Rapid Aerobic Inactivation and Facile Removal of Escherichia coli with

Feb 28, 2019 - Efficient Photocatalytic Fuel Cell via Simultaneous Visible-Photoelectrocatalytic Degradation and Electricity Generation on a Porous Co...
0 downloads 0 Views 1MB Size
Subscriber access provided by WEBSTER UNIV

Remediation and Control Technologies

Rapid Aerobic Inactivation and Facile Removal of Escherichia coli with Amorphous Zero-valent Iron Microspheres: Indispensable Roles of Reactive Oxygen Species and Iron Corrosion Products Hongwei Sun, Jian Wang, Yao Jiang, Wenjuan Shen, Falong Jia, Shaohui Wang, Xiaomei Liao, and Lizhi Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06499 • Publication Date (Web): 28 Feb 2019 Downloaded from http://pubs.acs.org on February 28, 2019

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 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 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.

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

Rapid Aerobic Inactivation and Facile Removal of Escherichia

2

coli

3

Indispensable Roles of Reactive Oxygen Species and Iron

4

Corrosion Products

5

Hongwei Sun†, Jian Wang§, Yao Jiang§, Wenjuan Shen†, Falong Jia†,*, Shaohui Wang§, Xiaomei

6

Liao§,*, Lizhi Zhang†,*

7



8

Environmental & Applied Chemistry, College of Chemistry, Central China Normal University,

9

Wuhan 430079, P. R. China

with

Amorphous

Zero-valent

Iron

Microspheres:

Key Laboratory of Pesticide & Chemical Biology of Ministry of Education, Institute of

10

§

11

China Normal University, Wuhan 430079, P. R. China

Hubei Key Lab of Genetic Regulation and Integrative Biology, School of Life Sciences, Central

12 13

* To whom correspondence should be addressed.

14

Fax: +86 027 67867535; E-mail: [email protected]. (F. J.)

15

Fax: +86 027 67861936; E-mail: [email protected]. (X. L.)

16

Fax: +86 027 67867535; E-mail: [email protected]. (L. Z.)

17

1

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 35

18

ABSTRACT: Zero valent iron (ZVI) is recently regarded as a promising alternative for water

19

disinfection, but still suffers from low efficiency. Herein we demonstrate that amorphous zero valent

20

iron microspheres (A-mZVI) exhibit both higher inactivation rate and physical removal efficiency

21

for the disinfection of Escherichia coli (E. coli) than conventional crystalline nanoscale ZVI

22

(C-nZVI) under aerobic condition. The enhanced E. coli inactivation performance of A-mZVI was

23

mainly attributed to more reactive oxygen species (ROSs), especially free •OH, generated by the

24

accelerated iron dissolution and molecular oxygen activation in bulk solution. In contrast, C-nZVI

25

preferred to produce surface bound •OH, and its bactericidal ability was thus hampered by the

26

limited physical contact between C-nZVI and E. coli. More importantly, hydrolysis of dissolved iron

27

released from A-mZVI produced plenty of loose FeOOH to wrap E. coli, increasing the dysfunction

28

of E. coli membrane. Meanwhile, this hydrolysis process lowered the stability of E. coli colloid and

29

caused its rapid coagulation and sedimentation, favoring its physical removal. These findings clarify

30

the indispensable roles of ROSs and iron corrosion products during the ZVI disinfection, and also

31

provide a promising disinfection material for water treatment.

32 33

Keywords: Amorphous zero valent iron; Disinfection; E. coli; Reactive oxygen species; Physical

34

removal;

35 36

INTRODUCTION

37

Over the past decades, nanoscale zero-valent iron (nZVI) has been widely used to remove various

38

environmental pollutants, including heavy metals,1, 2 halogenated organics and nitroaromatics,3-5 etc,

39

owing to its high redox activity, facile synthesis and environmental benignancy.6-8 Recently, nZVI

40

was also reported to inactivate a broad spectrum of microbes such as virus, coliphage and bacteria, 2

ACS Paragon Plus Environment

Page 3 of 35

Environmental Science & Technology

41

making it a promising antimicrobial agent for water disinfection.9-11 In comparison with chlorine,

42

nZVI is attractive due to free of carcinogenic disinfection byproducts (DBPs) such as

43

trihalomethanes (THMs), haloacetic acids (HAAs), bromate and chlorite, which are listed as primary

44

contaminants in drinking water by the United States Environmental Protection Agency (USEPA),

45

European Union and World Health Organization (WHO). Meanwhile, nZVI disinfection is capable

46

of removing co-existing contaminants like heavy metals and organic pollutants simultaneously,

47

which is advantageous over chlorination especially in the scenario of wastewater disinfection, where

48

large amounts of pollutants are present. These pollutants might act as precursors of DBPs and must

49

be removed in advance in the case of chlorination, which would increase the operation cost.12,

50

More importantly, the concomitant physical removal of bacteria from water along with inactivation

51

through magnetic separation,14 was strikingly advantageous for practical disinfection by nZVI,

52

because this physical removal process could reduce the chemical oxygen demand (COD) of water by

53

removing the debris of inactivated microbes, and eliminate the risk from regrowth of inactivated

54

bacteria and potential spread of antibiotic resistance genes through horizontal gene transfer.15-18 In

55

addition, nZVI is easier for handling during transport, storage and deployment than chlorine,

56

therefore benefits its application in the disinfection of non-central water supply, and point-of-use

57

water purification.19

13

58

Despite its merits, the disinfection efficiency of nZVI was still inadequate for practical

59

application, especially under aerobic conditions. Therefore, relatively high doses were required to

60

achieve satisfactory disinfection effects, which might improve cost and cause iron sludge.20, 21 It was

61

reported that the aerobic inactivation of microorganism by nZVI mainly involved the generation of

62

reactive oxygen species (ROSs) via oxygenation of nZVI, and the direct physical disruption of cell

63

membrane components by adhered nZVI particles.9,

22

Normally, the oxygenation of nZVI would 3

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 35

64

produce surface oxide layers, which subsequently passivated nZVI and decelerated the further

65

outward electron transfer from Fe0 core to the oxide surface for activation of oxygen and production

66

of ROSs.7, 23 In addition, easy aggregation of nZVI particles, which was driven by electrostatic and

67

magnetic attractions, resulted in poor dispersibility of nZVI and less adhesion onto cells.24 Thus,

68

these two reasons accounted for the unsatisfactory disinfection reactivity of nZVI. Although doping

69

nZVI with precious metals such as Pt and Pd could avoid the unsatisfactory corrosion of iron core

70

and accelerate outward electron transfer through galvanic mechanism,25, 26 the high cost of precious

71

metals disfavored their field application. Surface coating with polymeric stabilizers such as starch

72

and cellulose was a common strategy to improve the dispersion and mobility of nZVI,27, 28 but was

73

always faced with a trade-off between the stability and reactivity of nZVI, and the decreased

74

bactericidal effect after polymer coating.29-32 Obviously, it is vital to develop new strategies to

75

improve the disinfection performance of nZVI.

76

In 2005, Lowry et al. investigated the influence of nZVI annealing treatment on its catalyzed

77

reductive trichloroethylene (TCE) transformation with H2. They found that poorly ordered pristine

78

nZVI could activate H2 to reduce TCE to ethane, but annealed nZVI of good crystallinity could not.

79

Meanwhile, the oxidative dissolution of nZVI was also retarded after annealing treatment.33 Later,

80

Nurmi and his coworkers revealed that nZVI, being prepared by borohydride reduction, was poorly

81

ordered, but outperformed its crystalline counterpart prepared through thermal reduction of goethite

82

with H2, on reducing pollutants like benzoquinone and CCl4.34 Moreover, Wang et al. demonstrated

83

that the recrystallization of aged nZVI was responsible for its activity loss of bromate reduction.35

84

On the basis of these results, we proposed that the disordered structure of nZVI might benefit the

85

electron donating capacity of iron core, and for the first time prepared amorphous zero valent iron

86

microspheres (A-mZVI) through Fe(II) reduction with borohydride in the presence of 4

ACS Paragon Plus Environment

Page 5 of 35

Environmental Science & Technology

87

ethylenediamine (EDA). However, it is still unknown how the amorphous structure will affect the

88

ROSs generation and disinfection performances of ZVI, which is crucial for us to clarify the

89

pollutant removal ability of A-mZVI.

