The Phases of WS2 Nanosheets Influence Uptake, Oxidative Stress

Oct 24, 2018 - Self-Damaging Aerobic Reduction of Graphene Oxide by Escherichia coli: Role of GO-Mediated Extracellular Superoxide Formation...
0 downloads 0 Views 2MB Size
Subscriber access provided by The University of Texas at El Paso (UTEP)

Ecotoxicology and Human Environmental Health 2

The Phases of WS Nanosheets Influence Uptake, Oxidative Stress, Lipid Peroxidation, Membrane Damage and Metabolism in Algae Peng Yuan, Qixing Zhou, and Xiangang Hu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b04444 • Publication Date (Web): 24 Oct 2018 Downloaded from http://pubs.acs.org on October 25, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a 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 44

Environmental Science & Technology

1

The Phases of WS2 Nanosheets Influence Uptake, Oxidative Stress, Lipid

2

Peroxidation, Membrane Damage and Metabolism in Algae

3 4

Peng Yuan, Qixing Zhou, Xiangang Hu*

5

Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of

6

Education)/Tianjin Key Laboratory of Environmental Remediation and Pollution Control,

7

College of Environmental Science and Engineering, Nankai University, Tianjin 300350,

8

China

9 10

Corresponding author: Xiangang Hu, [email protected]

11

Fax: 0086-022-23507800

12

Tel.: 0086-022-23507800

13 14

ABSTRACT:

15

Applications of transition metal dichalcogenide (TMDC) nanosheets with different

16

phases have attracted much attention in various fields. However, the effects of TMDC

17

phases on environmental biology remain largely unknown. In this study, chemically

18

exfoliated WS2 nanosheets (Ce-WS2, mainly the 1T phase) and annealed exfoliated WS2

19

nanosheets (Ae-WS2, 2H phase) were fabricated to serve as representative TMDC

20

nanomaterials. Ce-WS2 showed higher levels of cellular uptake, oxidative stress, lipid 1

ACS Paragon Plus Environment

Environmental Science & Technology

21

peroxidation, membrane damage and inhibition of photosynthesis than Ae-WS2 in

22

Chlorella vulgaris. These differences were attributed to the higher electron conductivity

23

and higher separation efficiency of electrons and holes in the 1T phase, a typical feature

24

of Ce-WS2. Correspondingly, 2H-phase Ae-WS2 exhibited lower

25

photooxidation/reduction activity and a lower ability to generate reactive oxygen species

26

(mainly •OH) under visible-light irradiation. 1T-phase Ce-WS2 dissolved more readily

27

than Ae-WS2 and released more W ions into aqueous environments, but the W ions

28

exhibited negligible toxicity. Metabolomic analysis revealed that Ce-WS2 induced more

29

obvious alterations in metabolites (e.g., amino acids and fatty acids) and metabolic

30

pathways (e.g., starch and sucrose metabolism) than Ae-WS2. These alterations correlated

31

with cell membrane damage, oxidative stress and photosynthesis inhibition. The present

32

work provides insights into the environmentally friendly design of two-dimensional

33

TMDCs.

34 35

KEYWORDS:

36

Phytotoxicity, nanotoxicology, oxidative stress, phase, Chlorella vulgaris

37 38 39

INTRODUCTION

40

Two-dimensional (2D) transition metal dichalcogenide (TMDC) nanosheets have shown 2

ACS Paragon Plus Environment

Page 2 of 44

Page 3 of 44

Environmental Science & Technology

41

tremendous potential for use in various fields (e.g., electronic instruments, human

42

healthcare, energy storage and conversion, sensors and environmental protection).1-6

43

TMDC materials usually exist in more than one phase, and phase is a critical factor in

44

their fabrication and application.7 The structure of TDMCs, consisting of transition metal

45

atoms sandwiched between chalcogen atoms, can be either in octahedral coordination (1T

46

phase) or in trigonal prismatic coordination (2H phase), creating subtle differences in

47

electronic structures and other properties.1, 8 For example, the 1T phase is metallic, in

48

sharp contrast to the semiconducting 2H phase.2, 9 The phases of TMDCs affect the

49

surface activity of nanomaterials, and their roles in environmental applications and

50

behaviors have been studied.2, 10, 11 However, the effects of different phases of TMDC

51

nanosheets on environmental biology remain largely unknown, although the relevant

52

information is critical for evaluating the ecological risks and design of TMDCs.

