PDF (2 MB)

225. DO concentrations were at 3.21, 6.76, and 9.23 mg/L, respectively. The majority. 226. (80.23%) of SLMoS2 nanosheets dissolved with the supplement...
2 downloads 0 Views 2MB Size
Subscriber access provided by Bethel University

Ecotoxicology and Human Environmental Health

Dissolved Oxygen and Visible Light Irradiation Drive the Structural Alterations and Phytotoxicity Mitigation of Single-layer Molybdenum Disulfide Wei Zou, Qixing Zhou, Xingli Zhang, and Xiangang Hu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b00088 • Publication Date (Web): 04 Jun 2019 Downloaded from http://pubs.acs.org on June 4, 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 41

Environmental Science & Technology

TOC

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 41

1

Dissolved Oxygen and Visible Light Irradiation Drive the Structural Alterations

2

and Phytotoxicity Mitigation of Single−layer Molybdenum Disulfide

3

Wei Zou1, Qixing Zhou2, Xingli Zhang1, Xiangang Hu2,*

4

1 School

5

Environment and Pollution Control, Ministry of Education, Henan Key Laboratory for

6

Environmental Pollution Control, Henan Normal University, Xinxiang

7

453007, China.

8

2

9

Education)/Tianjin Key Laboratory of Environmental Remediation and Pollution

10

Control, College of Environmental Science and Engineering, Nankai University,

11

Tianjin 300350, China.

12

Corresponding authors: Xiangang Hu, [email protected].

13

Fax, 0086−022−23507800; Tel, 0086−022−23507800.

of Environment, Key Laboratory for Yellow River and Huai River Water

Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of

14 15

ABSTRACT

16

Understanding environmental fate is a prerequisite for the safe application of

17

nanoparticles.

18

transformation of single−layer molybdenum disulfide (SLMoS2, a 2D nanosheet

19

attracting substantial attention in various fields) remain largely unknown. The present

20

work found that the dissolution of SLMoS2 was pH and dissolved oxygen dependent

21

and that alterations in phase composition significantly occur under visible light

22

irradiation. The 1T phase was preferentially oxidized to yield soluble species (MoO42−

However,

the

fundamental

persistence

1

ACS Paragon Plus Environment

and

environmental

Page 3 of 41

Environmental Science & Technology

23

and SO42−), and the 2H phase remained as a residual. The transformed SLMoS2

24

exhibited a ribbon−like and multilayered structure and low colloidal stability due to

25

the loss of surface charge. Dissolved oxygen competitively captured the electrons of

26

SLMoS2 to generate superoxide radicals and accelerated the dissolution of

27

nanosheets. Compared to pristine 1T−phase SLMoS2, the transformed 2H−phase

28

SLMoS2 could not easily enter algal cells and induced a low developmental inhibition,

29

oxidative stress, plasmolysis, photosynthetic toxicity and metabolic perturbation. The

30

downregulation of amino acids and upregulation of unsaturated fatty acids contributed

31

to the higher toxicity of 1T−phase SLMoS2. The dissolved ions did not induce

32

apparent phytotoxicity. The connections between environmental transformation

33

(phase change and ion release) and phytotoxicity provide insights into the safe design

34

and evaluation of 2D nanomaterials.

35 36

KEYWORDS: nanoparticle, phytotoxicity, environmental transformation, phase

37

alteration, nanotoxicity

38 39

INTRODUCTION

40

Single−layer molybdenum disulfide (SLMoS2), a typical two−dimensional (2D)

41

transition metal dichalcogenide (TMD), has been widely applied in electronics and

42

optoelectronics,1 catalysis,2,

3

43

environmental protection.8,

Given the potential of environmental and organism

44

exposure to SLMoS2 during the nanomaterial lifecycle (e.g., fabrication, use and

9

energy storage,4 biology,5 biomedicine,6,

2

ACS Paragon Plus Environment

7

and

Environmental Science & Technology

45

disposal), environmental fate and safety should be considered in detail.10, 11 Compared

46

to pristine nanoparticles, environmentally transformed nanoparticles may exhibit

47

different morphologies, structures and stabilities, leading to the alteration of

48

ecological and health effects.12 For example, light irradiation has driven changes in

49

graphene morphology and reduced the toxicity of graphene to aquatic algae.13 Ion

50

release from metal and metallic oxide nanoparticles in water has contributed to the

51

enhancement or mitigation of nanotoxicity.14,

52

oxygen (DO) and visible light irradiation, which are both common environmental

53

factors, in the nanoparticle properties (e.g., phase change), environmental stability

54

(e.g., ion release) and ecotoxicity of SLMoS2 remains largely unknown.

55

15

However, the role of dissolved

The SLMoS2 surface can be quickly oxidized to produce molybdenum oxide

56

(MoO3) at defect sites after exposure to oxygen at temperatures above 340°C.16 In the

57

presence of oxygen or other strong oxidizing agents (hydroxyl radicals and H2O2),

58

SLMoS2 can be fast−etched and release soluble species.17 These results demonstrate

59

that SLMoS2 can be oxidized under extensive conditions. At room temperature,

60

sunlight−induced reactions (oxidation and reduction) also affect the oxidation state,

61

the generation of reactive oxygen species (ROS), and the persistence of

62

nanomaterials, although the process is slow.18 Photoactive nanomaterials, including

63

metal− and carbon−based nanoparticles, can absorb visible light and react with

64

oxygen to produce ROS.19 As a typical semiconductor material, SLMoS2 exhibits a

65

lower bandgap (1.8 eV) than that of TiO2 nanoparticles (3.0~3.2 eV),20 which is

66

beneficial for visible light adsorption. Herein, it was hypothesized that the 3

ACS Paragon Plus Environment

Page 4 of 41

Page 5 of 41

Environmental Science & Technology

67

environmental transformation of SLMoS2 can occur slowly in an oxygen−containing

68

aqueous phase with visible light irradiation at room temperature.

69

Nanotoxicity is highly relevant to the physicochemical properties (e.g., morphology,

70

colloidal stability, layer structure and crystal structure) of nanomaterials.13, 21,22 For

71

example, anatase TiO2 with octahedral coordination was more chemically active than

72

rutile TiO2 with square prismatic coordination in the production of ROS, which are

73

positively related to the toxicity of TiO2 in vitro.23 Environmental factors may change

74

the morphology and phase of SLMoS2, leading to the enhancement or mitigation of

75

nanotoxicity. The objective of the present study is to understand the physicochemical

76

transformation of SLMoS2 nanosheets in oxygen−containing water with visible light

77

irradiation and the effects on the aquatic toxicity of the nanosheets. pH, as another

78

environmental factor, is also taken into consideration. Specifically, the following

79

issues are addressed: (i) the chemical dissolution kinetics of SLMoS2; (ii) the

80

morphology, colloidal stability, layer structure and phase alterations of SLMoS2; and

81

(iii) the ecotoxicity (e.g., photosynthesis, ROS and metabolic profile) of pristine and

82

transformed SLMoS2 to a model species (Chlorella vulgaris). The results provide

83

insights into the safe design and evaluation of 2D nanomaterials. By analyzing the

84

connections between environmental transformation (phase change and ion release of

85

nanoparticles) and ecotoxicity, the present work will provide insights into the safety

86

evaluation and design of SLMoS2.

87 88

MATERIALS AND METHODS 4

ACS Paragon Plus Environment

Environmental Science & Technology

89

Environmental Transformation of SLMoS2 Nanosheets

90

Pristine SLMoS2 nanosheets (chemically exfoliated, single layer>99%) were obtained

91

from Nanjing XFNANO Materials Tech Co., Ltd., China. Deionized water (18.2

92

Ω/cm) was heated to boiling to remove DO and then cooled to room temperature in a

93

nitrogen−filled glovebox. SLMoS2 nanosheets (5 mg) were suspended in 200 mL of

94

prepared water. The SLMoS2 suspension (25 mg/L) was placed in a shaking incubator

95

(24°C, 150 rpm, humidity 80%) and incubated for eight weeks with and without

96

irradiation (light/dark = 14:10) by a xenon arc lamp (CEL−HXF300, Ceaulight,

97

Beijing, China) with a UV cutoff (λ < 420 nm) of 35 W/m2. To analyze the effects of

98

pH, the pH values of the SLMoS2 suspensions were adjusted to 3~11 by 0.1 mM HCl

99

and 0.1 mM NaOH. To study the effect of DO, the suspension was aerated with pure

100

oxygen (0.01, 0.05, 0.1, and 0.5 L/min), and the tested concentrations of DO were

101

3.21, 6.76, 9.23 and 9.52 mg/L (saturated), respectively. The groups placed in a

102

nitrogen−filled glovebox were set up as the control (0 mg/L DO). After eight weeks

103

of reaction, the SLMoS2 nanosheets were separated using a 0.1 μm polyether sulfone

104

(PES) filter and then lyophilized. To analyze the release kinetics of ions from the

105

SLMoS2 nanosheets, the filtrates on the 7th, 14th, 21st, 28th, and 56th days were

106

collected by ultrafiltration centrifugation (Amicon Ultra−15 3kD, Millipore, USA).