90

In this study, we compared the inactivation and physical removal performances of A-mZVI and

91

zero-valent iron nanoparticles (C-nZVI) toward a typical Gram-negative bacterium Escherichia coli

92

(E. coli) under aerobic condition, aiming to unravel the aerobic disinfection mechanism of A-mZVI

93

through investigating the production and evolution of ROSs in the A-mZVI/air and C-nZVI/air

94

systems, and clarified the effects of dissolution processes and corrosion products of A-mZVI and

95

C-nZVI on the inactivation and physical removal of E. coli.

96 97

EXPERIMENTAL SECTION

98

Synthesis and Characterization of ZVI. To prepare A-mZVI, 75 mL of EDA (0.9 mol/L) was

99

added into 300 mL of 0.075 mol/L FeCl2 aqueous solution, after thoroughly mixed for 3 min under

100

stirring, 75 mL of 1.6 mol/L NaBH4 aqueous solution was introduced. The reduction process was

101

finished in 25 min until no obvious bubble emission was observed. The obtained black precipitate

102

was collected, washed with deionized water and absolute ethanol for 3 times, and dried under

103

infrared lamp. All these procedures were conducted under argon gas protection to avoid oxidation of

104

iron. C-nZVI was prepared with similar procedures without adding EDA. The as-prepared ZVI

105

samples were characterized with scanning electron microscopy (SEM, TESCAN MIRA 3, Czech)

106

and X-ray diffraction (XRD, D8 Advance, Cu Kα radiation, λ = 0.15418 nm, Bruker, Germany). The

107

content of nitrogen in the synthesized A-mZVI and C-nZVI was measured by elementary analysis

108

(vario Micro cube, Elementar) to check the residue of EDA, and the release of EDA from A-mZVI

109

during use was evaluated by measuring total nitrogen (TN) concentrations with a TOC/TNb 5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 35

110

Analyzer (vario TOC select, Elementar). Fe0 content was quantified by acid digestion and hydrogen

111

gas production assay.33 Specific surface area (SSA) was determined by Brunauer-Emmett-Teller

112

(BET) method with a N2 adsorption system (Micromeritics Tristar 3000). Details of characterization

113

methods are provided in Supporting Information (SI).

114

Disinfection Experiments. E. coli strain BW25113 was purchased from Coli Genetic Stock Center

115

of Yale University, USA. To prepare the bacterial suspension, E. coli was inoculated into 50 mL of

116

autoclaved nutrient broth and incubated overnight, then the bacteria were harvested and washed. The

117

pellet was resuspended in 0.85% NaCl and the optical density at 670 nm (OD670 nm) was adjusted to

118

0.06. ZVI at predetermined doses were added into 50 mL of the obtained bacterial suspension and

119

incubated at 250 rpm and 37 ℃ in an orbital shaker. To determine the survival ratio and physical

120

removal efficiency of E. coli, 5 mL of the slurry samples were collected at regular intervals and

121

magnetically separated to obtain the supernatant, which was subsequently stained with

122

LIVE/DEADTM BacLightTM Bacterial Viability Kit and analyzed with flow cytometry

123

(LSRFortessaTM, BD, USA). This Kit contains two fluorescent nucleic acids dyes, namely SYTOTM

124

9 and propidium iodide (PI). SYTOTM 9 is cell membrane permeable and can bind with bacterial

125

DNA to emit green fluorescence (530 nm) when excited at 488 nm. PI can quench SYTOTM 9 due to

126

its higher DNA binding affinity to generate red fluorescence (excitation wavelength of 488 nm and

127

emission wavelength of 630 nm). Nevertheless, PI is not permeable to intact membrane, and thus the

128

live cells with intact membrane can only be stained green by SYTOTM 9, and dead cells with

129

disturbed membrane can be stained red by PI. The numbers of fluorescent particles (E. coli cells) in

130

200 μL samples were recorded by volumetric counting hardware. In the two-parameter dot plots of

131

PI versus SYTOTM 9, four clusters could be distinguished by quadrants gate (Q1-4), which were

6

ACS Paragon Plus Environment

Page 7 of 35

Environmental Science & Technology

132

ascribed to dead, injured, live cells and negative background signals (SI Figure S1), respectively.

133

The survival ratio and removal efficiency of E. coli were calculated according to Equation 1-3:

134

Survival ratio = N3/(N1+N2+N3)

(1)

135

C = (N1+N2+N3) / 200 μL

(2)

136

Removal efficiency (%) = (C0-Ct)/C0

(3)

137

where N1-3 were the cell numbers in Q1-3, respectively, and (N1+N2+N3) referred to the total cell

138

numbers. C0 and Ct were the bacterial concentrations in the supernatant at time 0 and t, respectively.

139

E. coli inactivation efficiency was also monitored by the heterotrophic plate count (HPC) method at

140

a predetermined initial E. coli concentration of 1×106 colony forming unit per milliliter (CFU/mL) to

141

facilitate the comparison of A-mZVI with other ZVI in the literature. Details of the method were

142

provided in SI.

143

The survival states of the cells in sediment samples were determined by a fluorescent

144

microscope (EVOS FL Auto, Life Technologies Corporation) after staining with LIVE/DEADTM

145

Kit. Images of green (SYTOTM 9) and red (PI) fluorescent channels were captured simultaneously

146

for the same sample, and their merged images were used to depict the viable states of E. coli. The

147

intensities of green and red fluorescence were also integrated with ImageJ,36 and their ratio

148

(Green/Red) was used as an index of the E. coli viability in the sediments. Note that the Green/Red

149

ratio of E. coli cells at 0 min could not be calculated without red fluorescence signals. The ratio

150

could only be obtained after E. coli cells were inactivated by ZVI and the bacterial membrane

151

became disturbed and permeable to PI. Details of the method were described in the SI.

152

Analytical Methods. The mixtures after ZVI disinfection were left to settle for different durations,

153

and the turbidity of the supernatant was measured with a portable turbidimeter (2100Q, HACH).

154

Concentrations of dissolved Fe2+ and Fe3+ were determined by the 1,10-phenanthroline colorimetric 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 35

155

method.37

156

5,5-dimethyl-1-pyrroline N-oxide (DMPO) were employed to detect •OH and •O2- with a JES FA

157

200 X-band spectrometer (JEOL, Japan). H2O2 concentration was tested with a modified

158

p-hydroxyphenylacetic acid (POHPAA) fluorescent method.38 Intracellular ROSs concentration was

159

measured fluorescently by utilizing 2,7-Dichlorodihydrofluorescein diacetate (DCFH-DA) as a

160

probe.39 Catalase and superoxide dismutase (SOD) activities were measured using Catalase Assay

161

Kit (S0051, Beyotime Institute of Biotechnology, China) and Superoxide Dismutase Assay Kit

162

(S0101, Beyotime Institute of Biotechnology), respectively. Membrane potential of E. coli was

163

measured with Rhodamine 123 (Rh123) probing method,22 and adenosine triphosphate (ATP)

164

production by E. coli was assayed by a BacTiter-GloTM Microbial Cell Viability Assay Kit (G8230,

165

Promega Corporation, USA). The sample preparation protocols of E. coli for SEM and transmission

166

electron microscopy (TEM, JEOL-2010, 200 kV) characterization were provided in SI. Magnetic

167

hysteresis loops of the oxidized ZVI were measured with a vibrating sample magnetometer (VSM) at

168

ambient temperature within the magnetic field of ± 10,000 Oe. Particle size distribution of ZVI-E.

169

coli suspension was measured by dynamic light scattering (Zetasizer Nano, Malvern), as described

170

in SI.

Electron

spin

resonance

(ESR)

spectra

and

the

radical

spin

trapper

171 172

RESULTS AND DISCUSSION

173

Synthesis and Characterization of ZVI. As shown in the SEM images, the as-synthesized C-nZVI

174

contained chains or aggregates assembled by spheres of 50-200 nm (Figure 1a), consistent with our

175

previous study.6 Differently, A-mZVI was composed of spheres in size of 0.8-1.2 μm (Figure 1b and

176

1c). XRD pattern suggested C-nZVI exhibited an obvious diffraction peak around 2θ value of 44.5°,

177

corresponding to α-Fe0 (JCPDS No. 87-722), which was the typical XRD pattern of nZVI produced 8

ACS Paragon Plus Environment

Page 9 of 35

Environmental Science & Technology

178

by NaBH4 reduction in solution.40,

179

A-mZVI, indicating the formation of amorphous structure of ZVI in the presence of EDA (Figure

180

1d). This amorphous nature of A-mZVI might benefit the production of ROSs and aerobic

181

disinfection of E. coli. The contents of N, C, H in C-nZVI were 0%, 0.47% and 0.48%, which were

182

0.90%, 1.16% and 0.55% in A-mZVI, respectively (SI Table S1). This comparison suggested the

183

EDA content in the as-prepared A-mZVI was rather low, as the washing with deionized water and

184

ethanol might remove most of EDA residue. Subsequently, the release of EDA from A-mZVI during

185

disinfection was evaluated. A-mZVI were subjected to aerobic corrosion for 1 h and then filtered.