53

Nanomaterials are exposed to a variety of environments during different parts of

54

their life cycles, such as fabrication, use and disposal.12, 13 Given the same elemental

55

composition, nanoscale forms with high surface activity usually induce more obvious

56

toxicity than bulk forms.14-16 It is well known that the composition, lateral size, thickness

57

and surface functionalization of nanomaterials play key roles in nanotoxicology.17-20

58

Phases determine the atomic arrangement of the nanomaterial surface and influence the

59

interface interactions between nanomaterials and the surrounding microenvironment.21

60

However, the effects of nanomaterial phases on organisms in the environment, especially 3

ACS Paragon Plus Environment

Environmental Science & Technology

61

aquatic organisms, are poorly studied. In the present study, WS2 nanosheets, a

62

representative 2D TMDC,21 were chosen to study the roles of phase in biological

63

responses to nanomaterials. In the present work, chemically exfoliated WS2 nanosheets

64

(Ce-WS2, mainly 1T phase) and annealed exfoliated WS2 nanosheets (Ae-WS2, 2H

65

phase) were fabricated and characterized prior to the biological experiments. Chlorella

66

vulgaris is widely used as a model organism to test the toxic effects of heavy metals,

67

organic pollutants and nanoparticles in aquatic environments.22 In the present study, the

68

effects of different phases of WS2 nanosheets on the biological responses of Chlorella

69

vulgaris, such as nanomaterial uptake, membrane damage, lipid peroxidation,

70

photosynthesis and toxicological mechanisms, were investigated.

71

In contrast to genes and proteins, metabolites are the direct output of biochemical

72

activity and are easily associated with biological phenomena and cellular biochemistry.23

73

However, studies examining only a few metabolites may overlook key information on

74

biological responses. Metabolomics with an untargeted analysis strategy can provide

75

global views of biological responses and identify the specific molecular mechanisms of

76

adverse effects.23, 24 Thus, in this study, the metabolic mechanisms of biological

77

responses are investigated using metabolomics. The present work provides insights into

78

the roles of nanomaterial phases, in particular for TMDCs, and their environmental risks,

79

which could be a useful reference for the evaluation of nanotoxicity and nanomaterial

80

design. 4

ACS Paragon Plus Environment

Page 4 of 44

Page 5 of 44

Environmental Science & Technology

81 82

MATERIALS AND METHODS

83

Preparation and Characterization of WS2 Nanosheets

84

Chemically exfoliated WS2 nanosheets (Ce-WS2) prepared by lithium intercalation were

85

obtained from Nanjing XFNANO Materials Tech Co., Ltd (Nanjing, China). Annealed

86

exfoliated WS2 nanosheets (Ae-WS2) were fabricated by annealing the Ce-WS2 samples

87

at 350℃ for 3 h under an argon atmosphere and then cooled to room temperature. To

88

reduce nanosheet restacking, WS2 nanosheet suspension was performed in an ultrasonic

89

dispersion (150 W for 10 min in an ice-water bath) prior to the characterization and

90

toxicology experiments. Detailed characterizations of Ce-WS2 and Ae-WS2 are provided

91

in the Supporting Information.

92

Algal Culture and Growth Inhibition by WS2 Nanosheets

93

Chlorella vulgaris (No. FACHB-8) was obtained from the Freshwater Algae Culture

94

Collection at the Institute of Hydrobiology, Wuhan, China. The algae were cultured in an

95

illumination incubator (14:10 h light:dark, 1800 lx, 24 ± 1 °C, 80% humidity). For the

96

WS2 nanosheets exposure experiment, algae in the logarithmic growth phase were

97

collected, washed, and then diluted to an initial concentration of 5 × 105 cells/mL in

98

250-mL glass flasks containing 100 mL of blue-green (BG-11) medium. To explore the

99

toxic mechanisms of nanomaterial phases and compare them with the reported results for

100

other 2D nanomaterials,15, 16, 25 we tested WS2 nanosheets (Ce-WS2 and Ae-WS2) at 0.1, 5

ACS Paragon Plus Environment

Environmental Science & Technology

101

1, and 10 mg/L by applying them to algal cells. Algae without nanomaterial exposure