107

Then, the contents of the dissolved Mo and S species were determined by inductively

108

coupled plasma mass spectrometry (ICP−MS, Agilent 7700, USA). Each treatment

109

was performed in triplicate.

110

Characterization of Pristine and Transformed SLMoS2 5

ACS Paragon Plus Environment

Page 6 of 41

Page 7 of 41

Environmental Science & Technology

111

The transformed SLMoS2 samples in the groups (pH=3, 7, and 11) with saturated DO

112

(9.52 mg/L) after 56 days of incubation (named T3−SLMoS2, T7−SLMoS2, and

113

T11−SLMoS2, respectively) were collected. Then, the morphology, structure, colloidal

114

stability, and chemical and optical properties of pristine and transformed SLMoS2

115

were characterized by transmission electron microscopy (TEM, HT7700, Hitachi,

116

Japan), atomic force microscopy (AFM, Agilent 5420, USA), Raman spectroscopy

117

(Thermo Scientific, DXR, USA), dynamic light scattering (DLS, Brookhaven, USA),

118

electrochemical workstation analysis (CHI660E, Shanghai Chenhua, China), UV−vis

119

analysis (T90, Purkinje General, China), Fourier transform infrared spectroscopy

120

(FTIR, Bruker Tensor 27, USA), X−ray photoelectron spectroscopy (XPS, Kratos,

121

Japan), and X−ray powder diffraction (XRD, Rigaku Ultima IV, Japan). The details

122

are provided in the supporting information (SI).

123

Free Radical Measurements

124

Peroxy radical (•O2−) was determined using the spin trap

125

5,5−dimethyl−1−pyrroline−N−oxide (DMPO, 10 mM) in methanol solution using an

126

electron spin resonance (ESR) spectroscope (MiniScope 400, Germany). •OH was

127

detected by monitoring the fluorescence intensity of 2−hydroxyterephthalic acid

128

(TAOH) at 435 nm in aqueous solution using a UV−vis spectrometer (T90, Purkinje

129

General, China). The details are provided in the SI.

130

C. vulgaris Cultivation

131

C. vulgaris and its culture medium (BG−11) were obtained from the Freshwater

132

Algae Culture Collection at the Institute of Hydrobiology, China. The environmental 6

ACS Paragon Plus Environment

Environmental Science & Technology

133

exposure of nanoparticles at high concentrations can occur in relevant wastewater

134

from manufacturing facilities, although the environmentally relevant concentrations

135

are currently unclear. Significant toxicity was observed when the SLMoS2

136

concentration was over 10 mg/L in a previous cytotoxicity study.24 Given the rapid

137

development of 2D nanomaterials, nanomaterials at 1, 10, and 25 mg/L were prepared

138

in algal BG−11 culture medium to compare the toxicology of pristine and transformed

139

SLMoS2. The initial density of algal cells was 1.5×105 CFU/mL. SLMoS2 was

140

exposed to algal cells in 250 mL glass flasks, and algal cells without SLMoS2 were

141

used as controls. To study the effects of released molybdate salts, the algae were also

142

exposed to 100 μM Na2MoO4 (the corresponding mass concentration of MoO42− was

143

15.9 mg/L), which was higher than the concentrations of ions released by SLMoS2 at

144

25 mg/L. The glass flasks were shaken at 150 rpm for 10 min once every 8 h and

145

placed in an illumination incubator (LRH−250 Gb, China) at 24.0 ±0.5°C and 80%

146

humidity. Each treatment was performed in triplicate. The effects of SLMoS2 on algal

147

division were measured by counting the number of algal cells by flow cytometry

148

(FCM, BD FACSCalibur, USA) at 24, 48, 72 and 96 h.

149

Cellular Ultrastructure Observation, Nanomaterial Uptake and ROS

150

The cell suspension (5 mL) was centrifuged at 9000 g for 5 min, and the supernatant

151

was discarded. The pellets were washed with phosphate buffered saline (PBS), and

152

then the algal cells were fixed in 2.5% glutaraldehyde at 4°C overnight, post fixed in

153

1% osmium tetroxide for 2 h, dehydrated in an ethanol gradient (30%, 50%, 70%,

154

80%, 90%, 95%, and 100%), and then embedded in an epoxy resin. Ultrathin sections 7

ACS Paragon Plus Environment

Page 8 of 41

Page 9 of 41

Environmental Science & Technology

155

(approximately 90 nm) of algal cells were cut using a diamond knife on an

156

ultramicrotome (EM FC7, Leica, Germany) and stained with uranyl acetate and lead

157

citrate for 15 min. The distribution of SLMoS2 was observed by TEM (HT7700,

158

Hitachi, Japan), and the uptake of SLMoS2 was quantified by ICP−MS (Agilent 7700,

159

USA). In brief, a known amount of 105 algal cells was collected and digested using

160

HNO3/H2O2 (v:v, 3:1) until no color was observed. After filtration through a 0.45 μm

161

water membrane, the concentration of Mo ions was determined by ICP−MS (Agilent

162

7700, USA). The endocytosis pathway and mechanisms of SLMoS2 by algal cells

163

were investigated using physical and pharmacological inhibitors. The algal cells were

164

preincubated at 4°C for 1 h to inhibit energy−dependent uptake, and then, the SLMoS2

165

nanosheets were exposed for 1 h at room temperature. To determine the specific

166

mechanisms of cellular uptake, the algal cells were pretreated with

167

methyl−betacyclodextrin (MβCD, 20 mM), chlorpromazine hydrochloride (CPZ, 100

168

μM), and 5−(N−ethyl−N−isopropyl) amiloride (EIPA, 50 μM) for 1 h. Subsequently,

169

SLMoS2 was exposed to the algae for 1 h. The cellular contents of Mo ions in algal

170

cells were detected as described above. Based on the S/Mo molar ratio (approximately

171

2) of pristine and transformed SLMoS2 (equation S1, details below), the concentration

172

of Mo ions was then converted to the SLMoS2 nanosheet content per 105 cells. The

173

measurement of ROS is presented in the SI.

174

Algal Metabolism

175

The algal suspension (30 mL) was centrifuged at 9000 g for 5 min to collect cells. To

176

completely break the cell walls, the collected algal cells underwent three cycles of 8

ACS Paragon Plus Environment

Environmental Science & Technology

177

freezing in liquid nitrogen and thawing at room temperature. Subsequently, 4.5 mL of

178

a methanol/chloroform/water (volume ratio=2.5:1:1) solution was added to the cell

179

suspension, and the cells were completely broken using an ultrasonic probe (150 W,

180

10 min) in an ice−water bath. The metabolites were extracted usingsonication (200 W,

181

30 min), followed by centrifugation at 9000 g for 5 min at 4°C. After sonication and

182

centrifugation, the supernatant was collected, and the pellet was extracted again as

183

described above. The supernatant was mixed with the previously collected

184

supernatant. Then, water (1.5 mL) was added to the above supernatant and

185

centrifuged at 9000 g for 5 min. The lower phase was separated and filtered through a

186

10 cm silica gel column, followed by nitrogen blowing. The upper phases, which

187

consisted of methanol and water, were dried by nitrogen blowing and lyophilization,

188

respectively, and then mixed with the residual lower−layer phase.

189

N−methyl−N−(trimethylsilyl)trifluoroacetamide (80 μL) and methoxamine

190

hydrochloride (20 mg/mL, 50 μL) were added as derivatives. After derivatization, the

191

samples (1 μL) were injected into a gas chromatography column (HP−5MS, Agilent,

192

USA) in split mode (1:25), and the metabolites were identified. Metabolic analysis

193

was conducted using gas chromatography with mass spectrometry (GC−MS;

194

6890N/5973, Agilent, USA). The MS was operated in full scan mode with a detection

195

slope of m/z 80−800. The metabolites were identified using the National Institute of

196

Standards and Technology (NIST 14.0) mass spectra library in ChemStation software.