186

The TN concentrations of filtrates were found to be 1.5, 1.8 and 3.2 mg/L for A-mZVI dosage of

187

200, 500 and 1000 mg/L, respectively (SI Figure S2). These values were much lower than the US

188

national criterions of nitrate nitrogen (N) and total ammonia nitrogen (TAN), which were 10 mg/L

189

for domestic water supply and 17 mg/L for fresh water, respectively.42, 43 TN concentrations of 1.5,

190

1.8 and 3.2 mg/L corresponded to EDA concentrations of 3.2, 3.9 and 6.9 mg/L, respectively,

191

assuming that all the nitrogen elements were originated from EDA. Since there is no guideline

192

concentration of EDA in water available, the lowest No Observed Adverse Effect Level (NOAEL, 9

193

mg/kg body weight per day) obtained from rat experiments was used to evaluate the toxicity and

194

overall risk of EDA oral intake, as recommended by WHO.44 The estimated daily intake (EDI) of

195

EDA through water consumption was calculated according to equation 4:45

196

41

However, no obvious diffraction peak was acquired for

EDI = Ci × CR / BW

(4)

197

where Ci is the concentration of EDA in water, CR represents the consumption rate of drinking

198

water (2 L/day), and BW stands for the average body weight (70 kg for adults). The EDIs were 0.09,

199

0.1 and 0.2 mg/kg body weight per day at A-mZVI dose of 200, 500 and 1000 mg/L, respectively,

200

which were far less than the NOAEL of 9 mg/kg body weight per day recommended by WHO. 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 35

201

Therefore, the toxicity and risk induced by the EDA residual in water were negligible.

202

Inactivation and Physical Removal of E. coli by ZVI. The as-prepared ZVI samples were then

203

used for the disinfection of E. coli under aerobic condition. The initial concentration of E. coli was

204

determined to be approximately 107 cells/mL by using flow cytometry (Figure 2a). In the

205

supernatant samples after C-nZVI disinfection, both the total and viable E. coli concentrations

206

decreased steadily, which were 8.4×105 and 5×105 cells/mL after 60 min, respectively. Impressively,

207

A-mZVI induced much sharper concentration decrease of total and live cells in the supernatant,

208

which were respectively down to 5×105 and 2×105 cells/mL after only 5 min (Figure 2a). The

209

survival ratios and removal efficiencies of E. coli were calculated according to Eqs 1-3. We found

210

that A-mZVI inactivated E. coli much more rapidly, and only 18% of the cells remained viable after

211

60 min, much lower than that (53%) of C-nZVI treatment (Figure 2b). More importantly, A-mZVI

212

also showed remarkably better physical removal performance. The removal efficiency of E. coli in

213

the C-nZVI system increased gradually to 95% after 1 h, but it just took 5 min for A-mZVI to reach

214

a comparable removal efficiency (Figure 2c).

215

Since the physical removal effect of ZVI would precipitate most of E. coli cells into the

216

sediment, it is essential to understand the fate of E. coli in this phase. Nonetheless, the survival status

217

of bacteria physically removed by ZVI has never been studied in previous reports.14, 46 Herein, we

218

utilized fluorescent microscope to check the viability of E. coli in the sediment samples, and found

219

that these cells were inactivated gradually, following similar trends with those in the supernatant (SI

220

Figure S3). As an index of E. coli viability, the fluorescent intensity ratio of Green/Red decreased

221

steadily from 4.82 at 5 min to 0.67 at 60 min in the C-nZVI/E. coli system, whereas the ratio was

222

1.24 after 5 min and further reduced to 0.46 at 40 min in the A-mZVI/E. coli system (Figure 2d).

223

This comparison confirmed that A-mZVI could also more efficiently inactivate E. coli in the 10

ACS Paragon Plus Environment

Page 11 of 35

224

Environmental Science & Technology

sediments.

225

The influence of ZVI dose on the inactivation and removal of E. coli was also investigated. As

226

expected, the inactivation of E. coli showed an obvious dose-dependent manner in the range of

227

20-400 mg/L, but A-mZVI exhibited apparent advantage over C-nZVI at identical dosage. For

228

example, virtually complete inactivation of E. coli was achieved at 400 mg/L of A-mZVI, whilst

229

16% of E. coli still survived at the same dose of C-nZVI (Figure 3a). Similar trends were observed in

230

the sediment samples via fluorescent microscopic images (SI Figure S4). The minimum doses of

231

C-nZVI and A-mZVI to physically remove 90% of E. coli from water were 50 and 20 mg/L,

232

respectively. This difference further confirmed the superiority of A-mZVI for physical removal of E.

233

coli (Figure 3b). Moreover, we conducted additional control experiments with crystalline

234

micro-scale zero valent iron (C-mZVI), which was prepared by annealing the A-mZVI at 500 °C for

235

4 h in Ar. SEM images indicated that the size of C-mZVI was in the range of 1-2 μm (SI Figure

236

S5a), similar with that of A-mZVI. As expected, the crystallinity of C-mZVI greatly improved after

237

annealing treatment, as revealed by the sharp diffraction peak at 2θ = 45° (JCPDS No. 87-722) in the

238

XRD pattern (SI Figure S5b). We thus compared the E. coli inactivation performance of A-mZVI,

239

C-nZVI and C-mZVI, and found that the disinfection capacity decreased following the order of

240

A-mZVI > C-nZVI > C-mZVI (SI Figure S5c). As both C-nZVI and C-mZVI were crystallized, the

241

lower disinfection activity of C-mZVI with larger particle size was consistent with previous reports

242

about the size effect on bactericidal capacity of ZVI.47 Although A-mZVI and C-mZVI had similar

243

particle size, C-mZVI of crystalline nature possessed a dramatically reduced antibacterial activity.

244

Therefore, we concluded that the improved disinfection performance of A-mZVI was arisen from its

245

amorphous structure, rather than the size effect. Moreover, the Fe0 content of A-mZVI (67%) was

246

only slightly higher than that (60%) of C-nZVI (SI Table S1), which further confirmed that the 11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 35

247

improved E. coli inactivation of A-mZVI was not related to the Fe0 content, but to the amorphous

248

structure.

249

The E. coli inactivation efficiency of A-mZVI was also compared with previous reports. Since

250

most of the literatures determined the inactivation efficiencies of E. coli by using the HPC method

251

with initial E. coli concentration of 1×106 CFU/mL, we conducted similar experiments by using

252

A-mZVI for better comparison. The inactivation efficiency (log (N0/N)) was 2.4 ± 0.2 after treated

253

with 100 mg/L of A-mZVI for 60 min under aerobic condition. After settling for 0.5 h, the E. coli

254

concentrations in the supernatant decreased by 3.8 ± 0.04 log (SI Table S2). This inactivation

255

efficiency of A-mZVI was higher than those of many ZVI reported in the literature, including nZVI,

256

electrosprayed nZVI, commercial Nanofer 25 (Nano Iron s.r.o., Czech Republic), commercial

257

reactive nanoscale iron particles (RNIP, Toda Kogyo Corp.), and microscale iron powder.14, 21, 29, 47, 48

258

The cost of the A-mZVI synthesized was estimated to be ca. 0.6 USD/g (SI Table S3), which was

259

relatively high because the synthesis of A-mZVI was conducted in laboratory scale and the reagents

260

used were supplied in small packages and of high purity (AR). Nevertheless, the price of A-mZVI

261

could be significantly decreased in industrial production by scaling-up the fabrication and by

262

adopting much cheaper reagents of technical grade to meet the need of practical disinfection.

263

Subsequently, the turbidity of the ZVI treated water was also monitored, which were 158.0 and

264

88.6 nephelometric turbidity unit (NTU) for A-mZVI and C-nZVI right after the batch experiments.

