Technology

89

thus generating some unexpected by-products (e.g., chlorinated (or brominated) products),38

90

and reducing the efficiency of SO4·-. Note that SO4·- for β-lactam antibiotic destruction was

91

primarily generated from the activated PDS (e.g., UV/PDS) in the previous studies, the

92

reactivity of PMS towards β-lactam antibiotics has always been ignored. As will be discussed

93

later, we discovered that PMS exhibited specific and high reactivity towards β-lactam

94

antibiotics without SO4·- generation, which was almost unaffected by the water matrices, and

95

thus showed high efficiency to eliminate β-lactam antibiotics in real water samples.

96

This work examined a wide range of β-lactam antibiotics, including three penicillins

97

[penicillin G (PG), ampicillin (AMP), and AMX], five cephalosporins [cefapirin (CFP),

98

cephalothin (CFL), cefradine (CFD), cefotaxime (CFT), and CFX] and two carbapenems

99

[meropenem (MPN) and imipenem (IPN)], and several structurally related substructure

100

chemicals [6-aminopenicillanic acid (6-APA), 7-aminocephalosporanic acid (7-ACA),

101

4-pyridineacetic

102

2-(2-Aminothiazole-4-yl)-2-methoxyiminoacetic acid (ATMAA), 2-thiopheneacetic acid, and

103

phenylglycine] (structures shown in Figure 1) to obtain a fundamental understanding of the

104

reactivity of β-lactam antibiotics towards PMS. Based on the reaction kinetics, reaction

105

stoichiometry, product identification, and structure-activity assessment, a two-electron

106

transfer reaction pathway that has not been discovered before for β-lactam antibiotics towards

107

PMS was proposed.

acid

(PTA),

3-methylcrotonic

acid

(3-MCA),

108 109 110

MATERIALS AND METHODS Chemicals. Sources of chemicals are provided in the Supporting Information (SI) Text 5

ACS Paragon Plus Environment

Environmental Science & Technology

111

S1.

112

Real Water Samples. Surface water (SW) from a river, groundwater (GW) from a

113

drinking water well, and wastewater (WW) from the effluent of a municipal wastewater

114

treatment plant were collected at locations in the southeast region of China. Samples were

115

filtered through 0.45-µm glass fiber filters immediately upon arrival in the laboratory and

116

stored at 5 °C before use. The characteristics of these water samples are shown in Table S1.

117

Experimental Procedures. Batch reactions were conducted in 100-mL amber glass

118

bottles wrapped with aluminum foil to prevent light. The solution was constantly mixed by

119

magnetic stirring at room temperature (25 °C). Reaction pH was controlled by 10 mM

120

phosphate buffer for all the experiments, and phosphate ions at 10 mM had no effect on

121

PMS-induced oxidation of β-lactam antibiotics. Reaction was initiated by adding PMS

122

(0.2-1.2 mM) to the solution containing 40 µM of β-lactam antibiotics at pH 7.0. Cl- (1-500

123

mM), HCO3- (1-50 mM), or humic acid (HA, 1-50 mg/L) was added into the above reaction

124

systems to evaluate their impact on PMS-induced oxidation. For the pH impact, the reaction

125

solution was controlled at pH 5.0-10.0 with phosphate buffer. The quenching agent (methanol

126

(MeOH, 1000 mM), tert-butyl alcohol (TBA, 1000 mM), and furfuryl alcohol (FFA, 10 mM))

127

was added to investigate the mechanism. DI water was used in all the above experiments.

128

Sample aliquots were taken at the predetermined time intervals, immediately quenched by

129

excess Na2S2O3. The quenched samples were filtered through 0.45 µm membrane

130

(GVWP01300, Millipore), stored in 2 mL amber vials at 5 °C, and analyzed within 24 h. The

131

same experimental procedure was conducted in evaluating PMS-induced oxidation of CFX in

132

real water matrices at pH 7.0. 6

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

Environmental Science & Technology

133

For comparison, the conventional radical processes, such as Co(II)/PMS and Ag(I)/PDS

134

were also evaluated to degrade β-lactam antibiotics. Co(II) and Ag(I) are regarded as the most

135

efficient metal activators for PMS and PDS, respectively, to generate SO4·-.10 40 µM of Co(II)

136

was added into the DI water containing CFX (40 µM) and phosphate buffer (10 mM, pH 7.0),

137

and then PMS was added to initiate the reaction. For Ag(I)/PDS, the reaction was controlled

138

at pH 3.0 with phosphate buffer owing to the solubility limitation of Ag(I) at neutral pH

139

condition. The experimental procedure was similar to that in PMS-induced oxidation. All the

140

experiments were conducted in duplicate or more.

141

Analytical Methods. A high performance liquid chromatography (HPLC, 1260, Agilent

142

Technology, USA) equipped with a Zorbax SB-C18 column (4.6 × 250 mm, 5 µm), and a UV

143

detector was used to monitor the concentration of β-lactam antibiotics. The mobile phase and

144

detection wavelengths are detailed in Table S2. The HPLC system (Utimate 3000, Dionex,

145

USA) was connected to a triple quadrupole mass spectrometry (TSQ Quantum Ultra EMR,

146

Thermo Fisher Scientific, USA) to analyze the transformation products. The chromatographic

147

and MS conditions are provided in Text S2. The EMX-8/2.7 spectrometer (BRUKER,

148

Germany) was used to analyze the electron paramagnetic resonance (EPR) spectra, with the

149

detailed conditions described in Text S3. A UV/Vis spectrophotometer (UV-1600PC,

150

Shanghai mapada Instruments Co., Ltd., China) was used to quantify PMS with the method

151

proposed by Liang et al.39 as detailed in Text S4.

152 153 154

RESULTS AND DISCUSSION PMS-Induced Rapid Degradation. Previous studies showed that the activated PS 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 35

155

process was efficient to destruct contaminants by the generated SO4·-.10 Our experiments

156

showed that PMS alone induced rapid degradation of β-lactam antibiotics without external

157

energy or catalyst. As Figure 2A shows, CFX ([CFX]0 = 40 µM) was completely degraded

158

after 3 min at pH 7.0 by PMS alone ([PMS]0 = 400 µM), a much faster degradation than that

159

by the conventional radical process, such as the PDS/Ag(I) or PMS/Co(II) systems ([PMS]0 =

160

[metal ion] = 400 µM). On the other hand, the decomposition of PMS in the reaction

161

involving only CFX and PMS was much slower than that in the radical process of PMS/Co(II)

162

(Figure S1), indicating high efficiency of PMS-induced oxidation of CFX than the radical

163

process. Moreover, while PMS alone showed high reactivity towards CFX, it was inert for the

164

degradation of anisole, which was frequently used as a probe for SO4·- and HO· (kHO· = 5.4 ×

165

109 M-1.s-1, kSO4-· = 4.9 × 109 M-1.s-1)40, 41 (data not shown). Hence, PMS exhibited a specific

166

reactivity towards CFX, likely via a different mechanism from radical-induced oxidation.