197

Statistical Analysis

198

All experiments were performed in triplicate. The results are presented as the mean ± 9

ACS Paragon Plus Environment

Page 10 of 41

Page 11 of 41

Environmental Science & Technology

199

standard deviation. The statistical significance was analyzed with one−way analysis

200

of variance (ANOVA) followed by Tukey’s test. A p value less than 0.05 was

201

considered statistically significant. All statistical analyses were performed using IBM

202

SPSS 22.0.

203 204

RESULTS AND DISCUSSION

205

Visible Light Irradiation, DO and Alkalinity Accelerate Ion Release from

206

SLMoS2 Nanosheets

207

The time− and pH−dependent dissolution of Mo and S species from SLMoS2

208

nanosheets is presented in Figure 1. The concentrations of released Mo and S species

209

at pH=11 after 56 days of reaction were 3.122 and 2.088 mg/L, respectively, which

210

were higher than those at pH=3 (0.631 and 0.424 mg/L for released Mo and S species,

211

respectively). MoS2 nanosheets produced soluble ions in aqueous media accompanied

212

by protons,10 and the hydroxyl ions in alkaline conditions contributed to the

213

dissolution of SLMoS2 (equation S1 in the SI). Importantly, SLMoS2 dissolved slowly

214

without oxygen (Figure 1a−1b). The concentrations of released Mo and S species at

215

pH=11 reached 6.218 and 4.405 mg/L, respectively, when DO was supplemented

216

(Figure 1c−1d). The details of dissolved Mo released from SLMoS2 under DO

217

conditions at different pH values are provided in Table S1. Visible light irradiation

218

also significantly increased the release of Mo and S (Figure 1e−f), with approximately

219

fourfold higher dissolved concentrations than those in the dark (Table S1). The

220

highest contents of Mo and S species after 56 days of irradiation in the alkaline 10

ACS Paragon Plus Environment

Environmental Science & Technology

221

groups (pH=11) were 12.035 and 7.836 mg/L, respectively. Noticeably, the soluble

222

ionic contents significantly increased with the DO concentration in the

223

visible−light−irradiated groups (Figure S1). The contents of dissolved Mo species at

224

the 56th day in the pH=11 groups reached 9.017, 10.518, and 11.568 mg/L when the

225

DO concentrations were at 3.21, 6.76, and 9.23 mg/L, respectively. The majority

226

(80.23%) of SLMoS2 nanosheets dissolved with the supplementation of saturated DO

227

(9.52 mg/L) and visible light irradiation in the pH=11 groups. In contrast, 20.81% and

228

41.45% of the nanosheets were dissolved under dark and single−DO treatment,

229

respectively (Table S1). The above data suggested that DO and visible light

230

irradiation significantly accelerated the chemical dissolution of SLMoS2. SLMoS2

231

was chemically oxidized, while molybdate ions (MoO42−) and sulfate (SO42−) were the

232

detected main products, as shown in equation S1. XPS spectra revealed the existence

233

of a Mo6+ 3d band in SLMoS2, ascribed to molybdenum trioxide (MoO3). The Mo6+

234

3d band in the XPS spectrum disappeared after visible light treatment (as presented

235

below), implying the dissolution of MoO3. MoO3 can also be dissolved to molybdate

236

in alkaline conditions (equation S2), further producing excess MoO42−. Therefore, the

237

molar ratios of soluble S/Mo were less than 2 in the visible−light−treated groups

238

(Figure S2), as a result of the release of extra MoO42−due to the photoreduction of

239

MoO3 in the SLMoS2 nanosheets.

240

Visible Light Irradiation, DO and Alkalinity Promote Phase Alteration

241

According to the UV−vis spectra (Figure 2a), pristine SLMoS2 exhibited a

242

considerable adsorption band at ~217 nm, which was attributed to the predominance 11

ACS Paragon Plus Environment

Page 12 of 41

Page 13 of 41

Environmental Science & Technology

243

of the metallic 1T phase.25, 26 With the release of soluble species, the absorption

244

intensities at 200−300 nm significantly decreased, indicating the dissolution of the 1T

245

phase. Compared to pristine SLMoS2, the adsorption band of transformed SLMoS2

246

was blueshifted (~10 nm). The characteristic peaks of the 2H phase at ~610 and ~670

247

nm became strong, and the absorbance at 500−800 nm increased (Figure 2b),

248

suggesting a reduction in the 1T phase and the enrichment of the 2H phase. The Mo

249

3d spectra in Figure 2c revealed that the 1T and 2H phases accounted for 72.9% and

250

27.1%, respectively, of pristine SLMoS2. However, the proportion of the 1T phase

251

significantly decreased with exposure to DO and visible light irradiation. The

252

proportion of the 1T component decreased from 72.9% in pristine SLMoS2 to 57.4%,

253

43.5%, and 11.4% in T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2, respectively,

254

which was consistent with the UV−vis spectra in Figure 2b. A significant decrease in

255

the 1T phase of SLMoS2 in the visible−light−irradiation group with the increase in

256

exposed DO is also observed in Figure S3, implying the crucial role of DO in the

257

phase alteration of SLMoS2. Additionally, the Mo 3d core−level spectrum of pristine

258

SLMoS2 consisted of three peaks located at ~235.5, ~232, and ~229 eV, which were

259

related to Mo6+ and 3d3/2 and 3d5/2 of Mo4+, respectively. Visible light irradiation

260

induced the dissolution of Mo oxides (Figure 1), resulting in the disappearance of the

261

3d Mo6+ band in T11−SLMoS2. The 3d Mo4+ peaks shifted to higher binding energies

262

in Figure 2c, confirming the enrichment of the 2H phase in transformed SLMoS2. The

263

alteration in the phase composition of SLMoS2 was further verified by the S 2p

264

core−level spectrum in Figure 2d. The peaks of the S 2p1/2 and S 2p3/2 bands also 12

ACS Paragon Plus Environment

Environmental Science & Technology

265

shifted to higher binding energies, suggesting the enrichment of the 2H phase in the

266

transformed SLMoS2, especially in T11−SLMoS2. Another convincing result was the

267

distinguishable broad band in transformed SLMoS2, which consisted of two peaks

268

ascribed to 2H doublets.

269

The crystal structure alteration of SLMoS2 was also investigated by XRD. The

270

characteristic XRD peaks centered at 2θ values of 14.92° (002), 39.28° (103) and

271

44.26° (006) were observed in all samples (Figure S4), suggesting that the major

272

component was SLMoS2 nanosheets.27 With the dissolution, the (002) reflection,

273

which was associated with the interplanar crystal spacing of SLMoS2,28 was

274

significantly blueshifted, and the width of the peak became narrow. The (002)

275

reflection peak of T11-SLMoS2 was centered at a 2θ of 14.46°, which was similar to

276

that (a 2θ of 14.4°) of pure 2H MoS2.28 The intensities of the (100), (103) (105) and

277

(112) reflections became stronger, while the (105) and (112) peaks were contracted in

278

transformed SLMoS2 compared to that of pristine SLMoS2, indicating that the

279

nanostructure of SLMoS2 was affected and transformed to the 2H phase with the

280

chemical dissolution of the 1T polymorph.

281

Morphological and Structural Alterations

282

As shown in Figure 3a, the lateral size of pristine SLMoS2 was approximately

283

100~200 nm, and the thickness was 1.243±0.051 nm, which were consistent with the

284

reported properties of SLMoS2 nanosheets.29 In addition, abnormal pores and dentate

285

edges existed on and around the nanosheets, which was ascribed to inherent defects of

286

SLMoS2.30 After treatment with visible light irradiation, the Mo and S atoms on the 13

ACS Paragon Plus Environment

Page 14 of 41

Page 15 of 41

Environmental Science & Technology

287

nanosheets were chemically dissolved, and the lateral size became substantially

288

smaller (Figure 3b−d). The large nanosheets also transformed to ribbon morphology,

289

and more defects were formed on the nanosheets, especially for T11−SLMoS2. The

290

thicknesses of the T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2 nanosheets increased

291

from approximately 1.243 nm to 2.598 nm (2.533±0.065 nm, n=3), 2.742 nm

292

(2.701±0.041 nm, n=3), and 4.267 nm (4.214±0.053 nm, n=3), respectively; these

293

results were consistent with the TEM images in Figure S5. As presented in the FTIR

294

results in Figure 3e, a significant peak ascribed to Mo−S stretching vibrations located

295

at ~600 cm−1 was observed for pristine SLMoS2.26 In contrast, the intensity of the

296

Mo−S stretching bond in the transformed SLMoS2 became very weak. The peak

297

located at ~800 cm−1 in pristine SLMoS2 and associated with the Mo−O stretching

298

bond also disappeared, particularly in T11−SLMoS2. These results indicated that the

299

chemical dissolution of SLMoS2 and Mo oxides significantly altered the functional

300

bonds on the surface of SLMoS2. The Raman phonon modes associated with the

301

in−plane (E2g) and out−of−plane (A1g) lattice vibrations of pristine SLMoS2 are

302

located at ~385 cm−1 and 403 cm−1, respectively.31 A previous study reported that the

303

intensity of Raman phonon modes (E2g and A1g) was positively correlated with defects

304

in two−dimensional nanosheets.32 In Figure 3f, the peak intensity of E2g and A1g bonds

305

significantly increased in the transformed SLMoS2, attributed to ion release (Figure 1)

306

and the abundance of defects on the nanosheets, denoted by red arrows in Figure 3d.