265

The turbidity of the supernatant in the A-mZVI system decreased sharply to 19.2 NTU after only 0.5

266

h of settling, much faster than that (45.4 NTU) in the C-nZVI system. The turbidity further

267

decreased with prolonged settling duration (SI Figure S6a). Interestingly, a linear relationship was

268

revealed between Ln (turbidityt/turbidity0) and settling time, where turbidityt and turbidity0 were the

269

turbidity at settling time t and 0 h, respectively (SI Figure S6b). Therefore, the turbidity of effluent 12

ACS Paragon Plus Environment

Page 13 of 35

Environmental Science & Technology

270

after A-mZVI disinfection can be controlled by tuning the settling duration according to the purpose

271

of application. In the case of wastewater disinfection, the wastewater discharge standard of China

272

(GB 8978-1996) regulates the total suspended solids (TSS) from 20 to 100 mg/L in different

273

scenarios, which corresponds to turbidity values of ca.13-100 NTU according to a general rule of

274

thumb. This suggests that 1 mg/L of TSS approximately equals to 1.0-1.5 NTU of turbidity.49 Then

275

less than 1.3 h of settling after A-mZVI disinfection was required to meet the wastewater discharge

276

standard (13 NTU, SI Figure S6b). In the context of drinking water disinfection, the criteria of

277

turbidity for central water supply and non-central water supply were 1 and 3 NTU, which could be

278

raised to 3 and 5 NTU when the quality of water source or the technical conditions of water

279

purification were limited, according to the drinking water standard of China (GB5749-2006). Then

280

the necessary settling durations were 6.8, 9.7 and 16 h to meet the turbidity standards of 5, 3 and 1

281

NTU, respectively (SI Figure S6b). Therefore, A-mZVI is potentially used for water disinfection

282

with post-disinfection settling process. Alternatively, settling process might be omitted if

283

disinfection be carried out in a flow-through column packed with ZVI.

284

Mechanism of Inactivation. To unravel the mechanism of promoted E. coli inactivation

285

performance of A-mZVI, we first investigated the contribution of various ROSs by scavenging

286

experiments, since ROSs played vital roles in the aerobic inactivation of bacteria by ZVI.9,

287

Intriguingly, the inactivation performance of the two ZVI samples changed in significantly different

288

manners after ROSs quenching. In the C-nZVI/E. coli system, the scavenging effects descended

289

following the order of SOD > catalase > tert-butyl alcohol (TBA), which were used to quench •O2-,

290

H2O2 and •OH, respectively. This trend suggested that •O2- and H2O2 were responsible for the

291

inactivation of E. coli, whereas •OH was less important, which was consistent with previous

292

reports.9, 47 As for the A-mZVI/E. coli system, the addition of SOD, catalase and TBA resulted in

50

13

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 35

293

similar changes of survival ratio, highlighting the indispensable role of •OH for the E. coli

294

disinfection (Figure 3c and SI  Figure S7), and also explaining the better bactericidal performance

295

of A-mZVI, because •OH is a much stronger oxidant than •O2- and H2O2.

296 297

Subsequently, we attempted to clarify how the amorphous structure altered ROSs production of ZVI. Generally, the evolution of ROSs by ZVI oxygenation could be described by Eqs 5-11:8, 51, 52

298

Fe0 → ≡Fe2+ + 2e-

(5)

299

≡Fe2+ → Fe2+

(6)

300

O2 + 2e- + 2H+ → H2O2

(7)

301

O2 + e- → •O2-

(8)

302

•O2- + e- + 2H+ → H2O2

(9)

303

≡Fe2+ + H2O2 → ≡Fe3+ + •OHsurface + OH-

(10)

304

Fe2+ + H2O2 → Fe3+ + •OH + OH-

(11)

305

where ≡ indicated surface bound species. As stated above, •OH more contributed to the disinfection

306

in the A-mZVI/E. coli system than in the C-nZVI/E. coli system, whereas •O2- and H2O2 served

307

mainly as the precursors of •OH in the A-mZVI/E. coli system. To clarify the ROSs production

308

processes, we first investigated the dissolution behavior of the two ZVI samples. The dissolution rate

309

of A-mZVI was higher, indicating that A-mZVI possessed stronger electron donating capacity than

310

C-nZVI (Figure 4a). Obviously, this stronger electron donating capacity should facilitate the ROSs

311

production via reduction of O2 by A-mZVI (Eqs 6, 7), which was confirmed by more •O2- generated

312

in the A-mZVI system (SI Figure S8 and Figure 4b). Moreover, more accessible electrons from iron

313

core and dissolved Fe2+ in the A-mZVI system would further accelerate the transformation of •O2-

314

into H2O2 and finally •OH (Eqs 9-11), and thus shorten the lifetime of •O2- and H2O2. Consequently,

315

•O2- and H2O2 would act as intermediates of •OH rather than direct bactericidal agents in the 14

ACS Paragon Plus Environment

Page 15 of 35

Environmental Science & Technology

316

A-mZVI system. To validate this hypothesis, we measured the concentrations of H2O2 and •OH.

317

Although the steady-state H2O2 were not detected due to its rapid in-situ decomposition, the

318

accumulated H2O2 concentrations produced by A-mZVI were higher than C-nZVI (Figure 4c),

319

confirming the enhanced transformation of •O2- into H2O2 by A-mZVI. A-mZVI also showed much

320

faster H2O2 depletion performance (Figure 4d), and the H2O2 decomposition rate constant of

321

A-mZVI was 0.708 min-1, much higher than that (0.109 min-1) of C-nZVI (SI Figure S9). Finally, the

322

•OH levels, as reflected by the ESR intensities of DMPO-OH adduct, were significantly elevated in

323

the A-mZVI system (Figure 4f). These results proved the accelerated transformation of H2O2 into

324

•OH by A-mZVI. The shortened lifetime of •O2- and H2O2 was further supported by the lowered

325

intracellular ROSs levels induced by A-mZVI (Figure 4e), because the limited lifespan rendered it

326

less likely for •O2- and H2O2 to penetrate the cytomembrane into cytoplasm before their

327

transformation, whereas their derivative •OH was also too active and short-lived to enter cells.53

328

As for the C-nZVI system, most of Fe2+ generated by oxidation of Fe0 would be

329

surface-associated rather than dissolved,6, 51 resulting in the decomposition of H2O2 dominantly at

330

the surface of C-nZVI to produce surface-bound •OH (•OHsurface, Eq. 10).54 In such a case, the

331

interaction between •OHsurface and E. coli would largely depend on the physical contact between

332

C-nZVI and cells. However, the poor dispersion of C-nZVI in the disinfection system disfavored the

333

direct contact between C-nZVI and E. coli and thus impeded the attack of •OHsurface to the cells.

334

Whereas in the case of A-mZVI, more Fe2+ was dissolved, resulting in more decomposition of H2O2

335

in the bulk solution to produce free •OH (Eq. 11), which could kill E. coli without the direct contact

336

between A-mZVI and cells. To verify the proposed mechanism, the fractions of •OHsurface among the

337

overall •OH (sum of •OHsurface and free •OH) were further investigated by using DMPO-trapped EPR

338

technique. F- was used to desorb the •OHsurface by forming strong hydrogen bond,55 thus the EPR 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 35

339

intensities without F- (I0) represented free •OH levels, and the intensities with F- addition (IF) was

340

attributed to overall •OH levels, assuming that F- could disassociate all the •OHsurface. We defined the

341

•OHsurface fractions as (IF-I0)/IF, and found the •OHsurface fractions (78.7-92.5%) of C-nZVI system

342

was much higher than those (36.1-57.9%) of A-mZVI system (Figure 4f). These results confirmed

343

that C-nZVI was dominated by •OHsurface whereas A-mZVI favored free •OH.

344

The intracellular ROSs did not play important role in E. coli inactivation in the A-mZVI-E. coli

345

system, because most of ROSs were generated and transformed in the extracellular bulk solution to

346

•OH, the predominant bactericidal ROSs. In the scavenging experiments, the scavengers

347

impermeable to the membrane were supposed to scavenge the extracellular ROSs. Therefore, the

348

significant inhibition of disinfection by TBA suggested the indispensable role of extracellular •OH

349

(Figure 3c). Although A-mZVI could inactivate E. coli much faster than C-nZVI, the intracellular

350

ROSs levels in the A-mZVI-E. coli system was much lower than those in the C-nZVI-E. coli system

351

(Figure 4e). In contrast, intracellular ROSs might play relatively more important role in the

352

C-nZVI-E. coli system because of its much higher intracellular ROSs concentrations. The

353

scavenging experiments also indicated the vital roles of extracellular •O2- and H2O2 for E. coli

354

inactivation, but neither •O2- nor H2O2 were efficient for the direct oxidative damage of E. coli cells.