167

The effect of PMS concentration on CFX degradation was evaluated at pH 7.0. The

168

degradation of CFX followed first-order kinetics at different initial concentrations of PMS

169

(Figure S2A). The pseudo-first-order rate constant (kobs in equation (1)) increased as the

170

PMS/CFX ratio was increased from 5 to 30. Plot of log (kobs) vs log (CPMS) exhibited a linear

171

relationship with the slope close to unity (Figure S2B). Hence kobs was first-order with respect

172

to PMS concentration. Apparent second-order rate constants (k2,app in equation (2)) were then

173

obtained to be 71.7 ± (3.76) M-1.s-1 by dividing kobs by the concentration of PMS.

174



[] 

= k  ∙ [CFX]

175



[] 

= k , ∙ [CFX] ∙ [PMS]

176

(1) (2)

We further investigated the effect of common water matrix (e.g., Cl-, HCO3- and HA) on 8

ACS Paragon Plus Environment

Page 9 of 35

Environmental Science & Technology

177

PMS-induced CFX degradation (Figure S3). The addition of HCO3- or HA showed negligible

178

effect on CFX degradation, while the presence of Cl- exhibited a slight promoting effect,

179

which increased with increasing concentration of Cl-. Cl- was supposed to react with PMS to

180

generate oxidative species, e.g., HClO, thus promoting the contaminant degradation.42 The

181

PMS-induced CFX degradation was further evaluated in real water matrices, e.g., SW, GW,

182

and WW samples. Results showed that the rapid degradation of CFX could be observed in the

183

real water samples (Figure S4). Compared to the DI water, PMS-induced degradation of CFX

184

was only slightly affected in SW, GW, and WW matrices. Therefore, PMS is highly selective

185

to oxidize CFX, thus a promising technology to remove CFX in real water matrices.

186

pH Impact. The impact of pH was examined for PMS-induced oxidation of CFX at pH

187

5.0-10.0. Generally, k2,app-CFX slightly decreased with increasing pH from 5.0 to 8.0, and

188

sharply decreased as pH was further increased to 9.0 and above (Figure 2B). CFX has two

189

pKa values at 2.56 and 6.88, corresponding to the carboxyl group on the dihydrothiazine ring

190

and the primary amine group on the phenylglycine side chain, respectively.28 PMS also

191

possesses two pKa values (pKa1 < 0, pKa2 = 9.4).8, 43 The speciation of PMS or CFX is strongly

192

influenced by solution pH (CFX  ⇌ CFX  + H  , pKa2, CFX = 6.88; HSO ⇌ H  + SO , pKa2,

193

PMS

194

reactivity. For example, protonated phenylglycine primary amine in CFX always shows

195

negligible reactivity towards oxidants, such as Cu(II)44 and ferrate(VI);34 while SO52- is less

196

reactive than HSO5-.19 Hence, the species-specific reactions between CFX and PMS species

197

were thus used to account for the pH-dependent variation of k2, app-CFX. The PMS-induced

198

oxidation of CFX could be expressed as follows:

= 9.4; Figure S5A) Note that different species of CFX and PMS might exhibit different

9

ACS Paragon Plus Environment

Environmental Science & Technology



d["#$] % = k , ∙ [CFX] % ∙ [PMS] % d&

*,+

= ' k (,) ∙ α( ∙ β) . ["#$] % ∙ [PMS] % (3) (,)

199

where [CFX]T and [PMS]T represent the total concentration of CFX and PMS, respectively;

200

ki,j is the species-specific second-order rate constant for the reaction between species i of CFX

201

and species j of PMS; and αi and βj represent the equilibrium distribution coefficients of CFX

202

and PMS species, respectively. The species-specific second-order rate constants (ki,j) were

203

determined by fitting the experimental data of k2,app-CFX to equation (4) with least-squares

204

nonlinear regressions by the Origin software. *,+

k , = ' k (,) ∙ α( ∙ β) (4) (,)

205

Note that the reaction between CFX0 and SO52- species was negligible owing to the

206

lower oxidation capacity of SO52- than HSO5-19 and the extremely low mole fractions of α1 ×

207

β2, i.e., CFX0-SO52- species (Figure S5B); thus, this reaction was not considered when fitting

208

experimental data to the kinetic model. As Figure 2B shows, the experimental k2,app-CFX across

209

the pH range could be well explained by the kinetic modeling (R2 = 0.9789), with the

210

obtained species-specific rate constants (ki,j) summarized in Table S3. The contribution of the

211

apparent species-specific reaction rate constants to the overall k2, app-CFX is also depicted in

212

Figure 2B. The results showed that the contribution of SO52- species to the overall reaction

213

was insignificant at pH lower than 8.0. SO52- exhibited relatively low reactivity to CFX (kCFX--

214

SO 52-

215

SO52- than HSO5-.19, 45 Compared to CFX- species, CFX0 was more susceptible to HSO5-

216

oxidation (76.81 ± (1.91) M-1.s-1 (kCFX0-HSO5-) vs 69.44 ± (2.18) M-1.s-1 (kCFX--HSO5-)). This result

= 8.01 ± (3.25) M-1.s-1), which could be attributed to the lower oxidation capacity of

10

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

Environmental Science & Technology

217

was opposite to previous reports that CFX0 was unreactive to oxidants such as Cu(II).44 It was

218

thus implied that reaction site on CFX by PMS oxidation was likely different from that by

219

Cu(II) oxidation, which will be further discussed in the following section. The different

220

reactivity trend observed in this study might be explained by electrostatic interactions. Indeed,

221

PMS (HSO5-) was previously reported to be effective for cationic dye degradation but inert to

222

anionic dye oxidation owing to electrostatic repulsion.19 In analogy, PMS (HSO5-) could get

223

close to the neutral CFX0 more easily than to the negatively-charged CFX-; thus, PMS

224

exhibited higher reactivity to CFX0. For PMS-induced oxidation of AMP and PG, the

225

pH-dependent k2,

226

species-specific reactions (Text S5, Figures S6 and S7).

app

could also be well explained by the kinetic modeling based on

227

Structure-Dependent Degradation. The PMS-induced degradation was further

228

investigated for many β-lactam antibiotics, including cephalosporins, penicillins, and

229

carbapenems, at pH 7.0. For all the selected compounds, the PMS-induced degradation

230

followed pseudo-first-order kinetics. The linear relationship between log (kobs) and log (CPMS)

231

with the slope close to unity confirmed their second-order reaction (Figures S8-S10), with the

232

obtained k2,app summarized in Figure 3. Generally, these k2,app values ranged from 2.0 to 71.7

233

M-1.s-1, which were comparable to k2,app of β-lactam antibiotics with peracetic acid