307

Moreover, the E2g and A1g vibrations in the Raman spectra of transformed SLMoS2

308

nanosheets exhibited considerable red and blueshifts, respectively. The blueshift in 14

ACS Paragon Plus Environment

Environmental Science & Technology

309

A1g vibrations indicated that the nanosheets transitioned from single−layer to

310

bulk−layer thickness. The redshift in E2g vibrations represented the stacking of

311

nanosheets.33 Generally, the layer number of MoS2 nanosheets can be evaluated by the

312

Raman shift of A1g−E2g.34 In the present study, the Raman shift of A1g−E2g for pristine

313

SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2 was 18.0, 20.7, 22.2, and 24.1

314

cm−1, respectively. Correspondingly, the layer numbers of T3−SLMoS2, T7−SLMoS2,

315

and T11−SLMoS2 were calculated to be 2, 3 and 4, respectively (Figure S6). It has

316

been reported that the loss of colloidal stability (as presented below, Figure 3g) can

317

cause the restacking of MoS2 nanosheets.35, 36 Chemical dissolution on the surfaces

318

and edges of SLMoS2 induced the destruction of ring structure and the enrichment of

319

defects (Figure 2d), which probably led to the scrolling and bending of SLMoS2.37

320

The SLMoS2 nanosheets were restacked, and the plane structure was scrolled; thus,

321

the layer number of transformed SLMoS2 increased, as supported by AFM (Figure 3)

322

and TEM (Figure S5) images.

323

Alterations in Colloidal Stability

324

The aggregation kinetics were assessed using the initial aggregation rate (d) when the

325

hydrodynamic diameter reached 1.5−fold higher than the initial size.38 As shown in

326

Figure 3g, the average size of pristine SLMoS2 increased with an initial aggregation

327

rate of 0.487 nm/s. In contrast, transformed SLMoS2 exhibited higher instability, and

328

the initial aggregation rates were 0.963, 1.271, and 1.763 nm/s for T3−SLMoS2,

329

T7−SLMoS2, and T11−SLMoS2, respectively. At 96 h, the hydrodynamic diameters of

330

pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2 were 1228, 1833, 1956, 15

ACS Paragon Plus Environment

Page 16 of 41

Page 17 of 41

Environmental Science & Technology

331

and 2145 nm, respectively. Furthermore, the zeta potential results revealed that

332

pristine SLMoS2 was dispersive (−32.5 mV~−39.9 mV) when the pH was over 7

333

(Figure 3h) due to the electrostatic repulsion among negatively charged nanosheets.35

334

In contrast, the zeta potential decreased after chemical dissolution (Figure 3h). The

335

zeta potentials of T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2 ranged from −30.3 to

336

−36.3 mV, −28.4 to −33.2 mV, and −27.5 to −31.1 mV, respectively. Thus,

337

dissolution and the alteration of phase, morphology and structure reduced the

338

dispersity of SLMoS2 in water.

339

Photoinduced Electronic Competition and Chemical Dissolution Assisted by DO

340

The UV−vis−near−infrared (UV−vis−NIR) diffuse reflection spectrum suggested that

341

the direct band gap of pristine SLMoS2 was 1.8 eV (Figure S7a), which was similar to

342

the data in the literature.39 The available wavelength for the band gap was 690.2 nm,

343

based on the threshold of the available wavelength (nm = 1242.375/band gap).

344

Moreover, Mott−Schottky plots showed that the flat−band potential of pristine

345

SLMoS2 was −0.45 eV (Figure S7b), which was similar to that of the conduction band

346

(CB).40 The CB potential was −0.45 eV, which is sufficient to oxidize O2 to produce

347

peroxy radical (•O2−)[Eθ(O2/•O2−) = −0.16 eV versus reversible hydrogen electrode].

348

However, •O2− generation was inhibited due to the rapid electron−hole recombination

349

of SLMoS2. Without DO, the signal of DMPO−•O2− in the SLMoS2 suspension was

350

very weak (Figure 4a and 4b). One study revealed that there is competition between

351

electron−hole recombination in materials and acquisition by other reactions (such as

352

oxygen reduction and oxidation).41 The ROS intensity was significantly strengthened 16

ACS Paragon Plus Environment

Environmental Science & Technology

353

after oxygen was aerated (Figure 4c). The continuous supplementation of DO

354

competitively captured the separated electrons to produce •O2− under light irradiation

355

(equation S3), accelerating the chemical dissolution of SLMoS2 (Figure 4a−4c).

356

According to the CB potential and band gap energy, the valence band (VB) potential

357

was calculated to be 1.35 eV, which cannot directly oxidize OH− to produce •OH

358

(Eθ(OH−/•OH) = 2.40 eV versus reversible hydrogen electrode). •OH radicals were

359

notably generated under alkaline conditions (Figure 4d), probably due to the

360

acquisition of electrons from OH− by soluble Mo6+ species under visible light

361

irradiation (equation S4). In sum, the electrons from SLMoS2 were competitively

362

utilized by O2 to produce •O2−under DO treatment and visible light irradiation, and

363

•OH radical was also generated, thus leading to the chemical dissolution of SLMoS2

364

(Figure 4e).

365

Mitigation of Algal Division, Cellular Uptake, Ultrastructure and ROS

366

The effects of the above photoinduced physicochemical transformations of SLMoS2

367

on algal ecotoxicity were investigated. For algal division, the initial number (1.5×105

368

CFU/mL) of algae increased with incubation time, as shown in Figure S8. During this

369

period, the division of algal cells was significantly inhibited by exposure to SLMoS2

370

in a dose−dependent manner. In contrast, transformed SLMoS2 induced significantly

371

lower inhibition than pristine SLMoS2, except for T3−SLMoS2 (Figure S8b−d). The

372

effects of T11−SLMoS2 on cell division were the lowest, and the inhibition

373

percentages were only 10.9−15.2%. The algal cells in the control groups presented

374

intact ultrastructural morphology, including cell walls, chloroplasts, cell nuclei, and 17

ACS Paragon Plus Environment

Page 18 of 41

Page 19 of 41

Environmental Science & Technology

375

other cytoplasmic compartments (Figure 5a). Mo ions (Na2MoO4) did not induce

376

obvious alterations in cell ultrastructure (Figure 5b). In contrast, pristine SLMoS2 was

377

visible in cellular cytoplasm and chloroplasts (Figure 5c, yellow arrows). The

378

nanosheet content in the pristine SLMoS2 groups (25 mg/L) reached 2.428 μg/105

379

cells (Figure S9a), and plasmolysis (red arrows) was also significantly induced by

380

pristine SLMoS2 (Figure 5c). The plasma membrane shrank and separated from the

381

cell walls. In addition, exposure to pristine SLMoS2 damaged and blurred

382

chloroplasts, and the biosynthesis of chlorophyll a was remarkably inhibited (up to

383

58%) compared to the control (Figure S9b). However, the damage to the cellular

384

ultrastructure by transformed SLMoS2 was weak (Figure 5d–5f). The cellular uptake

385

of transformed SLMoS2 was also reduced to 2.316, 1.806, and 1.103 μg/105 cells for

386

T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2, respectively, at 25 mg/L (Figure S9a).

387

Physical and pharmacological inhibitions were used to analyze the uptake

388

mechanisms

389

energy−dependent endocytosis of algal cells could be inhibited at low temperatures.