355

Comparatively, •OH is a much stronger oxidant, but the contribution of extracellular •OH was

356

negligible in the C-nZVI-E. coli system because most of •OH were produced in the form of surface

357

associated species and had limited physical contact with E. coli cells. Therefore, it is highly possible

358

that •O2- and H2O2 entered E. coli cells and thus became intracellular •OH to inactivate E. coli, as

359

revealed by the high intracellular ROSs levels induced by C-nZVI.56, 57 Nevertheless, it is still a great

360

challenge to elucidate the role of intracellular ROSs because of the limited techniques to quantify or

361

quench intracellular ROSs. 16

ACS Paragon Plus Environment

Page 17 of 35

Environmental Science & Technology

362

The responses of E. coli to the oxidative stress were then checked to further reveal the

363

promotion of ROSs mediated disinfection by using A-mZVI. Increasing concentrations of H2O2 and

364

•O2- would respectively trigger the biosynthesis of SOD and catalase, which in turn quench

365

excessive •O2- and H2O2 to keep their concentrations at acceptable levels.56,

366

elevation of SOD and catalase activities are supposed to indicate more ROSs production. As

367

expected, the activities of both SOD and catalase increased more quickly when E. coli were treated

368

with A-mZVI, in agreement with its greater ROSs generating capacity. Moreover, the

369

overwhelmingly yielded ROSs by A-mZVI, especially the •OH which cannot be quenched

370

effectively by catalase or SOD, would then cause the oxidative damage of the antioxidative

371

enzymes,39, 56 resulting in the sharp decrease of enzymatic activities after only 10 min. In contrast,

372

the activities of both enzymes kept increasing until 40 min and decreased more slowly in the

373

C-nZVI/E. coli system (Figure 5a and 5b). These results further confirmed the indispensable role of

374

ROSs on the enhanced E. coli inactivation of A-mZVI.

58

Therefore, the

375

Besides the oxidative inactivation induced by ROSs, the physical disruption of cell membrane

376

by adhered ZVI particles might also contribute to the disinfection of E. coli,9, 22, 59, 60 since quenching

377

ROSs could not inhibit the inactivation of E. coli completely (Figure 3c). Therefore, we checked the

378

physical contact between ZVI and E. coli by using SEM and TEM. In the C-nZVI/E. coli system,

379

quasi-spherical nanoparticles were found to adhere to the outer surface of E. coli loosely after 1h,

380

similar with previous reports.46,

381

numerous interconnected flakes to wrap the E. coli cells more tightly (Figure 6). The different

382

shapes and adsorption of reacted A-mZVI suggested the corrosion process of ZVI was greatly

383

changed by the amorphous structure. It is well documented that the corrosion of C-nZVI took place

384

mainly at its outer surface, featured by slow leaching of Fe2+ and the onsite growth of oxide shells.61

47

Interestingly, A-mZVI formed a thick envelope assembled by

17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 35

385

As a result, the adsorption of pristine or oxidized C-nZVI on the bacterial surface was hampered

386

owing to the poor dispersibility of C-nZVI. Differently, the fast dissolution of A-mZVI resulted in

387

more Fe2+ release, which was then hydrolyzed and oxidized to iron (hydro)oxide in the bulk

388

solution, thus the corrosion products of A-mZVI had easier access for the adsorption of E. coli. This

389

process was supported by the clear spatial separation of the spherical A-mZVI (denoted as

390

A-mZVI-2) from the laminar corrosion products (denoted as A-mZVI-1) in the SEM images (SI

391

Figure S10). More importantly, the iron (hydro)oxides species after corrosion were also altered by

392

the amorphous structure of A-mZVI. The XRD analysis showed that C-nZVI was converted to a

393

mixture of magnetite (Fe3O4, JCPDS No. 85-1436), akaganéite (β-FeOOH, JCPDS No. 75-1594) and

394

lepidocrocite (γ-FeOOH, JCPDS No. 70-0714), with magnetite as the dominant species, whereas

395

A-mZVI was mainly oxidized to lepidocrocite (Figure 6g), whose structure was more loosely than

396

that of magnetite and thus favored the sorption of E. coli cells. Meanwhile, it was reported that the

397

hydrolysis and oxidation of dissolved Fe2+ were able to produce polymeric Fe(OH)2 and Fe(OH)3

398

flocs as the precursors of lepidocrocite,62 and these flocs were also voluminous and favored the wrap

399

of E. coli cells.63 According to energy dispersive spectroscopy (EDS) analysis, the corrosion

400

products of A-mZVI (A-mZVI-1) possessed significantly higher carbon atomic percentage (36.41%)

401

than those (27.61% and 26.80%) of the remained spherical A-mZVI (A-mZVI-2) and used C-nZVI,

402

further confirming the higher E. coli adsorption capacity of the loose corrosion products of A-mZVI

403

(SI Figure S11).

404

The enhanced adsorption of A-mZVI corrosion products should cause more severe membrane

405

dysfunction of E. coli cells by physical disruption. Normally, the membrane potential and adenosine

406

triphosphate (ATP) synthesis capacity of cells were used as indicators of membrane dysfunction.

407

Membrane potential is crucial for the selectivity of plasma membrane,64 and the closely adhered ZVI 18

ACS Paragon Plus Environment

Page 19 of 35

Environmental Science & Technology

408

might block the ion channels and prevent the efflux of charged ions to construct the transmembrane

409

potential.65 After E. coli was treated with C-nZVI, the membrane potential maintained constant

410

during the initial 10 min, and dropped thereafter, but A-mZVI induced an instant membrane

411

potential decline as soon as E. coli was in contact with A-mZVI (Figure 5c). Moreover, the adsorbed

412

particles might isolate E. coli cells from their environment and prevent their consumption of

413

extracellular nutrients to synthesize ATP.59, 66 We collected the E. coli cells after different duration

414

of ZVI disinfection and measured their ATP synthesis rate (SI Figure S12), and found 5 min of

415

A-mZVI treatment drastically declined the rate from 4.75 to 0.87 nM/min, much faster than the case

416

of C-nZVI (Figure 5d). We thus concluded that the enhanced adhesion of corrosion products on E.

417

coli outer surface, originated from the amorphous structure of A-mZVI, also contributed to its better

418

disinfection performance.

419

It was known that the physical bacteria removal of nZVI relied on the adsorption of magnetic

420

iron oxide over bacterial surface and magnetic separation.14 However, the major corrosion product of

421

A-mZVI was lepidocrocite, whose saturation magnetization value was two orders of magnitude

422

lower than that (80-100 Am2/kg) of magnetite.62 As expected, the corrosion products of A-mZVI

423

were virtually non-magnetic (Figure 6h), and the sedimentation of the A-mZVI/E. coli mixture after

424

1 h of aerobic reaction was not affected by a magnet, different from the case of C-nZVI/E. coli

425

system (Figure 6i). These phenomena suggested that A-mZVI altered the E. coli physical removal

426

process thoroughly. Intrinsically, the suspension of E. coli is a kind of active colloid, and the rotation

427

of helical flagella endows E. coli with swimming motility, which is essential to maintain the cells

428

suspended against gravitational sedimentation.67 Therefore, the loss of motility would result in fast

429

precipitation and easy physical removal of E. coli. Although the inactivation of cells might decrease

430

motility, this reason could be ruled out because live cells were removed simultaneously with dead 19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 35

431

cells by ZVI and ROSs scavengers had little impact on the removal efficiencies (Figure 3d). As

432

stated above, A-mZVI could form thick envelope to wrap the E. coli cells tightly. This might inhibit

433

the rotation of helical flagella, thus decrease the mobility of E. coli cells and the stability of the E.

434

coli colloid, and subsequently cause the aggregation and gravitational sedimentation of bacterial

435

cells. To support this hypothesis, the particle size distribution of the ZVI/E. coli mixture was

436

measured. C-nZVI treatment induced a slow increase of particle size from ca. 750 nm (bare E. coli

437

cells) at 0 min to ca. 1800 nm (nZVI bearing E. coli cells and their aggregates) at 60 min, indicating

438

the slow adsorption of C-nZVI onto E. coli. In contrast, A-mZVI caused a drastic increase of particle

439

size from ca. 750 nm to ca. 5300 nm within 5 min, confirming the effective generation of

440

voluminous corrosion products and their fast coating of cells. After 5 min, the particle size decreased

441

steadily, corresponding to the precipitation of large E. coli-iron coagulate (SI Figure S13). The

442

possible physical removal processes of two systems are illustrated in Scheme 1.