234

(15.56-44.38 M-1.s-1),36 but were smaller than those for their reaction with ferrate(VI)

235

(110-770 M-1.s-1),34 ozone (103-106 M-1.s-1),46 and radical species such as SO4·- and HO· (109

236

M-1.s-1).47

237

As an oxidant, PMS is expected to attack the atoms/moiety with available electrons. The

238

thioether sulfur in the six- or five-membered ring, and the double bond in the six-membered 11

ACS Paragon Plus Environment

Environmental Science & Technology

239

dihydrothiazine ring are electron-rich and thus are regarded as potential reactive sites for

240

various oxidants such as SO4·-,47 O3,35 and MnO4-.32 If present, the phenylglycine primary

241

amine on the side chain is also electron-rich, reactive towards oxidants such as Cu(II)44 and

242

ferrate(VI).34 We thus hypothesized that the above three moieties were potential reactive sites

243

for PMS oxidation. To assess whether the double bond on the six-membered ring of

244

cephalosporins was a reactive site to PMS, we investigated the reactivity of the substructure

245

compound 3-MCA towards PMS. The finding that PMS could not induce degradation of

246

3-MCA (Figure 3A) suggested that the double bond on the six-membered dihydrothiazine

247

ring was not reactive to PMS. The study to assess the reactivities of phenylglycine primary

248

amine and thioether sulfur are discussed below.

249

Phenylglycine primary amine moiety. As Figure 3B shows, the phenylglycine-type

250

cephalosporins were more susceptible to oxidation by PMS than their non-phenylglycine-type

251

counter parts (e.g., CFX/CFD vs CFL/CFT), suggesting that the phenylglycine side chain

252

might be a reactive site for PMS. However, the reactivity of penicillins towards PMS seems to

253

be not affected by the presence of phenylglycine side chain (e.g., AMP/AMX vs PG). In

254

addition, when phenylglycine was tested for its reactivity towards PMS, degradation of

255

phenylglycine was negligible after 1 h (Figure 3A), indicating very low reactivity towards

256

PMS. The above results together indicated that the phenylglycine moiety was not a reaction

257

site for PMS, but may affect the PMS-induced oxidation of cephalosporins through altering

258

molecular electron distribution. According to the density functional theory (DFT) calculation,

259

the highest occupied molecular orbital (HOMO), characteristic of compound nucleophilicity,

260

is located at the six-membered dihydrothiazine ring for cephalosporins and the 12

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35

Environmental Science & Technology

261

five-membered thiazolidine ring for penicillins.36 The presence of phenylglycine moiety

262

increases the HOMO energy, and thus nucleophilicity at the six-membered ring of

263

cephalosporins.36 This could be the reason why the phenylglycine-type cephalosporins

264

showed higher reactivity towards PMS than the non phenylglycine-type cephalosporins,

265

similar to the reactivity trend of cephalosporins toward peracetic acid. In the case of

266

penicillins, the presence of phenylglycine moiety increases the HOMO energy but also shifts

267

the HOMO location away from the 5-membered thiazolidine ring, thus an overall less impact

268

of the phyenylglycine group on the compound reactivity was observed.36

269

In addition to phenylglycine, we also investigated the reactivity of PTA, ATMAA and

270

2-thiopheneacetic acid, substructure of CFP, CFT and CFL, respectively, towards PMS.

271

ATMAA and 2-thiopheneacetic acid were found to be inert towards PMS (Figure 3A). Hence,

272

CFT and CFL exhibited comparable reactivity towards PMS, and the reactive site was located

273

on the six-membered dihydrothiazine ring. PTA exhibited considerable reactivity towards

274

PMS, with k2,app = 6.8 ± (0.62) M-1.s-1. Hence, PTA significantly contributed to the higher

275

reactivity of CFP towards PMS than other non-phenylglycine cephalosporins.

276

Thioether sulfur. The thioether sulfur on the β-lactams’ five- or six-membered ring is

277

known to be susceptible to electrophilic attack by various oxidants.31 Thus, the thioether

278

sulfur on the ring was most likely the main reaction site for PMS. The thioether sulfur on the

279

six-membered dihydrothiazine ring showed higher reactivity towards PMS than that on the

280

five-membered thiazolidine ring (e.g., 6-APA vs 7-ACA). This difference could be explained

281

by the steric hindrance present in the latter. On the five-membered thiazolidine ring, the two

282

methyl groups adjacent to thioether might hinder the attack of oxidant on the thioether sulfur, 13

ACS Paragon Plus Environment

Environmental Science & Technology

283

thus contributing to the observed lower reactivity to PMS.34

284

Compared to penicillins and cephalosporins, the transformation of carbapenems has been

285

largely overlooked before. Although the thioether sulfur is not present on the five-membered

286

ring of carbapenems (i.e., MPN and IPN), PMS-induced rapid degradation was still observed

287

(Figure 3B), probably due to the thioether sulfur present in the side chain of carbapenems.

288

This hypothesis was supported by the considerable reactivity of PTA towards PMS. Note that

289

MPN reacted with PMS at 9.5 times as fast as IPN. Since the structural difference of MPN

290

and IPN only lies in the neighboring moiety linked to the side-chain thioether sulfur, the

291

side-chain thioether sulfur is important for carbapenems’ reactivity towards PMS and the

292

impact of the neighboring moiety to the side-chain thioether sulfur will be further elucidated

293

in a later section.

294

Reaction stoichiometry. Reaction stoichiometry could be determined from reaction ratio

295

between PMS and β-lactam antibiotics, because their oxidation products were not reactive

296

towards PMS (see the following section about transformation products). Decomposition of

297

PMS was monitored along with the degradation of β-lactam antibiotics. As Figure S11 shows,

298

near 10% of initial PMS was decomposed after the complete degradation of penicillins and

299

cephalosporins, except for CFP. Note that the initial concentration of PMS was 10 times

300

higher than that of β-lactam antibiotics and thus the reaction stoichiometry between PMS and

301

antibiotics was around 1:1. This result implied that only one reaction site was present on

302

β-lactam antibiotics for PMS oxidation, most likely the thioether sulfur on the six

303

(five)-membered rings. For CFP, the deviation of 1:1 reaction stoichiometry suggests that an

304

additional reaction site was present, most probably the thioether sulfur on the side chain. For 14

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

Environmental Science & Technology

305

carbapenems, however, the decomposition loss of PMS was around 20% and thus the reaction

306

stoichiometry between PMS and carbapenems was close to 2:1. This result implied that two

307

reaction sites were likely present on the carbapenem antibiotics.

308

Overall, the thioether sulfur, on the six- or five-membered rings and the side chain, was

309

the reaction site for PMS. The thioether sulfur on the six-membered dihydrothiazine ring

310

exhibited higher reactivity than that on the five-membered thiazolidine ring. Although the

311

phenylglyicine primary amine was not a direct reaction site, it could significantly enhance the

312

reactivity of the thioether sulfur on the six-membered dihydrothiazine ring towards PMS.