390

The uptakes of pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2 (25

391

mg/L) by algae at 4°C were reduced by 45.63%, 48.59%, 50.26, and 47.87%,

392

respectively (Figure S10a), compared to that of the control, suggesting the energy

393

dependence of SLMoS2 internalization. Furthermore, the algal cells were treated with

394

MβCD, CPZ, and EIPA to prevent caveolae−mediated endocytosis, clathrin−mediated

395

endocytosis, and uptake by micropinocytosis, respectively. Algal uptake of pristine

396

SLMoS2 (25 mg/L) with the pretreatments of MβCD, CPZ, and EIPA was reduced by

of

pristine

and

transformed

SLMoS2

18

ACS Paragon Plus Environment

by

algal

cells.

The

Environmental Science & Technology

397

21.68%, 5.77%, and 39.67%, respectively. By contrast, the uptake inhibition effects

398

of transformed SLMoS2 by pharmacological pretreatment were considerably more

399

significant, especially for T11−SLMoS2. MβCD, CPZ, and EIPA decreased the cellular

400

contents of T11−SLMoS2 (25 mg/L) by 36.78%, 19.56%, and 48.76%, respectively

401

(Figure S10b−10d). The above data demonstrated that the uptake mechanisms of

402

pristine and transformed SLMoS2 were slightly different and that micropinocytosis

403

was the dominant pathway for the uptake of SLMoS2 nanosheets. Therefore, the

404

bioavailability and cellular contents of transformed SLMoS2 were evidently lower

405

than those of pristine SLMoS2, owing to the higher aggregation of transformed

406

SLMoS2 (Figure 3g−h) and inhibited micropinocytosis. The intracellular ROS levels

407

in pristine SLMoS2 compared to those in the control were significantly increased by

408

186.9−313.2% (Figure S9c). The ROS levels in transformed SLMoS2, especially in

409

the T11−SLMoS2 group, were lower than those in pristine SLMoS2, which was

410

consistent with the results of nanomaterial uptake and photosynthesis.

411

Metabolic Mechanisms of Phytotoxicity Mitigation

412

Metabolomic analysis provides a global view on the perturbation of a metabolism.42

413

Fifty−three metabolites were identified by analyzing approximately 250 peaks for

414

each sample using GC−MS (Tables S2 and S3). The identified metabolites included

415

fatty acids, amino acids, carbohydrates, alkanes and other biomolecules. Hierarchical

416

clustering (HCL) analysis indicated that the samples could be divided into two

417

clusters:1) pristine SLMoS2 (1, 10, and 25 mg/L) and T3−SLMoS2 (10 and 25 mg/L)

418

and 2) others (the control and most of the transformed SLMoS2). The results 19

ACS Paragon Plus Environment

Page 20 of 41

Page 21 of 41

Environmental Science & Technology

419

suggested that pristine SLMoS2 significantly perturbed the cellular metabolism of

420

algae and the perturbation of metabolism was mitigated by environmental

421

transformation (Figure S11a). Furthermore, ANOVA with Tukey's test found that

422

SLMoS2 at 10 and 25 mg/L significantly downregulated the levels of amino acids and

423

fatty acids (p1)

436

between ROS and ornithine/threonine/tyrosine/serine is observed in Figure S12,

437

implying the inhibition of amino acid syntheses and the destruction of the defense

438

system induced by ROS. Serine takes part in the biosynthesis of purines, pyrimidines

439

and chlorophyll.45 Pristine SLMoS2 reduced the serine level, which supported the

440

decrease in chlorophyll a content. Furthermore, unsaturated fatty acid metabolism has 20

ACS Paragon Plus Environment

Environmental Science & Technology

441

been correlated with cell membrane fluidity.46 The levels of unsaturated fatty acids in

442

the transformed SLMoS2 groups were significantly lower than those in the pristine

443

SLMoS2 groups (Figure S11a), which was consistent with the results of plasmolysis in

444

Figure 5. Fatty acids are vigorously responsive to extra stimulation during the growth

445

process.47, 48 Herein, unsaturated and saturated fatty acids presented positive and

446

negative relationships, respectively, with the ROS level (Figure S12). Arachidonic

447

acid is a polyunsaturated fatty acid released from membrane phospholipids in

448

response to ROS.48 The arachidonic acid levels in transformed SLMoS2 treatments

449

were significantly lower than those in pristine SLMoS2 groups, supporting the

450

negligible plasmolysis of algal cells exposed to transformed SLMoS2. Salicylic acid is

451

a signaling molecule that plays an important role in the plant defense response.49

452

Pristine SLMoS2 inhibited the levels of salicylic acid, but the inhibition was

453

remarkably diminished in transformed SLMoS2 (Figure S11a), indicating the recovery

454

of the defense system of algal cells. By metabolomic analysis, the above results

455

explore the toxicological mechanisms by which environmental transformation

456

mitigates nanotoxicity.

457

Outlook on the Environmental Transformation of SLMoS2

458

SLMoS2 has been applied in various fields and presents great potential for release into

459

the environment during its life cycle (e.g., fabrication, use and disposal). The phase,

460

morphology and dissolution (accompanied by defects) of SLMoS2 play a dominant

461

role in surface chemistry, colloidal stability and activity,50 determining the uptake and

462

ecotoxicity of this material to organisms. DO and visible light irradiation are two 21

ACS Paragon Plus Environment

Page 22 of 41

Page 23 of 41

Environmental Science & Technology

463

fundamental factors in the water environment. However, the effects of DO and visible

464

light irradiation on the physicochemical transformation and nanostructure of SLMoS2,

465

especially ion release and phase transformation, remain largely unknown. The 1T

466

phase of pristine SLMoS2 was preferentially dissolved in alkaline conditions, and then

467

a ribbon structure and abundant defects on the surfaces of nanosheets were formed.

468

Transformed SLMoS2 nanosheets were thickened into multilayer structures, and the

469

colloidal stability was reduced. The phase and single−layer structure of SLMoS2 play

470

critical roles in photocatalysis and pollutant removal.51, 52 The environmental

471

transformations presented in this work unequivocally affect photocatalysis and

472

pollutant removal when using SLMoS2. Regarding biological responses, pristine

473

SLMoS2 (1T−rich) with a negative surface charge (pH>7.0) was stable (Figure 3g−f)

474

and was more effectively absorbed by algal cells than was transformed SLMoS2

475

(2H−rich), as illustrated in Figure 5. Transformed SLMoS2 (2H−rich phase) induced

476

lower intracellular ROS levels than pristine 1T−rich SLMoS2 (Figure S9c), indicating

477

the mitigation of nanotoxicity.53 Transformed SLMoS2 also triggered a weaker

478

perturbation of metabolism (Figure S11). In other words, the environmental risks of

479

SLMoS2 would be high in low−oxygen and dark conditions. Soluble ionic species that

480

coexist with nanomaterials are more bioavailable and have been confirmed to be the

481

drivers of the toxicological response.54, 55 In the present study, 100 μM Na2MoO4

482

showed no statistically significant inhibitory effect on algal cell growth (Figure 5).

483

The synthesis of chlorophyll a and intracellular ROS was also not remarkably

484

influenced by Na2MoO4 (Figure S9b and S9c). The above results indicated different 22

ACS Paragon Plus Environment

Environmental Science & Technology

485

environmental risks than those derived from released Ag ions from Ag

486

nanoparticles.55 One study found that the formation of abundant surface defects was

487

beneficial for the antibacterial activity of nanomaterials (e.g., carbon materials and

488

metal nanoparticles) on a Gram−negative bacterium (E. coli),56 and could also

489

strengthen the toxicity of nanomaterials to organisms (RT−W1 cells and embryos).57

490

The interdependence between the defect/phase composition and ionic dissolution of

491

SLMoS2 deserves substantial attention in future applications and risk evaluations.

492 493

ASSOCIATED CONTENT

494

Supporting Information

495

Methods regarding characterization techniques; the detection of peroxy radicals,

496

hydroxyl radicals and ROS; and a theoretical analysis of the photoinduced redox

497

reaction. Tables regarding the dissolution rate of SLMoS2 and metabolites of algal

498

cells. Figures regarding the chemical dissolution of SLMoS2 nanosheets, the molar

499

ratio of S/Mo species, XPS spectra, X-ray powder diffraction spectra, TEM images,

500

the dependence of E2g−A1g Raman shifts on the number of MoS2 layers, the inhibition

501

of algal cell growth, quantifications of cellular SLMoS2, the inhibition of chlorophyll

502

a, ROS generation in algal cell, and metabolite analysis.