102

were used as a control. To study the toxic effects of tungsten ions on Chlorella vulgaris,

103

we also exposed the algae to different concentrations (up to 50 μM) of tungstate salt

104

(Na2WO4). The Chlorella vulgaris was cultured from an initial concentration of 5 × 105

105

cells/mL, reached the stationary phase at 96 h and then gradually died afterward.26, 27

106

Therefore, the inhibitory effects of WS2 nanosheets and tungstate salt on algae growth

107

were evaluated by counting the cell numbers (Nt) using flow cytometry (FCM, Accuri

108

C6, Becton-Dickinson, USA) at 24 h, 48 h, 72 h and 96 h. The algal cells in the control

109

group (Nc) were also counted. The growth inhibition (%) = (Nc – Nt)/Nc × 100.

110

Electron Microscopic Observations

111

The surface alteration and cellular ultrastructure of algal cells were observed by scanning

112

electron microscopy (SEM) and transmission electron microscopy (TEM), respectively.

113

The details are presented in the Supporting Information.

114

Analysis of Photosynthesis

115

The chlorophyll fluorescence parameters of Chlorella vulgaris were measured every 24 h

116

using a pulse-amplitude-modulation fluorimeter (WATER-PAM, Heinz Walz, Effeltrich,

117

Germany).28 The maximum photochemical quantum yield of the photosystem (Fv/Fm)

118

was selected as an indicator of the photosynthetic capacity of Chlorella vulgaris. Algal

119

cells were dark-adapted for 10 min. The minimum fluorescence (F0) was measured using

120

a modulated light with low intensity to avoid the reduction of the PSII primary electron 6

ACS Paragon Plus Environment

Page 6 of 44

Page 7 of 44

Environmental Science & Technology

121

acceptor. The maximum fluorescence (Fm) was induced with a short saturating pulse of

122

white light. The Fv/Fm was calculated as (Fm – F0)/Fm.

123

Assessment of Oxidative Stress

124

The level of intracellular reactive oxygen species (ROS) was detected using a fluorescent

125

probe, 2′,7′-dichlorofluorescein diacetate (DCFH-DA). The intracellular total reduced

126

glutathione (GSH) content was determined using a kit (catalog number: A006-2) from

127

Nanjing Jiancheng Bioengineering Institute, China. Abiotic GSH oxidation by WS2

128

nanosheets was quantified using Ellman's assay. The details are presented in the

129

Supporting Information.

130

Lipid Peroxidation of the Cell Membrane

131

The lipid peroxidation of the cell membrane was determined by analyzing the content of

132

malondialdehyde (MDA). The cell collection, disruption and centrifugation steps were

133

the same as for the assessment of oxidative stress. The supernatants were used to analyze

134

the MDA content using a kit (Catalog Number: A003-1) from Nanjing Jiancheng

135

Bioengineering Institute, China, as directed in the manual.

136

Hemolysis Assay

137

Red blood cells (RBCs) have no fluid phase or receptor-mediated endocytosis and are

138

widely used to study nanomaterial interactions with the cell membrane29. WS2-induced

139

membrane damage was further investigated using a hemolysis assay in RBCs. Mouse

140

RBCs were diluted to 1 × 108 cells/mL in phosphate-buffered saline (PBS). 7

ACS Paragon Plus Environment

Environmental Science & Technology

141

Subsequently, 490 μL of RBC suspension was mixed with 10 μL of WS2 nanosheets to

142

achieve final concentrations of 0.1-0 mg/L. PBS and Triton X-100 were used as negative

143

and positive controls, respectively. The mixtures were gently stirred and incubated for 3 h

144

at 37 °C. The samples were centrifuged at 1000 g, and the absorbance of the supernatants

145

was measured at 541 nm on a microplate reader (Tecan Spark 10M, Switzerland). The %

146

hemolysis = (sample absorbance – negative control absorbance)/(positive control

147

absorbance – negative control absorbance) × 100.

148

Permeability of Cell Membranes

149

Algal cells were collected by centrifugation at 1500g for 5 min after treatment with WS2

150

nanosheets or tungstate salt for 96 h and then washed and resuspended in PBS. To assess

151

the permeabilization of algae, we used the fluorescent dye propidium iodide (PI) at 1

152

μg/L to stain cellular nucleic acids. The samples were incubated for 5 min in the dark at

153

room temperature. The fluorescence emission was measured by FCM.