313

Transformation products. To facilitate discerning the reaction mechanism, the

314

transformation products of β-lactam antibiotics by PMS were analyzed by LC/MS/MS. The

315

primary products were with molecular weight (MW) of M+16 and M+32 (M: MW of the

316

parent β-lactam antibiotic) (Table S4). For most cephalosporin antibiotics except CFP, only

317

two MW 363 (M+16) products were observed. The 363a and 363b product might be isomers

318

because they had different LC retention times but the same MW and MS fragment patterns.

319

Previous study has reported two MW 363 products for ferrate(VI) oxidation of CFX to be the

320

stereoisomeric CFX-(R)-sulfoxide and CFX-(S)-sulfoxide.34 We subsequently analyzed the

321

oxidation products of CFX by ferrate(VI) using the analytical methods in this study and found

322

that the stereoisomeric CFX-sulfoxides generated by ferrate(VI) oxidation matched with the

323

363a and 363b products in LC retention time and MS spectra. This evidence verified 363a and

324

363b as the CFX-sulfoxides, with the thioether sulfur on the six-membered dihydrothiazine

325

ring oxidized by PMS (Figure 4). Compared to PMS-induced oxidation of CFX, much more

326

oxidation products were previously reported in the oxidation of CFX by SO4·-, including 15

ACS Paragon Plus Environment

Environmental Science & Technology

327

oxidation of amine and thioether, hydroxylation of benzene and double bond, and β-lactam

328

ring opening of CFX.37

329

Different from the other cephalosporin antibiotics, oxidation of CFP by PMS yielded

330

three M+16 products and one M+32 product. Among the M+16 products, two products might

331

be the stereoisomeric sulfoxides of CFP from oxidation of the thioether sulfur on the

332

six-membered ring, and the other one likely from oxidation of the PTA side chain thioether

333

sulfur by PMS. For the M+32 product, both thioether sulfurs, on the six-membered ring and

334

the PTA side chain, were oxidized to the sulfoxides (Text S6, Figure S12 and S13).

335

For the PMS-induced oxidation of penicillin antibiotics, i.e., PG, AMP and AMX, two

336

M+16 products were also the only identified oxidation products. Similarly, they were

337

confirmed to be the corresponding sulfoxide products by matching their retention times and

338

MS spectra with those observed in the ferrate(VI)-induced oxidation.

339

Oxidation of the carbapenem antibiotic MPN by PMS yielded two M+16 products and

340

two M+32 products. Both pairs were likely the isomers because of the same MS fragments

341

but different retention times. Similar to MPN, the M+16 products also exhibited the MS

342

fragments of m/z 141 and 114, which were characteristics of the pyrrolidine side chain fused

343

to thioether sulfur and the β-lactam ring, respectively (Figure 4, Figure S14A and S14B). It

344

was thus suggested that the M+16 products possess intact pyrrolidine side chain and β-lactam

345

ring, and the reactive sites on MPN for PMS might be the thioether sulfur and the double

346

bond of the dihydropyrrole ring. However, since 3-MCA showed no reactivity towards PMS

347

(Figure 3A), the double bond on the dihydropyrrole ring was unlikely a reactive site for PMS.

348

Thus, the thioether sulfur on MPN was oxidized by PMS, generating the M+16 stereoisomeric 16

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

Environmental Science & Technology

349

sulfoxide products. The M+32 products were previously proposed as the sulfone products for

350

oxidation of penicillins or cephalosporins.34 If the sulfone products of MPN were generated,

351

the fragments of m/z 68 and 141, representative of the intact pyrrolidine side chain, were

352

expected in the MS spectra of M+32 products. However, m/z 84 and 157 fragments were

353

observed instead from the M+32 products (Figure S14C). The occurrence of m/z 84 and 157

354

implied hydroxylation of the side chain pyrrolidine ring, that is, 84 = 68 + 16 and 157 = 141 +

355

16. All the above evidence suggested the thioether sulfur of MPN was susceptible to oxidation

356

by PMS to generate stereoisomeric sulfoxides, but without further oxidation to sulfone

357

products. The much higher intensity of the sulfoxide products than that of the M+32 products

358

(Figure S15) also indicated that the thioether sulfur was the main reactive site on MPN for

359

PMS.

360

Proposed mechanism. The conventional PS activation process always relies on SO4·-

361

generation for contaminant degradation. SO4·- exhibits high reactivity towards β-lactam

362

antibiotics (up to 109 M-1.s-1)47 to generate diverse products.48 In this work, PMS-induced

363

oxidation of β-lactam antibiotics primarily generated the M+16 products, quite different from

364

the SO4·--induced oxidation.37 Moreover, the reaction stoichiometry of PMS with penicillins

365

and most cephalosporins (except CFP) was around 1:1, contrary to the fact that SO4·- reacts

366

non-selectively with the antibiotics as well as coexisting solutes or even water molecules. All

367

the experimental evidence indicated that SO4·- was not involved in the oxidation of β-lactam

368

antibiotics by PMS. This hypothesis was further verified by the radical scavenging study and

369

EPR technique. Specifically, the PMS-induced oxidation of CFX was not affected after

370

addition of excess radical scavengers, i.e., MeOH and TBA (Figure S16). Moreover, EPR 17

ACS Paragon Plus Environment

Environmental Science & Technology

371

analysis unambiguously demonstrated that SO4·- and HO· were not generated (Figure S17).

372

Thus, the PMS induced-oxidation of β-lactam antibiotics underwent via the non-radical

373

process rather than radical process. In the real water samples, such as WW, water matrix

374

could not activate PMS to generate radicals, on the basis that PMS decomposition was

375

negligible by water matrix alone, and both CFX degradation and PMS decomposition were

376

not affected by excess MeOH (Figure S18). Hence, the oxidation of CFX by PMS was still

377

via the non-radical process in real water matrix. Previous study reported that 1O2 likely acted

378

as the reactive species in the non-radical activation of PMS by BQ.17 However, addition of

379

FFA, an efficient 1O2 scavenger, did not influence the PMS-induced degradation of CFX

380

(Figure S16); thus, the involvement of 1O2 was excluded (Text S7).