503 504

AUTHOR INFORMATION

505

Corresponding Author

506

*E–mail: [email protected] (X.H.). Phone: +86–022–23507800. Fax: +86– 23

ACS Paragon Plus Environment

Page 24 of 41

Page 25 of 41

507

Environmental Science & Technology

022–66229562.

508 509

NOTES

510

The authors declare that there are no competing financial interests.

511 512

ACKNOWLEDGEMENTS

513

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

514

China (grant nos. 21577070 and 21307061), the China Postdoctoral Science

515

Foundation (No. 2018M642757), and the Opening Foundation of Ministry of

516

Education Key Laboratory of Pollution Processes and Environmental Criteria (No.

517

201801). This work was also supported by the Doctoral Scientific Research

518

Foundation of Henan Normal University (No. 5101219170133).

519 520

REFERENCES

521

(1) Autere, A.; Jussila, H.; Dai, Y.; Wang, Y.; Lipsanen, H.; Sun, Z. Nonlinear

522

Optics with 2D Layered Materials. Adv. Mater. 2018, 30, 1705963.

523

(2) Yang, P.; Zou, X.; Zhang, Z.; Hong, M.; Shi, J.; Chen, S.; Shu, J.; Zhao, L.;

524

Jiang, S.; Zhou, X.; Huan, Y.; Xie, C.; Gao, P.; Chen, Q.; Zhang, Q.; Liu, Z.; Zhang,

525

Y. Batch Production of 6−Inch Uniform Monolayer Molybdenum Disulfide Catalyzed

526

by Sodium in Glass. Nat. Commun. 2018, 9, 929.

527

(3) Gupta, U.; Rao, C. N. R. Hydrogen Generation by Water Splitting using MoS2

528

and Other Transition Metal Dichalcogenides. Nano Energy 2017, 41, 49−65. 24

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 41

529

(4) Lei, Z.; Zhan, J.; Tang, L.; Zhang, Y.; Wang, Y. Recent Development of Metallic

530

(1T) Phase of Molybdenum Disulfide for Energy Conversion and Storage. Adv.

531

Energy Mater. 2018, 8, 1703482.

532

(5) Majd, S. M.; Salimi, A.; Ghasemi, F. An Ultrasensitive Detection of miRNA−155

533

in Breast Cancer via Direct Hybridization Assay using Two−Dimensional

534

Molybdenum Disulfide Field−Effect Transistor Biosensor. Biosens. Bioelectron. 2018,

535

105, 6−13.

536

(6) Chen, S. C.; Lin, C. Y.; Cheng, T. L.; Tseng, W. L. 6−Mercaptopurine−Induced

537

Fluorescence Quenching of Monolayer MoS2 Nanodots: Applications to Glutathione

538

Sensing, Cellular Imaging, and Glutathione−Stimulated Drug Delivery. Adv. Funct.

539

Mater. 2017, 27, 1702452.

540

(7) Xie, W.; Gao, Q.; Wang, D.; Guo, Z.; Gao, F.; Wang, X.; Cai, Q.; Feng, S. S.;

541

Fan, H.; Sun, X.; Zhao, L. Doxorubicin−loaded Fe3O4@MoS2−PEG−2DG Nanocubes

542

as

543

Chemo−Photothermal Therapy of Breast Cancer. Nano Res. 2018, 11, 2470−2487.

544

(8) Alam, I.; Guiney, L. M.; Hersam, M. C.; Chowdhury, I. Antifouling Properties of

545

Two−Dimensional Molybdenum Disulfide and Graphene Oxide. Environ. Sci. : Nano.

546

2018, 5, 1628−1639.

547

(9) Hirunpinyopas, W.; Prestat, E.; Worrall, S. D.; Haigh, S. J.; Dryfe, R. A. W.;

548

Bissett, M. A. Desalination and Nanofiltration through Functionalized Laminar MoS2

549

Membranes. ACS Nano 2017, 11, 11082−11090.

550

(10) Wang, Z.; von demBussche, A.; Qiu, Y.; Valentin, T. M.; Gion, K.; Kane, A. B.;

a

Theranostic

Platform

for

Magnetic

Resonance

25

ACS Paragon Plus Environment

Imaging−Guided

Page 27 of 41

Environmental Science & Technology

551

Hurt, R. H. Chemical Dissolution Pathways of MoS2 Nanosheets in Biological and

552

Environmental Media. Environ. Sci. Technol. 2016, 50, 7208−7217.

553

(11) Zou, W.; Zhou, Q.; Zhang, X.; Hu, X. Environmental Transformations and Algal

554

Toxicity of Single−Layer Molybdenum Disulfide Regulated by Humic Acid. Environ.

555

Sci. Technol. 2018, 52, 2638−2648.

556

(12) Ding, X.; Wang, J.; Rui, Q.; Wang, D. Long−Term Exposure to Thiolated

557

Graphene Oxide in the Range of μg/L Induces Toxicity in Nematode Caenorhabditis

558

Elegans. Sci. Total Environ. 2018, 616, 29−37.

559

(13) Hu, X.; Zhou, M.; Zhou, Q. Ambient Water and Visible−Light Irradiation Drive

560

Changes in Graphene Morphology, Structure, Surface Chemistry, Aggregation, and

561

Toxicity. Environ. Sci. Technol. 2015, 49, 3410−3418.

562

(14) Mansano, A. S.; Souza, J. P.; Cancino−Bernardi, J.; Venturini, F. P.; Marangoni,

563

V. S.; Zucolotto, V. Toxicity of Copper Oxide Nanoparticles to Neotropical Species

564

Ceriodaphniasilvestrii and Hyphessobrycon Eques. Environ. Pollut. 2018, 243,

565

723−733.

566

(15) Marchioni, M.; Gallon, T.; Worms, I.; Jouneau, P.−H.; Lebrun, C.; Veronesi, G.;

567

Truffier−Boutry, D.; Mintz, E.; Delangle, P.; Deniaud, A.; Michaud−Soret, I. Insights

568

into Polythiol−Assisted AgNP Dissolution Induced by Bio−Relevant Molecules.

569

Environ. Sci. :Nano 2018, 5, 1911−1920.

570

(16) Yamamoto, M.; Einstein, T. L.; Fuhrer, M. S.; Cullen, W. G. Anisotropic

571

Etching of Atomically Thin MoS2. J. Phys. Chem. C 2013, 117, 25643−25649.

572

(17) Wang, Z.; Zhu, W.; Qiu, Y.; Yi, X.; von demBussche, A.; Kane, A.; Gao, H.; 26

ACS Paragon Plus Environment

Environmental Science & Technology

573

Koski, K.; Hurt, R. Biological and Environmental Interactions of Emerging

574

Two−Dimensional Nanomaterials. Chem. Soc. Rev. 2016, 45, 1750−1780.

575

(18) Qu, X.; Alvarez, P. J. J.; Li, Q. Photochemical Transformation of Carboxylated

576

Multiwalled Carbon Nanotubes: Role of Reactive Oxygen Species. Environ. Sci.

577

Technol. 2013, 47, 14080−14088.

578

(19) Wang, D.; Liu, L.; Zhang, F.; Tao, K.; Pippel, E.; Domen, K. Spontaneous Phase

579

and Morphology Transformations of Anodized Titania Nanotubes Induced by Water

580

at Room Temperature. Nano Lett. 2011, 11, 3649−3655.

581

(20) Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P. Y.; Mak, K. F.; Kim, C.−J.;

582

Muller, D.; Park, J. High−Mobility Three−Atom−Thick Semiconducting Films with

583

Wafer−Scale Homogeneity. Nature 2015, 520, 656−660.

584

(21) Sharifi, S.; Behzadi, S.; Laurent, S.; Forrest, M. L.; Stroeve, P.; Mahmoudi, M.

585

Toxicity of Nanomaterials. Chem. Soc. Rev. 2012, 41, 2323−2343.

586

(22) Wang, W.; Sedykh, A.; Sun, H.; Zhao, L.; Russo, D. P.; Zhou, H.; Yan, B.; Zhu,

587

H. Predicting Nano−Bio Interactions by Integrating Nanoparticle Libraries and

588

Quantitative Nanostructure Activity Relationship Modeling. ACS Nano 2017, 11,

589

12641−12649.

590

(23) Braydich−Stolle, L. K.; Schaeublin, N. M.; Murdock, R. C.; Jiang, J.; Biswas, P.;

591

Schlager, J. J.; Hussain, S. M. Crystal Structure Mediates Mode of Cell Death in TiO2

592

Nanotoxicity. J. Nanopart. Res. 2009, 11, 1361−1374.