154

Comet assay

155

After exposure to WS2 nanosheets for 96 h, algal cells were collected. DNA damage was

156

assessed using a comet assay kit (Catalog number, KGA240-50, Keygen Biotech, China)

157

according to the manufacturer’s specifications. The details are presented in the

158

Supporting Information.

159

Electron Spin Resonance

160

Electron spin resonance (ESR) measurements were carried out at ambient temperature 8

ACS Paragon Plus Environment

Page 8 of 44

Page 9 of 44

Environmental Science & Technology

161

using a Magnettech MiniScope 400 ESR spectrometer operated at a microwave

162

frequency of 9.4 GHz and a magnetic field modulation frequency of 100 kHz. A solar

163

simulator consisting of a 300 W xenon lamp filtered to provide simulated sunlight was

164

used in the ESR studies. The superoxide and hydroxyl radicals were determined using the

165

spin trap 5,5-dimethyl-1-pyrroline-N-oxide (DMPO).30 2,2,6,6-Tetramethylpiperidine

166

(TEMP) was used to demonstrate the generation of singlet oxygen.30 The presence of free

167

electrons was detected using the spin label 2,2,6,6-tetramethylpiperidine-1-oxyl

168

(TEMPO).30

169

Dissolution of WS2 Nanosheets

170

WS2 nanosheets were prepared at 10 mg/L in deionized water and, separately, in BG-11

171

medium. The dissolved ions from 10 mg/L WS2 nanosheets were separated by

172

centrifuging 10 mL samples at 5000 g in a centrifugal ultrafilter (Amicon Ultra-15 3 kD,

173

Millipore, MA) for 30 min after different experimental durations (3 h, 24 h, 48 h, 72 h

174

and 96 h). The concentrations of dissolved W ions were directly determined using

175

inductively coupled plasma mass spectrometry (ICP-MS, Elan drc-e, PerkinElmer, USA).

176

Cellular Uptake

177

After 96 h of exposure, algal cells were washed repeatedly with PBS, then collected and

178

digested using HNO3/H2O2 (3:1) until no color was observed. After the cell digest was

179

filtered through a 0.22-µm water membrane, the concentrations of W ions were measured

180

by ICP-MS. Physical and pharmacological inhibitors were used to further investigate the 9

ACS Paragon Plus Environment

Environmental Science & Technology

181

pathway of nanosheet uptake by algal cells. The cells were precooled at 4 °C for 1 h to

182

inhibit energy-dependent uptake, and then WS2 nanosheets were added and incubated

183

with the algae for another 1 h. To investigate the specific mechanism of endocytosis, the

184

cells were pretreated with methyl-beta-cyclodextrin (MβCD, 20 mM), chlorpromazine

185

hydrochloride (CPZ, 100 μM), or 5-(N-ethyl-N-isopropyl) amiloride (EIPA, 50 μM) for 1

186

h. Subsequently, WS2 nanosheets were added and coincubated with the algae for 1 h. The

187

internalized WS2 nanosheets were quantified using ICP-MS.

188

Metabolomic Analysis

189

After exposure to WS2 nanosheets for 96 h, the algal cells were collected. Intracellular

190

metabolites were extracted from cells by liquid–liquid extraction and subjected to

191

nitrogen blow-off, lyophilization, derivatization, and subsequent analysis by gas

192

chromatography−mass spectrometry (GC-MS, 6890N/5973, Agilent, USA).22 The details

193

are presented in the Supporting Information.

194

Statistical Analysis

195

All treatments included three replicates, and the results are presented as the mean ±

196

standard deviation. The data were analyzed using one-way analysis of variance

197

(ANOVA) and compared using Tukey’s test with SPSS 22.0 software. A p value less

198

than 0.05 was considered statistically significant. The metabolomic data were imported

199

into SPSS 22.0 software for univariate analysis. Metabolic pathway analysis was

200

performed using MetaboAnalyst version 4.0 (http://www.metaboanalyst.ca), and 10

ACS Paragon Plus Environment

Page 10 of 44

Page 11 of 44

Environmental Science & Technology

201

SIMCA-P 13.0 software was used for multivariate analysis of metabolites.