381

The non-radical process might proceed through two-electron transfer (i.e., oxygen atom

382

transfer), which involves the heterolytic breakage of the peroxide bond in PMS and an oxygen

383

atom transfer from PMS to nucleophiles (Nu) (Scheme A in Figure 5).1, 49 The two-electron

384

transfer mechanism was verified by the 1:1 reaction stoichiometry between the thioether

385

sulfur and PMS. The oxidation of alkyl sulfides by PMS was previously assumed to proceed

386

via the attack of peroxide on the sulfur atom to form an intermediate, and the subsequent

387

rate-limiting heterolysis (and rearrangement) of the intermediate to form the sulfoxide.50

388

Similarly, a plausible mechanism for the oxidation of β-lactam antibiotics by PMS likely

389

underwent via the electrophilic attack of the peroxide on the thioether sulfur to generate an

390

intermediate, which was further decomposed to sulfoxide products by the rate-limiting

391

heterolysis and rearrangement (Scheme B in Figure 5). This reaction mechanism only

392

involves two electron transfer from the thioether sulfur to peroxide bond without formation of 18

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35

Environmental Science & Technology

393

radical species. Compared to the typical radical process, all of the oxidation capacity of PMS

394

was utilized in this non-radical process, rendering this technique cost-effective and highly

395

efficient for the oxidation of β-lactam antibiotics.

396

In fact, the above mechanism involving heterolysis and rearrangement was primarily

397

proposed for the mono-substituted peroxide, e.g., PMS.49 Unlike PMS, the rearrangement of

398

the intermediate between PDS and thioether sulfur was not likely to occur owing to the lack

399

of proton in PDS molecule. Hence, PDS exhibited low reactivity towards the thioether sulfur

400

in β-lactam antibiotics (Figure 2A). In this work, the sulfoxide products of β-lactam

401

antibiotics were the end products by PMS oxidation without further oxidation to a sulfone

402

product, quite different from the previous studies that the sulfoxide product (DMSO) of

403

dimethyl sulfide (DMS) was further oxidized to the corresponding sulfone product

404

(DMSO2).50 Compared to the simple DMSO molecule, it might be difficult for the second

405

PMS molecule to approach the sulfoxide sulfur on the strained five (or six)-membered ring

406

owing to the steric hindrance effect. Consequently, the stoichiometry reaction between the

407

thioether sulfur and PMS was around 1:1, and the PMS decomposition was almost ceased

408

after complete conversion of thioether sulfur to sulfoxide.

409

The reactivity of reduced sulfurs towards PMS was positively related to their electron

410

density, e.g., DES (DMS) < H2S < HS-.50 The fusion of β-lactam ring with the five (or

411

six)-membered ring made the molecule structure more strained, thus resulting in

412

non-planarity of the molecule with large angle and torsional rotation.31 As a result, more

413

electron density of the thioether sulfur on the ring is exposed outside, thus exhibiting high

414

reactivity towards oxidants, e.g., PMS. This hypothesis was verified by the larger k2, app for 19

ACS Paragon Plus Environment

Environmental Science & Technology

415

penicillins and cephalosporins than those for carbapenem antibiotics or alkyl sulfides.50

416

Although the thioether sulfur is located at the side chain of MPN rather than on the ring, its

417

linkage with two five-membered rings makes the electron density more exposed to the

418

electrophilic attack, rendering the thioether sulfur on MPN more reactive towards PMS than

419

that on IPN.

420

Environmental Significance. To our best knowledge, this is the first study to report the

421

high reactivity and fast reaction of β-lactam antibiotics with PMS not relying on SO4·-

422

generation, highlighting a new and facile strategy for alleviating β-lactam antibiotic

423

contamination. Compared to the conventional PMS activation with SO4·- generation,

424

PMS-induced degradation exhibits low cost because no external energy or catalyst is required,

425

and high efficiency because of complete utilization of the oxidation capacity of PMS. In

426

contrast to the non-selective reactivity of SO4·-, PMS shows specific high reactivity towards

427

the thioether sulfur on β-lactam antibiotics, which is almost unaffected by background ions

428

and natural organic matter in water matrices. Hence, PMS maintains remarkable treatment

429

efficiency for β-lactam antibiotics in complicated environmental matrices, e.g., wastewater.

430

The β-lactam ring is responsible for the antibacterial activity of β-lactam antibiotics,51

431

and hydrolytic cleavage of the β-lactam ring could eliminate antibacterial activity.52 However,

432

the hydrolysis of β-lactam antibiotics is dependent on compound structure, pH and

433

temperature.53 Compared to the rapid oxidation of β-lactam antibiotics by PMS, the

434

hydrolysis could be neglected at environmental conditions. Although photodegradation may

435

be the most important route to eliminate cephalosporins in surface waters (t1/2 up to hours),

436

their transformation products were found to be more toxic than the parent cephalosporins.28 20

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35

Environmental Science & Technology

437

Recently, various oxidants, such as ferrate(VI),34 ozone,54 chlorine,55 and advanced oxidation

438

processes56, 57 have been investigated for eliminating β-lactam antibiotics and their biological

439

activity. The remarkable reduction of antibacterial activity of β-lactam antibiotics was

440

observed after treatment by ferrate(VI)34 and ozone.54 However, photocatalytic transformation

441

products of AMX with UV/TiO2 still exhibited significant activity to Enterococcus faecalis

442

(ATCC 14506), 57 and the oxidation of AMX by a photo-Fenton process (Fe2+/H2O2/stimulated

443

solar light) generated even more toxic products than AMX based on Daphnia magna

444

biossays.56 Free available chlorine (FAC) is among the most frequently utilized disinfectants,

445

but the oxidation of β-lactam antibiotics, e.g., cefazolin, by chlorine generated chlorinated

446

products, which might possess high genotoxicity.55 The rapid oxidation of β-lactam antibiotics

447

by PMS generated the stereoisomeric sulfoxide products, which were reported to exhibit

448

significantly lower antibacterial activity than the parent β-lactam antibiotics. For example,

449

PG-(R)-sulfoxide and CFX-(R)-sulfoxide were determined to be ∼15% and ∼83% as active

450

as PG and CFX, respectively, against B. subtilis ATTC 6051; while their corresponding

451

(S)-sulfoxides were found to possess less than 1% of their parent β-lactam antibiotics’

452

antibacterial activity.54 Therefore, PMS oxidation may effectively reduce the antibacterial

453

activity of β-lactam antibiotics, providing a promising way to rapidly eliminate β-lactam

454

antibiotics and significantly reduce their biological activity in water treatment.

455 456

ASSOCIATED CONTENT

457

Supporting Information.

458

Text S1-S7, Tables S1-S4 and Figures S1-S18. This material is available free of charge via the 21

ACS Paragon Plus Environment

Environmental Science & Technology

459

Internet at http://pubs.acs.org.

460 461

ACKNOWLEDGMENTS

462

We sincerely thank the National Natural Science Foundation of China (51509175), Science

463

and Technology Planning Project of Suzhou (SS201666, SS201722) for financially supporting

464

this work.

465 466

REFERENCES

467

(1) Wang, Z.; Bush, R. T.; Sullivan, L. A.; Chen, C.; Liu, J. Selective oxidation of arsenite by

468

peroxymonosulfate with high utilization efficiency of oxidant. Environ. Sci. Technol. 2014, 48

469

(7), 3978-3985.