593

(24) Chng, E. L. K.; Sofer, Z.; Pumera, M. MoS2Exhibits Stronger Toxicity with

594

Increased Exfoliation. Nanoscale 2014, 6, 14412−14418. 27

ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41

Environmental Science & Technology

595

(25) Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M.

596

Photoluminescence from Chemically Exfoliated MoS2. Nano Lett. 2011, 11,

597

5111−5116.

598

(26) Wang, Y.; Ou, J. Z.; Chrimes, A. F.; Carey, B. J.; Daeneke, T.; Alsaif, M. M. Y.

599

A.; Mortazavi, M.; Zhuiykov, S.; Medhekar, N.; Bhaskaran, M.; Friend, J. R.; Strano,

600

M. S.; Kalantar−Zadeh, K. Plasmon Resonances of Highly Doped Two−Dimensional

601

MoS2. Nano Lett. 2015, 15, 883−890.

602

(27) Djamil, J.; Hansen, A. L.; Backes, C.; Bensch, W.; Schurmann, U.; Kienle, L.;

603

Duvel, A.; Heitjans, P. Using Light, X-rays and Electrons for Evaluation of the

604

Nanostructure of Layered Materials. Nanoscale 2018, 10, 21142−21150.

605

(28) Liu, Q.; Li, X.; He, Q.; Khalil, A.; Liu, D.; Xiang, T.; Wu, X.; Song, L.

606

Gram−Scale Aqueous Synthesis of Stable Few−Layered 1T−MoS2: Applications for

607

Visible−Light−Driven Photocatalytic Hydrogen Evolution. Small 2015, 11,

608

5556−5564.

609

(29) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.−J.; Loh, K. P.; Zhang, H. The

610

Chemistry

611

Nanosheets. Nat. Chem. 2013, 5, 263−275.

612

(30) Hong, J.; Hu, Z.; Probert, M.; Li, K.; Lv, D.; Yang, X.; Gu, L.; Mao, N.; Feng, Q.;

613

Xie, L.; Zhang, J.; Wu, D.; Zhang, Z.; Jin, C.; Ji, W.; Zhang, X.; Yuan, J.; Zhang, Z.

614

Exploring Atomic Defects in Molybdenum Disulphide Monolayers. Nat. Commun.

615

2015, 6, 6293.

616

(31) Lee, C.; Yan, H.; Brus, L. E.; Heinz, T. F.; Hone, J.; Ryu, S. Anomalous Lattice

of

Two−Dimensional

Layered

Transition

28

ACS Paragon Plus Environment

Metal

Dichalcogenide

Environmental Science & Technology

617

Vibrations of Single− and Few−Layer MoS2. ACS Nano 2010, 4, 2695−2700.

618

(32) Mignuzzi, S.; Pollard, A. J.; Bonini, N.; Brennan, B.; Gilmore, I. S.; Pimenta, M.

619

A.; Richards, D.; Roy, D. Effect of Disorder on Raman Scattering of Single−Layer

620

MoS2. Phys. Rev. B 2015, 91, 195411.

621

(33) Kretschmer, S.; Komsa, H.−P.;Boggild, P.; Krasheninnikov, A. V. Structural

622

Transformations in Two−Dimensional Transition−Metal Dichalcogenide MoS2 under

623

an Electron Beam: Insights from First−Principles Calculations. J. Phys. Chem. Lett.

624

2017, 8, 3061−3067.

625

(34) Velicky, M.; Bissett, M. A.; Woods, C. R.; Toth, P. S.; Georgiou, T.; Kinloch, I.

626

A.; Novoselov, K. S.; Dryfe, R. A. W. Photoelectrochemistry of Pristine Mono− and

627

Few−Layer MoS2. Nano Lett. 2016, 16, 2023−2032.

628

(35) Chou, S. S.; De, M.; Kim, J.; Byun, S.; Dykstra, C.; Yu, J.; Huang, J. X.; Dravid,

629

V. P. Ligand Conjugation of Chemically Exfoliated MoS2. J. Am. Chem. Soc. 2013,

630

135, 4584−4587.

631

(36) Liu, W.; Yang, X.; Zhang, Y.; Xu, M.; Chen, H. Ultra−Stable Two−Dimensional

632

MoS2 Solution for Highly Efficient Organic Solar Cells. RSC Adv. 2014, 4,

633

32744−32748.

634

(37) Li, Y.; Yang, N.; Du, T.; Wang, X.; Chen, W. Transformation of Graphene

635

Oxide by Chlorination and Chloramination: Implications for Environmental Transport

636

and Fate. Water Res. 2016, 103, 416−423.

637

(38) Bouchard, D.; Zhang, W.; Powell, T.; Rattanaudompol, U. S. Aggregation

638

Kinetics and Transport of Single−Walled Carbon Nanotubes at Low Surfactant 29

ACS Paragon Plus Environment

Page 30 of 41

Page 31 of 41

Environmental Science & Technology

639

Concentrations. Environ. Sci. Technol. 2012, 46, 4458−4465.

640

(39) Liu, C.; Kong, D.; Hsu, P.C.; Yuan, H.; Lee, H.W.; Liu, Y.; Wang, H.; Wang, S.;

641

Yan, K.; Lin, D.; Maraccini, P. A.; Parker, K. M.; Boehm, A. B.; Cui, Y. Rapid Water

642

Disinfection using Vertically Aligned MoS2 Nanofilms and Visible Light. Nat.

643

Nanotechnol. 2016, 11, 1098−1104.

644

(40) Zhang, M.; Bai, X.; Liu, D.; Wang, J.; Zhu, Y. Enhanced Catalytic Activity of

645

Potassium−Doped Graphitic Carbon Nitride Induced by Lower Valence Position.

646

Appl. Catal., B 2015, 164, 77−81.

647

(41) Tong, H.; Ouyang, S. X.; Bi, Y. P.; Umezawa, N.; Oshikiri, M.; Ye, J. H.

648

Nano−Photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24,

649

229−251.

650

(42) Gioria, S.; Vicente, J. L.; Barboro, P.; La Spina, R.; Tomasi, G.; Urban, P.;

651

Kinsner−Ovaskainen, A.; Francois, R.; Chassaigne, H. A Combined Proteomics and

652

Metabolomics Approach to Assess the Effects of Gold Nanoparticles in Vitro.

653

Nanotoxicology 2016, 10, 736−748.

654

(43) Zhang, W.; Tan, N. G. J.; Fu, B.; Li, S. F. Y. Metallomics and NMR−based

655

Metabolomics of Chlorella sp. Reveal the Synergistic Role of Copper and Cadmium

656

in Multi−Metal Toxicity and Oxidative Stress. Metallomics 2015, 7, 426−438.

657

(44) Tan, W.; Peralta−Videa, J. R.; Gardea−Torresdey, J. L. Interaction of Titanium

658

Dioxide Nanoparticles with Soil Components and Plants: Current Knowledge and

659

Future Research Needs−A Critical Review. Environ. Sci. : Nano 2018, 5, 257−278.

660

(45) Toubiana, D.; Batushansky, A.; Tzfadia, O.; Scossa, F.; Khan, A.; Barak, S.; 30

ACS Paragon Plus Environment

Environmental Science & Technology

661

Zamir, D.; Fernie, A. R.; Nikoloski, Z.; Fait, A. Combined Correlation−Based

662

Network and mQTL Analyses Efficiently Identified Loci for Branched−Chain Amino

663

Acid, Serine to Threonine, and Proline Metabolism in Tomato Seeds. Plant J. 2015,

664

81, 121−133.

665

(46) Kumar, M.; Bijo, A. J.; Baghel, R. S.; Reddy, C. R. K.; Jha, B. Selenium and

666

Spermine Alleviate Cadmium Induced Toxicity in the Red Seaweed Gracilaria Dura

667

by Regulating Antioxidants and DNA Methylation. Plant Physiol. Biochem. 2012, 51,

668

129−138.

669

(47) Akhavan, O. Graphene Scaffolds in Progressive Nanotechnology/Stem

670

Cell−based Tissue Engineering of the Nervous System. J. Mater. Chem. B 2016, 4,

671

3169−3190.

672

(48) Martinez, J.; Moreno, J. J. Role of Ca2+−Independent Phospholipase A2 on

673

Arachidonic Acid Release Induced by Reactive Oxygen Species. Arch. Biochem.

674

Biophys. 2001, 392, 257−62.