202

RESULTS AND DISCUSSION

203

Characterization of Ce-WS2 and Ae-WS2

204

As shown in Figure S1, the lateral sizes of Ce-WS2 ranged from 77 nm to 133 nm, and

205

the peak of the lateral size distribution was located at 94 nm. Similarly, the lateral sizes of

206

Ae-WS2 ranged from 51 nm to 146 nm, and the peak of the lateral size distribution was

207

located at 98 nm. The average thicknesses of Ce-WS2 and Ae-WS2 were 1.4 nm and 1.3

208

nm, respectively, which were very close to the height of monolayer WS2 2.

209

The alteration of hydrodynamic diameters was used to analyze nanomaterial

210

aggregation.11, 31 As shown in Figure S2, the hydrodynamic diameters increased by

211

107-128%, suggesting the aggregation of nanomaterials in the BG-11 medium. The

212

above results are consistent with a previous study in which high ionic strength was shown

213

to stimulate nanomaterial aggregation.32, 33 The data concerning zeta potential in Figure

214

S2 also confirmed that the tested nanosheets were metastable in BG-11 at a pH of

215

approximately 7.0. Increased zeta potential and aggregation led to the precipitation of

216

nanomaterials, which, in turn, affected the interaction of the nanomaterials with

217

organisms. In the present study, Ce-WS2 and Ae-WS2 showed no significant difference in

218

zeta potential or aggregation, meaning that these factors will not interfere with the study

219

of the role of phase in WS2 nanosheet toxicity.

220

Figure 1a shows X-ray powder diffraction (XRD) patterns for the Ce-WS2, Ae-WS2 11

ACS Paragon Plus Environment

Environmental Science & Technology

221

and bulk 2H WS2 samples. Ae-WS2 showed similar peaks to bulk 2H WS2. In contrast to

222

the standard 14.3° (002) peak observed in the Ae-WS2 and bulk 2H-WS2 samples, the

223

XRD patterns of Ce-WS2 samples showed a new (002)new peak at 9.4°, which was

224

consistent with 1T-WS2.9 The Raman spectra of the Ce-WS2 and Ae-WS2 nanosheets are

225

provided in Figure 1b. Two prominent peaks corresponding to the in-plane E12g and

226

out-of-plane A1g modes of 2H WS2 were observed for all samples. However, the

227

spectrum of Ce-WS2 exhibited much richer Raman features than Ae-WS2, with several

228

new peaks in the low frequency region, which corresponded to the distorted 1T-phase

229

Raman active modes and were not possible in the 2H phase. The intensities of Raman

230

modes in the low frequency region (J1, J2, and J3) for Ae-WS2 almost disappeared after

231

annealing treatment. The Raman spectrum of Ae-WS2 was very similar to that of the 2H

232

phase,2, 9 indicating that Ce-WS2 was transformed into 2H-WS2 by the annealing

233

treatment. The above results were also supported by the results of high-resolution

234

transmission electron microscopy (HRTEM). As shown in Figure 1c and 1d, the trigonal

235

lattice and honeycomb lattice facets for the 1T phase and 2H phase can be clearly

236

visualized in Ce-WS2 and Ae-WS2 nanosheets, respectively. XPS is an efficient

237

technique to evaluate the composition of 2D TMDC materials in the 1T and 2H phases.10,

238

34

239

the W4+ 4f7/2 and W4+ 4f5/2 in the 2H-WS2 (1T-WS2) phase, respectively. Ce-WS2

240

contained the signals of both the 1T and 2H forms, indicating that Ce-WS2 was a mixture

The signals at ~ 33 (32) eV and ~ 35 (34) eV, shown in Figure 1e, are attributable to

12

ACS Paragon Plus Environment

Page 12 of 44

Page 13 of 44

Environmental Science & Technology

241

of the 1T and 2H phases. 1T was the main phase of Ce-WS2, accounting for ~ 61%.

242

When the nanosheets annealed into Ae-WS2, they transformed to 100% 2H phase and

243

closely matched the bulk 2H-phase WS2.

244

Growth Inhibition

245

As shown in Figure S3a, dose-dependent algal growth inhibition was evident for all

246

durations of exposure. After 24 h of exposure, there was no significant inhibitory effect at

247

0.1 or 1 mg/L, but algal growth was significantly inhibited at 10 mg/L. The inhibition

248

rates were 24.0%, 11.3%, and 7.2% for Ce-WS2, Ae-WS2 and bulk WS2, respectively.