470

(2) Liu, H.; Bruton, T. A.; Doyle, F. M.; Sedlak, D. L. In situ chemical oxidation of

471

contaminated groundwater by persulfate: Decomposition by Fe(III)-and Mn(IV)-containing

472

oxides and aquifer materials. Environ. Sci. Technol. 2014, 48 (17), 10330-10336.

473

(3) Lee, H.; Kim, H. I.; Weon, S.; Choi, W.; Hwang, Y. S.; Seo, J.; Lee, C.; Kim, J. H.

474

Activation of persulfates by graphitized nanodiamonds for removal of organic compounds.

475

Environ. Sci. Technol. 2016, 50 (18), 10134-10142.

476

(4) Ahn, Y. Y.; Yun, E. T.; Seo, J. W.; Lee, C.; Kim, S. H.; Kim, J. H.; Lee, J. Activation of

477

peroxymonosulfate by surface-loaded noble metal nanoparticles for oxidative degradation of

478

organic compounds. Environ. Sci. Technol. 2016, 50 (18), 10187-10197.

479

(5) Zhang, B. T.; Zhang, Y.; Teng, Y. H.; Fan, M. H. Sulfate radical and its application in

480

decontamination technologies. Crit. Rev. Env. Sci. Tec. 2015, 45 (16), 1756-1800. 22

ACS Paragon Plus Environment

Page 22 of 35

Page 23 of 35

Environmental Science & Technology

481

(6) Johnson, R. L.; Tratnyek, P. G.; Johnson, R. O. B. Persulfate persistence under thermal

482

activation conditions. Environ. Sci. Technol. 2008, 42 (24), 9350-9356.

483

(7) Waldemer, R. H.; Tratnyek, P. G.; Johnson, R. L.; Nurmi, J. T. Oxidation of chlorinated

484

ethenes by heat-activated persulfate: Kinetics and products. Environ. Sci. Technol. 2006, 41

485

(3), 1010-1015.

486

(8) Guan, Y. H.; Ma, J.; Li, X. C.; Fang, J. Y.; Chen, L. W. Influence of pH on the formation

487

of sulfate and hydroxyl radicals in the UV/peroxymonosulfate system. Environ. Sci. Technol.

488

2011, 45 (21), 9308-9314.

489

(9) Qian, Y.; Guo, X.; Zhang, Y.; Peng, Y.; Sun, P.; Huang, C. H.; Niu, J.; Zhou, X.; Crittenden,

490

J. C. Perfluorooctanoic acid degradation using UV–persulfate process: Modeling of the

491

degradation and chlorate formation. Environ. Sci. Technol. 2016, 50 (2), 772-781.

492

(10) Anipsitakis, G. P.; Dionysiou, D. D. Radical generation by the interaction of transition

493

metals with common oxidants. Environ. Sci. Technol. 2004, 38 (13), 3705-3712.

494

(11) Zou, J.; Ma, J.; Chen, L. W.; Li, X. C.; Guan, Y. H.; Xie, P. C.; Pan, C. Rapid

495

acceleration of ferrous iron/peroxymonosulfate oxidation of organic pollutants by promoting

496

Fe(III)/Fe(II) cycle with hydroxylamine. Environ. Sci. Technol. 2013, 47 (20), 11685-11691.

497

(12) Ahmad, M.; Teel, A. L.; Watts, R. J. Persulfate activation by subsurface minerals. J.

498

Contam. Hydrol. 2010, 115 (1–4), 34-45.

499

(13) Ahmad, M.; Teel, A. L.; Watts, R. J. Mechanism of persulfate activation by phenols.

500

Environ. Sci. Technol. 2013, 47 (11), 5864-5871.

501

(14) Fang, G. D.; Gao, J.; Dionysiou, D. D.; Liu, C.; Zhou, D. M. Activation of persulfate by

502

quinones: Free radical reactions and implication for the degradation of PCBs. Environ. Sci. 23

ACS Paragon Plus Environment

Environmental Science & Technology

503

Technol. 2013, 47 (9), 4605-4611.

504

(15) Fang, G.; Liu, C.; Gao, J.; Dionysiou, D. D.; Zhou, D. Manipulation of persistent free

505

radicals in biochar to activate persulfate for contaminant degradation. Environ. Sci. Technol.

506

2015, 49 (9), 5645.

507

(16) Zhou, Y.; Jiang, J.; Gao, Y.; Pang, S. Y.; Yang, Y.; Ma, J.; Gu, J.; Li, J.; Wang, Z.; Wang,

508

L. H.; Yuan, L. P.; Yang, Y. Activation of peroxymonosulfate by phenols: Important role of

509

quinone intermediates and involvement of singlet oxygen. Water Res. 2017, 125, 209-218.

510

(17) Zhou, Y.; Jiang, J.; Gao, Y.; Ma, J.; Pang, S. Y.; Li, J.; Lu, X. T.; Yuan, L. P. Activation of

511

peroxymonosulfate by benzoquinone: A novel nonradical oxidation process. Environ. Sci.

512

Technol. 2015, 49 (21), 12941-12950.

513

(18) Zhang, T.; Chen, Y.; Wang, Y. R.; Le Roux, J.; Yang, Y.; Croue, J. P. Efficient

514

peroxydisulfate activation process not relying on sulfate radical generation for water pollutant

515

degradation. Environ. Sci. Technol. 2014, 48 (10), 5868-5875.

516

(19) Lei, Y.; Chen, C. S.; Ai, J.; Lin, H.; Huang, Y. H.; Zhang, H. Selective decolorization of

517

cationic dyes by peroxymonosulfate: non-radical mechanism and effect of chloride. RSC Adv.

518

2016, 6 (2), 866-871.

519

(20) Chen, J.; Wang, Y.; Qian, Y.; Huang, T. Fe(III)-promoted transformation of β-lactam

520

antibiotics: Hydrolysis vs oxidation. J. Hazard. Mater. 2017, 335, 117-124.

521

(21) Leung, H. W.; Minh, T. B.; Murphy, M. B.; Lam, J. C. W.; So, M. K.; Martin, M.; Lam,

522

P. K. S.; Richardson, B. J. Distribution, fate and risk assessment of antibiotics in sewage

523

treatment plants in Hong Kong, South China. Environ. Int. 2012, 42 (0), 1-9.

524

(22) Cha, J. M.; Yang, S.; Carlson, K. H. Trace determination of beta-lactam antibiotics in 24

ACS Paragon Plus Environment

Page 24 of 35

Page 25 of 35

Environmental Science & Technology

525

surface water and urban wastewater using liquid chromatography combined with electrospray

526

tandem mass spectrometry. J. chromatogr., A 2006, 1115 (1-2), 46-57.

527

(23) Gulkowska, A.; He, Y.; So, M. K.; Yeung, L. W. Y.; Leung, H. W.; Giesy, J. P.; Lam, P. K.