675

(49) Chen, Z.; Silva, H.; Klessig, D. F. Active Oxygen Species in the Induction of

676

Plant Systemic Acquired Resistance by Salicylic Acid. Science 1993, 262,

677

1883−1886.

678

(50) Kwak, I. H.; Kwon, I. S.; Abbas, H. G.; Jung, G.; Lee, Y.; Debela, T. T.; Yoo, S.

679

J.; Kim, J.−G.; Park, J.; Kang, H. S. Nitrogen−rich 1T−MoS2Layered Nanostructures

680

using Alkyl Amines for High Catalytic Performance toward Hydrogen Evolution.

681

Nanoscale 2018, 10, 14726−14735.

682

(51) Ghim, D.; Jiang, Q.; Cao, S. S.; Singamaneni, S.; Jun, Y. S. Mechanically 31

ACS Paragon Plus Environment

Page 32 of 41

Page 33 of 41

Environmental Science & Technology

683

Interlocked 1T/2H Phases of MoS2 Nanosheets for Solar Thermal Water Purification.

684

Nano Energy 2018, 53, 949−957.

685

(52) Wang, J.; Wang, N.; Guo, Y.; Yang, J.; Wang, J.; Wang, F.; Sun, J.; Xu, H.; Liu,

686

Z.; Jiang, R. Metallic−Phase MoS2 Nanopetals with Enhanced Electrocatalytic

687

Activity for Hydrogen Evolution. ACS Sustain. Chem. Eng. 2018, 6, 13435−13442.

688

(53) Pieper, H.; Chercheja, S.; Eigler, S.; Halbig, C. E.; Filipovic, M. R.; Mokhir, A.

689

Endoperoxides Revealed as Origin of the Toxicity of Graphene Oxide. Angew. Chem.,

690

Int. Ed. 2016, 55, 405−407.

691

(54) van Leeuwen, H. P.; Duval, J. F. L.; Pinheiro, J. P.; Blust, R.; Town, R. M.

692

Chemodynamics and Bioavailability of Metal Ion Complexes with Nanoparticles in

693

Aqueous Media. Environ. Sci. : Nano 2017, 4, 2108−2133.

694

(55) Bollyn, J.; Willaert, B.; Kerre, B.; Moens, C.; Arijs, K.; Mertens, J.; Leverett, D.;

695

Oorts, K.; Smolders, E. Transformation−Dissolution Reactions Partially Explain

696

Adverse Effects of Metallic Silver Nanoparticles to Soil Nitrification in Different

697

Soils. Environ. Toxicol. Chem. 2018, 37, 2123−2131.

698

(56) Perreault, F.; De Faria, A. F.; Nejati, S, Elimelech, M. Antimicrobial Properties

699

of Graphene Oxide Nanosheets: Why Size Matters. ACS Nano, 2015, 9, 7226−7236.

700

(57) George, S.; Lin, S.; Ji, Z.; Thomas, C. R.; Li, L.; Mecklenburg, M.; Meng, H.;

701

Wang, X.; Zhang, H.; Xia, T.; Hohman, J. N.; Lin, S.; Zink, J. I.; Weiss, P. S.; Nel, A.

702

E. Surface Defects on Plate−Shaped Silver Nanoparticles Contribute to its Hazard

703

Potential in a Fish Gill Cell Line and Zebrafish Embryos. ACS Nano, 2012, 6,

704

3745−3759. 32

ACS Paragon Plus Environment

Environmental Science & Technology

705 706 707

Figure Legends

708

Figure 1. Kinetics of pH−dependent chemical dissolution. (a−b) The contents of

709

dissolved Mo and S species from SLMoS2 nanosheets without dissolved oxygen (DO)

710

and light irradiation. (c−d) The contents of Mo and S species released by SLMoS2

711

nanosheets treated with DO but without light irradiation. (e−f) The contents of

712

dissolved Mo and S species from SLMoS2 nanosheets treated with saturated DO and

713

light irradiation. The saturated concentration of DO was 9.52 mg/L. The values in

714

parentheses represent the final concentrations of released ions (mg/L) on the 56th day.

715 716

Figure 2. Effect of pH on the ion release and phase composition of SLMoS2. (a)

717

UV−vis spectra of pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2. (b)

718

XPS spectra of pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2. The

719

changed intensity in phase−associated bonds (a) and the deconvolution analysis (b)

720

confirmed the alterations comprising the disappearance of the 1T phase and the

721

enrichment of the 2H phase proportion under visible light irradiation.

722 723

Figure 3. The alterations in morphology, surface chemistry, layer structure, and

724

colloidal stability of SLMoS2 driven by dissolved oxygen and visible light irradiation.

725

(a~d) AFM images of pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2.

726

(e) FTIR spectra. (f) Raman spectra. The shifts of the E2g and A1g modes are denoted 33

ACS Paragon Plus Environment

Page 34 of 41

Page 35 of 41

Environmental Science & Technology

727

by arrows. Correlations between the A1g−E2g shifts and the layer number of MoS2

728

nanosheets are presented in the SI. (g) Time profile of the average hydrodynamic

729

diameter of SLMoS2 at pH 7.0. d1, d2, d3, and d4 represent the initial aggregation rates

730

of pristine SLMoS2, T3−SLMoS2, T7−SLMoS2, and T11−SLMoS2, respectively. (h)

731

Zeta potential.

732 733

Figure 4. Generation of free redials and photochemical reaction mechanisms. (a–c)

734

the production of peroxy radicals (•O2−) by pristine SLMoS2 with or without

735

dissolved oxygen (DO), (d) the generation of hydroxyl radicals (•OH) by pristine and

736

transformed SLMoS2, and (e) the photochemical reaction mechanisms of SLMoS2.

737 738

Figure 5. Ultrastructure of algal cells and internalization of SLMoS2 as shown by

739

TEM images. (a) The control without nanomaterial exposure, (b) 100 μM Na2MoO4

740

exposure, (c) 25 mg/L pristine SLMoS2, (d) 25 mg/L T3−SLMoS2, (e) 25 mg/L

741

T7−SLMoS2, and (f) 25 mg/L T11−SLMoS2. The yellow and red arrows indicate

742

SLMoS2 and plasmolysis, respectively. Cw, cell wall; S, starch grain; Chl,

743

chloroplast; Cn, cell nucleus.

744 745

Figure 6. Pathways of the metabolic responses and amino acid and fatty acid

746

disturbances in algal cells treated with pristine and transformed SLMoS2. The

747

metabolic pathways were established based on the Kyoto Encyclopedia of Genes and

748

Genomes (KEGG) database. Red and green words represent down− and upregulated 34

ACS Paragon Plus Environment

Environmental Science & Technology

749

metabolites induced by pristine SLMoS2, respectively, compared to those of the

750

control. The metabolites labeled with up and down purple arrows are the metabolites

751

that are upregulated and downregulated, respectively, by transformed SLMoS2

752

compared to pristine SLMoS2.

753 754

755 756

Figure 1.

757

35

ACS Paragon Plus Environment

Page 36 of 41

Page 37 of 41

Environmental Science & Technology

758 759

Figure 2.

36

ACS Paragon Plus Environment

Environmental Science & Technology

760 761

Figure 3.

37

ACS Paragon Plus Environment

Page 38 of 41

Page 39 of 41

Environmental Science & Technology

762 763

Figure 4.

764 765 38

ACS Paragon Plus Environment

Environmental Science & Technology

766 767

Figure 5.

768 769 770 771 772 773 774

39

ACS Paragon Plus Environment

Page 40 of 41

Page 41 of 41

Environmental Science & Technology

Glycine

Glucose

Serine

Glyceric acid-3-phosphate

Phenylalanine Tyrosine

ROS

Pyruvate

Valine

Plasmolysis

Isoleucine Metabolic disturbance

Chl

Phosphoenolpyruvic acid

Acetyl-CoA

Arginine

Threonine Isoleucine

Oxaloacetate

Critrate

Malic acid

Isocitric acid

Fumarate

Aconitic acid

Succinate

775 776

Fatty acid metabolism 9,12-Octadecadienoic acid 9-Hexadecenoic acid Eicosatrienoic acid Octadecenoic acid Arachidonic acid Stearic acid

Asparate Methionine

Nucleus

Lactic acid

leucine Asparagine

Alanine

Salicylic acid

2-Oxoglutarate Butanoic acid

Figure 6.

40

ACS Paragon Plus Environment

Dodecanoic acid Palmitic Acid Ornithine

Urea

Glutamate

Glutamine

Glutathione

5-Oxoproline