249

The inhibition effect was lower than those of other 2D materials, such as graphene oxide

250

and MoS2.27, 35 For all the materials, the growth inhibition increased as exposure time

251

increase (24-72 h). After 72 h, the inhibition rate decreased with increasing exposure

252

time. The above phenomenon may be due to the increase in algal cell density from 5.0 ×

253

105 to 6.2 × 106 cells/mL. In addition, the nanomaterials exhibited obvious aggregation

254

after 72 h (Figure S2), which may reduce their toxicity. Aggregated graphene exhibited

255

lower toxicity than the dispersed form.36 The inhibition rate induced by Ce-WS2 was

256

significantly higher than that induced by Ae-WS2 during the entire exposure experiment.

257

The differences in specific toxicity are discussed below, along with the mechanisms.

258

Morphological and Ultrastructural Changes in Algal Cells

259

The SEM images of the control group showed intact, smooth and well-shaped surfaces

260

(Figure 2a). After exposure to WS2 nanosheets, the cells shrunk, and some wrinkles 13

ACS Paragon Plus Environment

Environmental Science & Technology

261

(denoted by black arrows) and holes (denoted by red arrows) appeared on the surface

262

(Figure 2b and 2c). TEM images in Figure 2d-2i show the ultrastructural differences in

263

algal cells between the control and exposed groups. In the control group, an intact cell

264

wall was observed, and the cytoplasm was closely attached to the cell membrane (Figure

265

2d). In contrast, the cell walls exposed to WS2 nanosheets were damaged (as denoted by

266

the red arrows), and serious plasmolysis was also observed (as denoted by black arrows)

267

in Figure 2e and 2f. Similar phenomena were also observed in graphene-, graphene

268

oxide- and MoS2-exposed Chlorella vulgaris in our previous study27, 35, demonstrating

269

that these two-dimensional nanomaterials caused similar physical damage to algal cells.

270

In particular, the adsorption of WS2 nanomaterials on cell walls was observed (as denoted

271

by the purple arrows in Figure 2h and 2i). Cell wrapping by the aggregated sheets is an

272

important mechanism that may be involved in 2D sheet toxicity. Wrapping by nanosheets

273

hinders matter exchange (e.g., nutrient substances) between cells and the surrounding

274

environment, resulting in damage to the cell wall and membrane37, 38 . Moreover,

275

nanoparticles could also be found inside the cells, especially in chloroplasts, as denoted

276

by the green arrows in Figure 2h and 2i. The specific uptake pathways of nanomaterials

277

are discussed below.

278

Cellular Uptake

279

The cellular uptake of nanoparticles and the associated pathway can have direct

280

consequences for intracellular localization and cytotoxicity39, 40. ICP-MS showed 14

ACS Paragon Plus Environment

Page 14 of 44

Page 15 of 44

Environmental Science & Technology

281

quantitatively that the cellular Ce-WS2 and Ae-WS2 content reached 0.335 and 0.155

282

μg/106 cells, respectively, after 10 mg/L exposure (Figure 2j). The cellular internalization

283

of WS2 was significantly lower than the internalization of MoS2 by Chlorella vulgaris27,

284

which was consistent with the fact that the toxicity of WS2 nanosheets was lower than

285

that of MoS2 nanosheets. Moreover, the cellular content of Ce-WS2 was 1.79-fold higher

286

than that of Ae-WS2 at 1 mg/L. The correlation coefficient between cellular uptake and

287

growth inhibition rate was determined to be 0.93 (Figure S4a). To further evaluate the

288

effect of phase on the uptake mechanism, we used ICP-MS to quantify the uptake of

289

nanosheets under physical and pharmacological inhibition. Energy-dependent

290

endocytosis can be inhibited by exposing algal cells to low temperatures40. The uptake of

291

Ce-WS2 and Ae-WS2 was reduced by 37.09% and 38.23%, respectively, at 4 °C

292

compared to control conditions, indicating the energy dependence of the internalization

293

process (Figure 2k). However, the internalization of WS2 was still detected at 4 °C,

294

indicating the existence of energy-independent pathways as well, for example, passive