528

S.; Martin, M.; Richardson, B. J. The occurrence of selected antibiotics in Hong Kong coastal

529

waters. Mar. Pollut. Bull. 2007, 54 (8), 1287-1293.

530

(24) Li, S.; Shi, W.; Liu, W.; Li, H.; Zhang, W.; Hu, J.; Ke, Y.; Sun, W.; Ni, J. A duodecennial

531

national synthesis of antibiotics in China's major rivers and seas (2005–2016). Sci. Total

532

Environ. 2017, 615, 906-917.

533

(25) Huang, T.; Fang, C.; Qian, Y.; Gu, H.; Chen, J. Insight into Mn(II)-mediated

534

transformation of β-lactam antibiotics: The overlooked hydrolysis. Chem. Eng. J. 2017, 321,

535

662-668.

536

(26) Cheng, X. Q.; Zhang, C.; Wang, Z. X.; Shao, L. Tailoring nanofiltration membrane

537

performance for highly-efficient antibiotics removal by mussel-inspired modification. J.

538

Membrane Sci. 2016, 499, 326-334.

539

(27) Ahmed, M. J.; Theydan, S. K. Adsorption of cephalexin onto activated carbons from

540

Albizia lebbeck seed pods by microwave-induced KOH and K2CO3 activations. Chem. Eng. J.

541

2012, 211–212, 200-207.

542

(28) Wang, X. H.; Lin, A. Y. C. Phototransformation of cephalosporin antibiotics in an

543

aqueous environment results in higher toxicity. Environ. Sci. Technol. 2012, 46 (22),

544

12417-12426.

545

(29) Li, B.; Zhang, T. Biodegradation and adsorption of antibiotics in the activated sludge

546

process. Environ. Sci. Technol. 2010, 44 (9), 3468-3473. 25

ACS Paragon Plus Environment

Environmental Science & Technology

547

(30) Watkinson, A. J.; Murby, E. J.; Costanzo, S. D. Removal of antibiotics in conventional

548

and advanced wastewater treatment: Implications for environmental discharge and wastewater

549

recycling. Water Res. 2007, 41 (18), 4164-4176.

550

(31) Deshpande, A. D.; Baheti, K. G.; Chatterjee, N. R. Degradation of β-lactam antibiotics.

551

Curr. Sci. 2004, 87 (12), 1684-1695.

552

(32) Li, L.; Wei, D.; Wei, G.; Du, Y. Oxidation of cefazolin by potassium permanganate:

553

Transformation products and plausible pathways. Chemosphere 2016, 149, 279-285.

554

(33) Navalon, S.; Alvaro, M.; Garcia, H. Reaction of chlorine dioxide with emergent water

555

pollutants: Product study of the reaction of three β-lactam antibiotics with ClO2. Water Res.

556

2008, 42 (8–9), 1935-1942.

557

(34) Karlesa, A.; De Vera, G. A. D.; Dodd, M. C.; Park, J.; Espino, M. P. B.; Lee, Y.

558

Ferrate(VI) oxidation of β-Lactam antibiotics: Reaction kinetics, antibacterial activity

559

changes, and transformation products. Environ. Sci. Technol. 2014, 48 (17), 10380-10389.

560

(35) Dodd, M. C.; Kohler, H. P. E.; Von Gunten, U. Oxidation of antibacterial compounds by

561

ozone and hydroxyl radical: Elimination of biological activity during aqueous ozonation

562

processes. Environ. Sci. Technol. 2009, 43 (7), 2498-2504.

563

(36) Zhang, K.; Zhou, X.; Du, P.; Zhang, T.; Cai, M.; Sun, P.; Huang, C. H. Oxidation of

564

β-lactam antibiotics by peracetic acid: Reaction kinetics, product and pathway evaluation.

565

Water Res. 2017, 123, 153-161.

566

(37) Serna. Galvis, E. A.; Ferraro, F.; SilvaAgredo, J.; TorresPalma, R. A. Degradation of

567

highly consumed fluoroquinolones, penicillins and cephalosporins in distilled water and

568

simulated hospital wastewater by UV254 and UV254/persulfate processes. Water Res. 2017, 122, 26

ACS Paragon Plus Environment

Page 26 of 35

Page 27 of 35

Environmental Science & Technology

569

128-138.

570

(38) Wang, Y.; Le Roux, J.; Zhang, T.; Croué, J. P. Formation of brominated disinfection

571

byproducts from natural organic matter isolates and model compounds in a sulfate

572

radical-based oxidation process. Environ. Sci. Technol. 2014, 48 (24), 14534-14542.

573

(39) Liang, C.; Huang, C. F.; Mohanty, N.; Kurakalva, R. M. A rapid spectrophotometric

574

determination of persulfate anion in ISCO. Chemosphere 2008, 73 (9), 1540-1543.

575

(40) Zhang, R.; Sun, P.; Boyer, T. H.; Zhao, L.; Huang, C. H. Degradation of pharmaceuticals

576

and metabolite in synthetic human urine by UV, UV/H2O2 and UV/PDS. Environ. Sci. Technol.

577

2015, 49 (5), 3056-3066.

578

(41) Furman, O.; Teel, A.; Ahmad, M.; Merker, M.; Watts, R. Effect of basicity on persulfate

579

reactivity. J. Environ. Eng. 2011, 137 (4), 241-247.

580

(42) Chen, J.; Zhang, L.; Huang, T.; Li, W.; Wang, Y.; Wang, Z. Decolorization of azo dye by

581

peroxymonosulfate activated by carbon nanotube: Radical versus non-radical mechanism. J.

582

Hazard. Mater. 2016, 320, 571-580.

583

(43) Zhang, T.; Zhu, H.; Croué, J. P. Production of sulfate radical from peroxymonosulfate

584

induced by a magnetically separable CuFe2O4 spinel in water: Efficiency, stabilityand

585

mechanism. Environ. Sci. Technol. 2013, 47 (6), 2784-2791.

586

(44) Chen, J.; Sun, P.; Zhang, Y.; Huang, C. H. Multiple roles of Cu(II) in catalyzing

587

hydrolysis and oxidation of β-lactam antibiotics. Environ. Sci. Technol. 2016, 50 (22),

588

12156-12165.

589

(45) Guan, Y. H.; Ma, J.; Ren, Y. M.; Liu, Y. L.; Xiao, J. Y.; Lin, L. Q.; Zhang, C. Efficient

590

degradation of atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate 27

ACS Paragon Plus Environment

Environmental Science & Technology

591

oxidation via the formation of hydroxyl and sulfate radicals. Water Res. 2013, 47 (14),

592

5431-5438.

593

(46) Dodd, M. C.; Buffle, M. O.; Von Gunten, U. Oxidation of antibacterial molecules by

594

aqueous ozone:  Moiety-specific reaction kinetics and application to ozone-based wastewater

595

treatment. Environ. Sci. Technol. 2006, 40 (6), 1969-1977.