443 444

Environmental Implications. ZVI is a promising antimicrobial agent because of its low cost, high

445

activity and low toxicity. However, the antimicrobial activity of ZVI is often hampered greatly by

446

the presence of oxygen due to the growth of passivation layer and the subsequent decreased outward

447

electron transfer from iron core, which limit its use. In this study, we have revealed that

448

amorphorization is an effective strategy to improve the bactericidal ability of ZVI under aerobic

449

conditions by enhancing the electron donating capacity of iron core to produce more ROSs,

450

especially free •OH in bulk solution. More importantly, the production of abundant loose corrosion

451

products by oxygenation of A-mZVI facilitates the simultaneous inactivation and physical removal

452

of E. coli, very promising for water self-cleanup during disinfection. Since the bacteria can be

453

immobilized by the iron hydroxides and easily separated from the treated water, the release of 20

ACS Paragon Plus Environment

Page 21 of 35

Environmental Science & Technology

454

antibiotic resistance genes (ARGs) from these bacteria can be suppressed. With further

455

detoxification processes of the bacteria-containing iron sludge, the spread of ARGs in water systems

456

might be eliminated. Therefore, A-mZVI might reduce our dependence on chlorine and antibiotics to

457

tackle problems caused by carcinogenic disinfection byproducts, resistant bacteria and ARGs, etc.

458 459

AUTHOR INFORMATION

460

Corresponding Author

461

*Phone/Fax: +86-27-6786 7535; e-mail: [email protected]; [email protected];

462

[email protected].

463

Notes

464

The authors declare no competing financial interest.

465 466

Acknowledgments: This research was financially supported by Natural Science Funds for

467

Distinguished Young Scholars (Grant 21425728), National Natural Science Foundation of China

468

(Grant 41601543 and 21777050), China Postdoctoral Science Foundation (Grant 2018T110782 and

469

2017M620327), Science Funds for Outstanding Postdocs of Hubei Province, China (Grant Z13), the

470

program of China Scholarship Council (Grant 201706775080), the 111 Project (Grant B17019), and

471

the CAS Interdisciplinary Innovation Team of the Chinese Academy of Sciences.

472 473

ASSOCIATED CONTENT

474

Supporting Information: characterization of ZVI; E. coli inactivation efficiency monitored by HPC

475

method; fluorescence images intensity processing procedures by using ImageJ; measurement of

476

dissolved iron; H2O2 determination; intracellular ROSs level detection; CAT and SOD activity 21

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 35

477

analysis; membrane potential measurement; ATP production rate assay; bacterial sample preparation

478

protocols for SEM and TEM; dot plots of bacterial samples in flow cytometer; TN concentrations;

479

fluorescent images of E. coli in sediment; SEM, XRD and E. coli inactivation performance of

480

C-mZVI; turbidity profile of treated water; ESR spectra of DMPO trapped •O2-; kinetic fitting of

481

H2O2 decomposition; SEM images and EDS of used ZVI; kinetics of ATP generation; particle size

482

distribution of E. coli-ZVI mixture; Comparison of the elementary and Fe0 content between the

483

C-nZVI and A-mZVI; Comparison of the E. coli inactivation performance by employing ZVI in

484

different studies.

485 486

References

487

(1) Rangsivek, R.; Jekel, M. R., Removal of dissolved metals by zero-valent iron (ZVI): Kinetics,

488

equilibria, processes and implications for stormwater runoff treatment. Water Res. 2005, 39, (17),

489

4153-4163.

490

(2) Zou, Y. D.; Wang, X. X.; Khan, A.; Wang, P. Y.; Liu, Y. H.; Alsaedi, A.; Hayat, T.; Wang, X.

491

K., Environmental Remediation and Application of Nanoscale Zero-Valent Iron and Its Composites

492

for the Removal of Heavy Metal Ions: A Review. Environ. Sci. Technol. 2016, 50, (14), 7290-7304.

493

(3) Liu, Y. Q.; Phenrat, T.; Lowry, G. V., Effect of TCE concentration and dissolved groundwater

494

solutes on NZVI-promoted TCE dechlorination and H2 evolution. Environ. Sci. Technol. 2007, 41,

495

(22), 7881-7.

496

(4) Agrawal, A.; Tratnyek, P. G., Reduction of nitro aromatic compounds by zero-valent iron metal.

497

Environ. Sci. Technol. 1996, 30, (1), 153-160.

498

(5) Jia, F. L.; Liu, J.; Zhang, L. Z., Copper Ions Promoted Aerobic Atrazine Degradation by Fe@

499

Fe2O3 Nanowires. Acta Chim. Sinica 2017, 75, (6), 602-607. 22

ACS Paragon Plus Environment

Page 23 of 35

Environmental Science & Technology

500

(6) Ai, Z. H.; Gao, Z. T.; Zhang, L. Z.; He, W. W.; Yin, J. J., Core–shell structure dependent

501

reactivity of Fe@Fe2O3 nanowires on aerobic degradation of 4-chlorophenol. Environ. Sci. Technol.

502

2013, 47, (10), 5344-5352.

503

(7) Qin, H. J.; Li, J. X.; Yang, H. Y.; Pan, B. C.; Zhang, W. M.; Guan, X. H., Coupled effect of

504

ferrous ion and oxygen on the electron selectivity of zerovalent iron for selenate sequestration.

505

Environ. Sci. Technol. 2017, 51, (9), 5090-5097.

506

(8) Mu, Y.; Ai, Z. H.; Zhang, L. Z., Phosphate shifted oxygen reduction pathway on Fe@Fe2O3

507

core–shell nanowires for enhanced reactive oxygen species generation and aerobic 4-chlorophenol

508

degradation. Environ. Sci. Technol. 2017, 51, (14), 8101-8109.

509

(9) Kim, J.; Lee, C.; Love, D. C.; Sedlak, D. L.; Yoon, J.; Nelson, K. L., Inactivation of MS2

510

coliphage by ferrous ion and zero-valent iron nanoparticles. Environ. Sci. Technol. 2011, 45, (16),

511

6978-6984.

512

(10) You, Y. W.; Han, J.; Chiu, P. C.; Jin, Y., Removal and inactivation of waterborne viruses using

513

zerovalent iron. Environ. Sci. Technol. 2005, 39, (23), 9263-9269.

514

(11) Oprckal, P.; Mladenovic, A.; Vidmar, J.; Pranjic, A. M.; Milacic, R.; Scancar, J., Critical

515

evaluation of the use of different nanoscale zero-valent iron particles for the treatment of effluent

516

water from a small biological wastewater treatment plant. Chem. Eng. J. 2017, 321, 20-30.

517

(12) Liu, J. L.; Li, X. Y., Biodegradation and biotransformation of wastewater organics as precursors

518

of disinfection byproducts in water. Chemosphere 2010, 81, (9), 1075-1083.

519

(13) Boyer, T. H.; Singer, P. C., Bench-scale testing of a magnetic ion exchange resin for removal of

520

disinfection by-product precursors. Water Res. 2005, 39, (7), 1265-1276.

521

(14) Chen, Q.; Gao, M.; Li, J.; Shen, F. X.; Wu, Y.; Xu, Z. Q.; Yao, M. S., Inactivation and

522

magnetic separation of bacteria from liquid suspensions using electrosprayed and nonelectrosprayed 23

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 35

523

nZVI particles: Observations and mechanisms. Environ. Sci. Technol. 2012, 46, (4), 2360-2367.

524

(15) Oguma, K.; Katayama, H.; Ohgaki, S., Photoreactivation of Legionella pneumophila after

525

inactivation by low- or medium-pressure ultraviolet lamp. Water Res. 2004, 38, (11), 2757-2763.

526

(16) Chávez de Paz, L. E.; Hamilton, I. R.; Svensäter, G., Oral bacteria in biofilms exhibit slow

527

reactivation from nutrient deprivation. Microbiology 2008, 154, (7), 1927-1938.

528

(17) Chang, P. H.; Juhrend, B.; Olson, T. M.; Marrs, C. F.; Wigginton, K. R., Degradation of

529

extracellular antibiotic resistance genes with UV254 treatment. Environ. Sci. Technol. 2017, 51, (11),

530

6185-6192.

531

(18) Huddleston, J. R., Horizontal gene transfer in the human gastrointestinal tract: potential spread

532

of antibiotic resistance genes. Infect. Drug. Resist. 2014, 7, 167-76.

533

(19) Li, J.; Chen, Q.; Li, X. Y.; Yao, M. S., Rapid point-of-use water purification using nanoscale

534

zero valent iron (nZVI) particles. Sci. Bull. 2014, 59, (29-30), 3926-3934.