295

diffusion.39 To further identify the energy-dependent pathways, we treated algae with

296

MβCD to block caveolae-mediated endocytosis5, 40, CPZ to prevent clathrin-mediated

297

endocytosis5, and EIPA to prevent uptake by micropinocytosis.41 MβCD and EIPA

298

reduced the uptake of Ce-WS2 by 19.03% and 34.68%, respectively (Figure 2k). MβCD,

299

CPZ, and EIPA reduced the uptake of Ae-WS2 by 41.50%, 31.30% and 52.77%,

300

respectively (Figure 2k), suggesting a difference between the uptake pathways of Ce-WS2 15

ACS Paragon Plus Environment

Environmental Science & Technology

301

and Ae-WS2. The above data suggest that macropinocytosis played a dominant role in the

302

uptake of both phases of WS2 nanosheets. In contrast, clathrin-mediated endocytosis was

303

not very important in the uptake process, especially for Ce-WS2. Moreover, Ae-WS2 was

304

more sensitive than Ce-WS2 to these pharmacological inhibitors.

305

Photosynthesis and Oxidative Stress

306

As shown in Figure S5, after the algae were exposed to WS2 nanosheets for 24 h, the

307

Fv/Fm ratios (the maximum photochemical quantum yields) were slightly higher than

308

those of control cells. As the exposure time increased, the Fv/Fm ratios decreased. The

309

Fv/Fm values in the treated group at 72 h were significantly decreased compared with

310

those in the control, implying the inhibition of photosystem II (PSII).42 Then, the Fv/Fm

311

values increased slightly at 96 h, which was consistent with the results showing growth

312

inhibition (Figure S3). In addition, the Fv/Fm ratios in the Ce-WS2 groups were lower

313

than those in the Ae-WS2 groups, which was also consistent with the internalization of

314

WS2 nanosheets by algae. To explore the mechanisms underlying the decrease in

315

photosynthesis, we investigated oxidative stress. Both Ce-WS2 and Ae-WS2 triggered an

316

increase in intracellular ROS levels in algal cells, and the former induced significantly

317

higher intracellular ROS levels than the latter (Figure S3b). Figure S3c shows that WS2

318

nanosheets at 1 mg/L and 10 mg/L inhibited reduced glutathione (GSH) by 16.1-68.6%,

319

with especially strong inhibition by Ce-WS2. The correlation coefficients of ROS

320

generation with cellular GSH and growth inhibition ratio were determined to be 0.90 and 16

ACS Paragon Plus Environment

Page 16 of 44

Page 17 of 44

Environmental Science & Technology

321

0.88, respectively (Figure S4b), suggesting that oxidative stress played an important role

322

in nanotoxicity. Furthermore, abiotic GSH oxidation was used as a model system to

323

examine the possibility of oxidative stress mediated by Ce-WS2 and Ae-WS2 using

324

Ellman's assay. In Figure S6, Ce-WS2 had stronger GSH oxidation potential than

325

Ae-WS2, which was consistent with the results of biotic experiments in algal cells.

326

Lipid Peroxidation and Membrane Damage

327

Cell membrane damage is an important mechanism of nanomaterial cytotoxicity. In

328

addition to the physical damage to the cell membrane inflicted by their sharp edges or

329

rough surfaces, nanomaterials can cause membrane damage and cytotoxicity through

330

ROS generation and lipid peroxidation.43, 44 As shown in Figure 3a, both Ce-WS2 and

331

Ae-WS2, but especially Ce-WS2, caused significant lipid peroxidation of the membrane,

332

which was consistent with ROS generation and the failure of membrane integrity in TEM

333

images. RBCs have no fluid phase or receptor-mediated endocytosis and are widely used

334

to study nanomaterial interactions with the cell membrane.29 WS2 nanosheet-induced

335

membrane damage was investigated using a hemolysis assay in RBCs. As shown in

336

Figure 3b, WS2 nanosheets exhibited a slight hemolytic effect, and the hemolytic effect

337

of Ce-WS2 was stronger than that of Ae-WS2. Furthermore, the cellular permeability and

338

integrity of algal cells were examined by PI uptake. The green (P2) and red (P1) regions

339

in the flow cytometry images indicate intact cells and membrane-damaged cells

340

(membrane permeabilization), respectively (Figure 3c). Ce-WS2 induced a much higher 17

ACS Paragon Plus Environment

Environmental Science & Technology

341

level of membrane permeabilization than Ae-WS2, which was consistent with the changes

342

in membrane peroxidation.