596

(47) Rickman, K. A.; Mezyk, S. P. Kinetics and mechanisms of sulfate radical oxidation of

597

β-lactam antibiotics in water. Chemosphere 2010, 81 (3), 359-365.

598

(48) He, X.; Mezyk, S. P.; Michael, I.; Fatta-Kassinos, D.; Dionysiou, D. D. Degradation

599

kinetics and mechanism of β-lactam antibiotics by the activation of H2O2 and Na2S2O8 under

600

UV-254 nm irradiation. J. Hazard. Mater. 2014, 279 (0), 375-383.

601

(49) Edwards, J. O. Peroxide reaction mechanisms; Interscience Publishers: New York, 1962;

602

p 67-106.

603

(50) Betterton, E. A. Oxidation of alkyl sulfides by aqueous peroxymonosulfate. Environ. Sci.

604

Technol. 1992, 26 (3), 527-532.

605

(51) Page, M. I. The mechanisms of reactions of beta-lactam antibiotics. Accounts Chem. Res.

606

1984, 17 (4), 144-151.

607

(52) Pereira, J. H. O. S.; Reis, A. C.; Homem, V.; Silva, J. A.; Alves, A.; Borges, M. T.;

608

Boaventura, R. A. R.; Vilar, V. J. P.; Nunes, O. C. Solar photocatalytic oxidation of

609

recalcitrant natural metabolic by-products of amoxicillin biodegradation. Water Res. 2014, 65

610

(0), 307-320.

611

(53) Mitchell, S. M.; Ullman, J. L.; Teel, A. L.; Watts, R. J. pH and temperature effects on the

612

hydrolysis of three β-lactam antibiotics: Ampicillin, cefalotin and cefoxitin. Sci. Total Environ. 28

ACS Paragon Plus Environment

Page 28 of 35

Page 29 of 35

Environmental Science & Technology

613

2014, 466–467 (0), 547-555.

614

(54) Dodd, M. C.; Rentsch, D.; Singer, H. P.; Kohler, H. P. E.; Gunten, U. V. Transformation

615

of β-lactam antibacterial agents during aqueous ozonation: Reaction pathways and

616

quantitative bioassay of biologically-active oxidation products. Environ. Sci. Technol. 2010,

617

44 (15), 5940-5948.

618

(55) Li, L.; Wei, D.; Wei, G.; Du, Y. Transformation of cefazolin during chlorination process:

619

Products, mechanism and genotoxicity assessment. J. Hazard. Mater. 2013, 262 (0), 48-54.

620

(56) Trovo, A. G.; Nogueira, R. F. P.; Aguera, A.; Fernandez-Alba, A. R.; Malato, S.

621

Degradation of the antibiotic amoxicillin by photo-Fenton process - Chemical and

622

toxicological assessment. Water Res. 2011, 45 (3), 1394-1402.

623

(57) Dimitrakopoulou, D.; Rethemiotaki, I.; Frontistis, Z.; Xekoukoulotakis, N. P.; Venieri, D.;

624

Mantzavinos, D. Degradation, mineralization and antibiotic inactivation of amoxicillin by

625

UV-A/TiO2 photocatalysis. J. Environ. Manage. 2012, 98, 168-174.

29

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 35

Figure 1. Structures of β-lactam antibiotics and related compounds examined in this study. 6-APA:

6-aminopenicillanic

4-pyridineacetic

acid,

acid,

7-ACA:

3-MCA:

7-aminocephalosporanic

3-methylcrotonic

acid,

acid,

PTA:

ATMAA:

2-(2-Aminothiazole-4-yl)-2-methoxyiminoacetic acid. Note: 7-ACA and 6-APA are the core structure for penicillins and cephaloporins, respectively; PTA, ATMAA, 2-thiopheneacetic acid represent the substructure on the side chain of CFP, CFT and CFL, respectively; phenylglcine is the substructure of phenylglycine-type β-lactam antibiotics, including CFX, CFD, AMP, and AMX.

30

ACS Paragon Plus Environment

Page 31 of 35

Environmental Science & Technology

(A)

1.0

PMS PDS PMS + 400 µM Co(II) PDS + 40 µM Ag(I) PDS + 200 µM Ag(I) PDS + 400 µM Ag(I)

0.8

C/C0

0.6 0.4 0.2 0.0 0

30

60

90

120

150

180

t (min)

(B)

80

Experimental k2,app Modeled k2,app

60

kCFX -HSO .αCFX .βHSO 0

0

-1

_

5

-1

k2,app, M .s

_

5

40

kCFX -HSO .αCFX .βHSO _

_

_

5

20

_

5

kCFX -SO .αCFX .βSO _

2-

_

5

2-

5

0 5

6

7

8

9

10

pH Figure 2. Comparison of PS with the metal-activated PS process for CFX degradation (A), and effect of pH and kinetic modeling for the reaction rate constants of CFX with PMS (B). [CFX] = 40 µM, [PS] = 400 µM, 10 mM phosphate buffer used to control pH at 7.0 for PS or PMS/Co(II) system, but at pH 3.0 for PDS/Ag(I) system. Error bars indicate standard deviations of triplicates.

31

ACS Paragon Plus Environment

Environmental Science & Technology

Page 32 of 35

(A)

0 7-ACA 6-APA PTA phenylglycine 3-MCA ATMAA 2-thiopheneacetic acid

ln(C/C0)

-1 -2 -3 -4 -5 0

500

1000

1500 3500

3600

t (s)

cephalosporin

penicillin carbapenem substructure (B)

75

45

-1

-1

k2, app (M .s )

60

30

15

0 CFX CFD CFP CFL CFT

PG

AMX AMP MPN IPN 6-APA7-ACA PTA

Figure 3. Degradation of structurally related chemicals by PMS (A), and the apparent second-order rate constants for various β-lactam antibiotics (B). [β-lactam antibiotic] = 40 µM, [PMS] = 400 µM, pH 7.0 (10 mM phosphate buffer). Error bars indicate standard deviations of triplicates. Note: Thioether sulfur exists in all the investigated β-lactam antibiotics, and phenylglycine only presents in CFX, CFD, AMP and AMX.

32

ACS Paragon Plus Environment

Page 33 of 35

Environmental Science & Technology

Figure 4. Transformation products of β-lactam antibiotics by PMS oxidation. Note: R1 and R3 represent the side chain of β-lactam in pencillins and cephalosporins, and R2 is the side chain fused to six-membered dihydrothiazine ring of cephalosporins.

33

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 5. Proposed mechanism of PMS-induced oxidation of β-lactam antibiotics with cephalosporin as a representative. Nu: nucleophiles.

34

ACS Paragon Plus Environment

Page 34 of 35

Page 35 of 35

Environmental Science & Technology

TOC Art:

35

ACS Paragon Plus Environment