535

(20) Diao, M. H.; Yao, M. S., Use of zero-valent iron nanoparticles in inactivating microbes. Water

536

Res. 2009, 43, (20), 5243-5251.

537

(21) Auffan, M.; Achouak, W.; Rose, J.; Roncato, M. A.; Chaneac, C.; Waite, D. T.; Masion, A.;

538

Woicik, J. C.; Wiesner, M. R.; Bottero, J. Y., Relation between the redox state of iron-based

539

nanoparticles and their cytotoxicity toward Escherichia coli. Environ. Sci. Technol. 2008, 42, (17),

540

6730-6735.

541

(22) Lv, Y. C.; Niu, Z. Y.; Chen, Y. C.; Hu, Y. Y., Bacterial effects and interfacial inactivation

542

mechanism of nZVI/Pd on Pseudomonas putida strain. Water Res. 2017, 115, 297-308.

543

(23) Stefaniuk, M.; Oleszczuk, P.; Ok, Y. S., Review on nano zerovalent iron (nZVI): From

544

synthesis to environmental applications. Chem. Eng. J. 2016, 287, 618-632.

545

(24) Phenrat, T.; Saleh, N.; Sirk, K.; Tilton, R. D.; Lowry, G. V., Aggregation and sedimentation of 24

ACS Paragon Plus Environment

Page 25 of 35

Environmental Science & Technology

546

aqueous nanoscale zerovalent iron dispersions. Environ. Sci. Technol. 2007, 41, (1), 284-290.

547

(25) Crane, R. A.; Scott, T. B., Nanoscale zero-valent iron: Future prospects for an emerging water

548

treatment technology. J. Hazard. Mater. 2012, 211-212, 112-125.

549

(26) Choi, H.; Al-Abed, S. R.; Agarwal, S.; Dionysiou, D. D., Synthesis of reactive nano-Fe/Pd

550

bimetallic system-impregnated activated carbon for the simultaneous adsorption and dechlorination

551

of PCBs. Chem. Mater. 2008, 20, (11), 3649-3655.

552

(27) He, F.; Zhao, D. Y.; Liu, J. C.; Roberts, C. B., Stabilization of Fe−Pd nanoparticles with sodium

553

carboxymethyl cellulose for enhanced transport and dechlorination of trichloroethylene in soil and

554

groundwater. Ind. Eng. Chem. Res. 2007, 46, (1), 29-34.

555

(28) He, F.; Zhao, D. Y., Preparation and characterization of a new class of starch-stabilized

556

bimetallic nanoparticles for degradation of chlorinated hydrocarbons in water. Environ. Sci. Technol.

557

2005, 39, (9), 3314-3320.

558

(29) Li, Z. Q.; Greden, K.; Alvarez, P. J. J.; Gregory, K. B.; Lowry, G. V., Adsorbed polymer and

559

NOM limits adhesion and toxicity of nano scale zerovalent iron to E. coli. Environ. Sci. Technol.

560

2010, 44, (9), 3462-3467.

561

(30) Zhou, L.; Thanh, T. L.; Gong, J.; Kim, J. H.; Kim, E. J.; Chang, Y. S., Carboxymethyl cellulose

562

coating decreases toxicity and oxidizing capacity of nanoscale zerovalent iron. Chemosphere 2014,

563

104, 155-161.

564

(31) Ayush, V.; Francesco, S., Effect of surface properties on nanoparticle–cell interactions. Small

565

2010, 6, (1), 12-21.

566

(32) Chen, J. W.; Xiu, Z. M.; Lowry, G. V.; Alvarez, P. J. J., Effect of natural organic matter on

567

toxicity and reactivity of nano-scale zero-valent iron. Water Res. 2011, 45, (5), 1995-2001.

568

(33) Liu, Y. Q.; Choi, H.; Dionysiou, D.; Lowry, G. V., Trichloroethene hydrodechlorination in 25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 35

569

water by highly disordered monometallic nanoiron. Chem. Mater. 2005, 17, (21), 5315-5322.

570

(34) Nurmi, J. T.; Tratnyek, P. G.; Sarathy, V.; Baer, D. R.; Amonette, J. E.; Pecher, K.; Wang, C.;

571

Linehan, J. C.; Matson, D. W.; Penn, R. L.; Driessen, M. D., Characterization and properties of

572

metallic iron nanoparticles:  Spectroscopy, electrochemistry, and kinetics. Environ. Sci. Technol.

573

2005, 39, (5), 1221-1230.

574

(35) Wang, Q. L.; Lee, S.; Choi, H., Aging study on the structure of Fe0-nanoparticles: Stabilization,

575

characterization, and reactivity. J. Phys. Chem. C 2010, 114, (5), 2027-2033.

576

(36) Yániz, J.; Palacín, I.; Vicente-Fiel, S.; Gosálvez, J.; López-Fernández, C.; Santolaria, P.,

577

Evaluation of a commercial kit based on acridine orange/propidium iodide to assess the plasma

578

membrane integrity of ram sperm. Span. J. Agric. Res. 2013, 11, (2), 362-365.

579

(37) Harvey Jr, A. E.; Smart, J. A.; Amis, E., Simultaneous spectrophotometric determination of iron

580

(II) and total iron with 1, 10-phenanthroline. Anal. Chem. 1955, 27, (1), 26-29.

581

(38) Chen, N.; Huang, Y. H.; Hou, X. J.; Ai, Z. H.; Zhang, L. Z., Photochemistry of hydrochar:

582

Reactive oxygen species generation and sulfadimidine degradation. Environ. Sci. Technol. 2017, 51,

583

(19), 11278-11287.

584

(39) Sun, H. W.; Li, G. Y.; An, T. C.; Zhao, H. J.; Wong, P. K., Unveiling the photoelectrocatalytic

585

inactivation mechanism of Escherichia coli: Convincing evidence from responses of parent and

586

anti-oxidation single gene knockout mutants. Water Res. 2016, 88, 135-143.

587

(40) Wang, C. B.; Zhang, W. X., Synthesizing nanoscale iron particles for rapid and complete

588

dechlorination of TCE and PCBs. Environ. Sci. Technol. 1997, 31, (7), 2154-2156.

589

(41) Chekli, L.; Bayatsarmadi, B.; Sekine, R.; Sarkar, B.; Shen, A. M.; Scheckel, K. G.; Skinner, W.;

590

Naidu, R.; Shon, H. K.; Lombi, E.; Donner, E., Analytical characterisation of nanoscale zero-valent

591

iron: A methodological review. Anal. Chim. Acta 2016, 903, 13-35. 26

ACS Paragon Plus Environment

Page 27 of 35

Environmental Science & Technology

592

(42) US EPA. Gold Book-Water Quality for Water, 1986.

593

(43) US PEA. Aquatic Life: Ambient Water Quality Criteria for Ammonia-Freshwater, 2013,

594

EPA-822-R-13-001.

595

(44) World Health Organization, Concise international chemical assessment document 15:

596

1,2-diaminoethane (ethylenediamine), Geneva, 1999

597

(45) Sun, H. W.; An, T. C.; Li, G. Y.; Qiao, M.; Wei, D. B., Distribution, possible sources, and

598

health risk assessment of SVOC pollution in small streams in Pearl River Delta, China. Environ. Sci.

599

Pollut. Res. 2014, 21, (17), 10083-10095.

600

(46) Chen, Q.; Li, J.; Wu, Y.; Shen, F. X.; Yao, M. S., Biological responses of Gram-positive and

601

Gram-negative bacteria to nZVI (Fe0), Fe2+ and Fe3+. RSC Adv. 2013, 3, (33), 13835-13842.

602

(47) Lee, C.; Kim, J. Y.; Lee, W. I.; Nelson, K. L.; Yoon, J.; Sedlak, D. L., Bactericidal effect of

603

zero-valent iron nanoparticles on Escherichia coli. Environ. Sci. Technol. 2008, 42, (13), 4927-4933.

604

(48) Dong, H. R.; Xie, Y. K.; Zeng, G. M.; Tang, L.; Liang, J.; He, Q.; Zhao, F.; Zeng, Y. L.; Wu,

605

Y. A., The dual effects of carboxymethyl cellulose on the colloidal stability and toxicity of nanoscale

606

zero-valent iron. Chemosphere 2016, 144, 1682-1689.

607

(49) Turbidity and TSS (accessed on Feb 21, 2019).

608

http://www.lakesuperiorstreams.org/understanding/param_turbidity.html

609

(50) Lefevre, E.; Bossa, N.; Wiesner, M. R.; Gunsch, C. K., A review of the environmental

610

implications of in situ remediation by nanoscale zero valent iron (nZVI): Behavior, transport and

611

impacts on microbial communities. Sci. Total Environ. 2016, 565, 889-901.

612

(51) He, D.; Ma, J. X.; Collins, R. N.; Waite, T. D., Effect of structural transformation of

613

nanoparticulate zero-valent iron on generation of reactive oxygen species. Environ. Sci. Technol.

614

2016, 50, (7), 3820-3828. 27

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 35

615

(52) Mu, Y.; Jia, F. L.; Ai, Z. H.; Zhang, L. Z., Molecular Oxygen Activation with Nano Zero-valent

616

Iron for Aerobic Degradation of Organic Contaminants and the Performance Enhancement. Acta

617

Chim. Sinica 2017, 75, (6), 538-543.

618

(53) Sun, H. W.; Li, G. Y.; Nie, X.; Shi, H. X.; Wong, P. K.; Zhao, H. J.; An, T. C., Systematic

619

approach to in-depth understanding of photoelectrocatalytic bacterial inactivation mechanisms by

620

tracking the decomposed building blocks. Environ. Sci. Technol. 2014, 48, (16), 9412-9419.

621

(54) Liu, W.; Ai, Z. H.; Cao, M. H.; Zhang, L. Z., Ferrous ions promoted aerobic simazine

622

degradation with Fe@Fe2O3 core–shell nanowires. Appl. Catal. B-Environ. 2014, 150-151, 1-11.

623

(55) Hou, X. J.; Huang, X. P.; Jia, F. L.; Ai, Z. H.; Zhao, J. C.; Zhang, L. Z., Hydroxylamine

624

promoted goethite surface Fenton degradation of organic pollutants. Environ. Sci. Technol. 2017, 51,

625

(9), 5118-5126.

626

(56) Ezraty, B.; Gennaris, A.; Barras, F.; Collet, J. F., Oxidative stress, protein damage and repair in

627

bacteria. Nat. Rev. Microbiol. 2017, 15, (7), 385-396.

628

(57) Kakhlon, O.; Cabantchik, Z. I., The labile iron pool: characterization, measurement, and

629

participation in cellular processes. Free Radical Bio. Med. 2002, 33, (8), 1037-1046.

630

(58) Cabiscol Català, E.; Tamarit Sumalla, J.; Ros Salvador, J., Oxidative stress in bacteria and

631

protein damage by reactive oxygen species. Int. Microbiol. 2000, 3, (1), 3-8.

632

(59) Akhavan, O.; Ghaderi, E.; Esfandiar, A., Wrapping bacteria by graphene nanosheets for

633

isolation from environment, reactivation by sonication, and inactivation by near-infrared irradiation.

634

J. Phys. Chem. B 2011, 115, (19), 6279-6288.

635

(60) Kim, J. Y.; Park, H. J.; Lee, C.; Nelson, K. L.; Sedlak, D. L.; Yoon, J., Inactivation of

636

Escherichia coli by nanoparticulate zerovalent iron and ferrous ion. Appl. Environ. Microbiol. 2010,

637

76, (22), 7668-70. 28

ACS Paragon Plus Environment

Page 29 of 35

Environmental Science & Technology

638

(61) Liu, A. R.; Liu, J.; Zhang, W. X., Transformation and composition evolution of nanoscale zero

639

valent iron (nZVI) synthesized by borohydride reduction in static water. Chemosphere 2015, 119,

640

1068-1074.

641

(62) Cornell, R. M.; Schwertmann, U., The iron oxides: structure, properties, reactions, occurrences

642

and uses. John Wiley & Sons: 2003.

643

(63) Noubactep, C., On the mechanism of microbe inactivation by metallic iron. J. Hazard. Mater.

644

2011, 198, 383-6.

645

(64) Shin, E. H.; Li, Y.; Kumar, U.; Sureka, H. V.; Zhang, X.; Payne, C. K., Membrane potential

646

mediates the cellular binding of nanoparticles. Nanoscale 2013, 5, (13), 5879-86.

647

(65) Warren, E. A. K.; Payne, C. K., Cellular binding of nanoparticles disrupts the membrane

648

potential. RSC Adv. 2015, 5, (18), 13660-13666.

649

(66) Xiu, Z. M.; Jin, Z. H.; Li, T. L.; Mahendra, S.; Lowry, G. V.; Alvarez, P. J. J., Effects of

650

nano-scale zero-valent iron particles on a mixed culture dechlorinating trichloroethylene.

651

Bioresource Technol. 2010, 101, (4), 1141-1146.

652

(67) Schwarz-Linek, J.; Arlt, J.; Jepson, A.; Dawson, A.; Vissers, T.; Miroli, D.; Pilizota, T.;

653

Martinez, V. A.; Poon, W. C. K., Escherichia coli as a model active colloid: A practical

654

introduction. Colloids and Surfaces B: Biointerfaces 2016, 137, 2-16.

655 656

29

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 35

657 658

Figure 1. The typical SEM images of (a) C-nZVI and (b, c) A-mZVI. (d) XRD of A-mZVI and

659

C-nZVI.

660 661

Figure 2. (a) The temporary change of E. coli concentrations in the supernatants during ZVI

662

inactivation processes. (b) The ratio of survived E. coli among the total cells in the supernatants 30

ACS Paragon Plus Environment

Page 31 of 35

Environmental Science & Technology

663

treated by C-nZVI and A-mZVI, respectively. (c) The removal efficiencies of total E. coli cells from

664

the supernatants by C-nZVI and A-mZVI as a function of time. (d) The temporary trend of the

665

bacterial viability in the sediment samples as reflected by the intensity ratios of green and red

666

fluorescence. The ZVI dose was 100 mg/L, and the initial concentration of E. coli was OD670nm =

667

0.06 in 0.85% NaCl.

668 669

Figure 3. (a) The survival ratio and (b) removal efficiency of E. coli in the supernatant as a function

670

of ZVI dose. The impacts of ROSs scavengers on (c) inactivation of E. coli and (d) removal

671

efficiency of E. coli in the supernatant. The duration of disinfection was 60 min, and the initial

672

concentration of E. coli was OD670nm = 0.06 in 0.85% NaCl. For scavenging experiments, the ZVI

673

dose was 200 mg/L and the concentrations of SOD, catalase and TBA were 100 mg/L, 100 mg/L and

674

1% (v/v), respectively.

31

ACS Paragon Plus Environment

Environmental Science & Technology

Page 32 of 35

675 676

Figure 4. (a) The temporary change of dissolved iron concentrations from C-nZVI and A-mZVI

677

during the disinfection process. (b) Time profiles of ESR intensities of DMPO-•O2- adduct. (c)

678

Accumulation of H2O2 as a function of time. (d) Decomposition profile of H2O2 by C-nZVI and

679

A-mZVI. (e) Evolution of intracellular ROSs level within E. coli treated by C-nZVI and A-mZVI. (f)

680

Time profiles of ESR intensities generated by DMPO trapped •OH (symbols + lines), and the

681

proportions of surface bound •OH among overall •OH production (bars). The ZVI dose was 100

682

mg/L and the concentration of F- was 2 × 10-3 mol/L.

32

ACS Paragon Plus Environment

Page 33 of 35

Environmental Science & Technology

683 684

Figure 5. The evolution of enzymatic activities of (a) catalase and (b) superoxide dismutase (SOD)

685

as a function of ZVI inactivation time. (c) The time profile of bacterial membrane potential

686

represented by the fluorescent intensity of Rh123 stained cells. (d) The impact of ZVI on ATP

687

synthesis capacity of E. coli. The ZVI dose was 100 mg/L.

33

ACS Paragon Plus Environment

Environmental Science & Technology

Page 34 of 35

688 689

Figure 6. The SEM (a, b, c) and TEM (d, e, f) images of the E. coli cells before (a, d) and after 1 h

690

of disinfection treating by C-nZVI (b, e) and A-mZVI (c, f). (g) The XRD patterns and (h) the

691

magnetic hysteresis loops of the corrosion products generated by C-nZVI and A-mZVI after 1 h of

692

aerobic disinfection. (i) The sedimentation curves of ZVI-E. coli mixture after 1 h of aerobic

693

disinfection, w/ and w/o a magnet.

694 695

Scheme 1. The possible physical removal processes of E. coli from supernatant by C-nZVI and

696

A-mZVI. 34

ACS Paragon Plus Environment

Page 35 of 35

697

Environmental Science & Technology

TOC Art Figure

698

35

ACS Paragon Plus Environment