343

Genotoxicity

344

Shape, size and surface chemicals are important parameters that can affect the potential

345

cytotoxicity of nanomaterials20,29,36. The main difference between the two tested WS2

346

nanosheets was the phases between 1T and 2H, and the roles of the phases of WS2

347

nanosheets in nanotoxicity are largely unknown. Previous reports showed that

348

graphene-based 2D nanomaterials induce genetic damage45, 46. In the present work, the

349

genotoxicity induced by WS2 nanosheets was measured by monitoring DNA damage to

350

algal cells using a comet assay. As shown in Figure S7, there was a slight increase in tail

351

DNA in the WS2 nanosheet-treated group at 10 mg/L, and there was no significant

352

difference (p > 0.05) between Ce-WS2 and Ae-WS2. This indicates that 10 mg/L WS2 has

353

low genotoxicity, which is consistent with the observation that WS2 is less toxic than

354

graphene materials.25

355

ESR Detection of ROS and Photogenerated Electrons

356

Generation of oxidative stress in cells is considered a major factor in

357

nanoparticle-induced cytotoxicity.47 The algal exposure experiment was performed under

358

visible-light irradiation. Through semiconductor photocatalysis, WS2 nanosheets may

359

produce photogenerated ROS, causing damage to the exposed organism.48, 49 To test this

360

hypothesis, we measured the generation of hydroxyl radical (•OH) and superoxide (O2•−) 18

ACS Paragon Plus Environment

Page 18 of 44

Page 19 of 44

Environmental Science & Technology

361

from irradiated Ce-WS2 and Ae-WS2. No ESR signals were observed for Ce-WS2 or

362

Ae-WS2 without irradiation (Figure 4a and 4b). After 30 min of irradiation in the

363

presence of Ce-WS2 and Ae-WS2 nanosheets, four lines with relative intensities of

364

1:2:2:1 were observed – the characteristic spectrum of the spin adduct DMPO/•OH.30

365

Furthermore, the addition of DMSO (a specific scavenger for hydroxyl radicals) almost

366

completely suppressed the signal intensity of irradiated Ce-WS2 and Ae-WS2,

367

demonstrating that hydroxyl radicals were generated from irradiated Ce-WS2 and

368

Ae-WS2, whereas no superoxide radicals were produced. With the same amount and

369

recording time, the ESR signal intensity generated from photoexcited Ce-WS2 was

370

approximately 1.33-fold higher than that from Ae-WS2, indicating that Ce-WS2 has a

371

stronger ability to photogenerate hydroxyl radicals. TEMP itself is ESR silent but can

372

react with singlet oxygen to produce a nitroxide radical with a distinctive three-line ESR

373

spectrum.30 Figure S8 shows that no singlet oxygen was generated from irradiated

374

Ce-WS2 or Ae-WS2.

375

Previous research has shown that both visible and near-infrared (NIR) light excitation

376

can cause electron-hole pairs in WS2 nanosheets50. To explore the differential ability of

377

light to generate electrons and holes on the surface of Ce-WS2 and Ae-WS2 nanosheets,

378

we used ESR spectroscopy to investigate the light-induced formation of electrons. The

379

spin label TEMPO was shown not to react with oxidative intermediates formed in

380

photoexcitation of photocatalysis, such as •OH, O2•−, 1O2 and holes.51 In the present 19

ACS Paragon Plus Environment

Environmental Science & Technology

381

study, TEMPO was used to verify electrons photogenerated by nanomaterials under

382

simulated sunlight. TEMPO exhibited a stable triplet ESR spectrum, with a relative

383

intensity of 1:1:1. As shown in Figure 4c and 4d, the signal intensity was unchanged

384

when TEMPO was mixed with WS2 nanosheets before irradiation or irradiated without

385

WS2 nanosheets (control). In contrast, an obvious reduction in the ESR signal intensity

386

was observed for the reaction of TEMPO with WS2 nanosheets under irradiation for 15

387

min. The reduction of TEMPO signal intensity caused by Ce-WS2 and Ae-WS2 was

388

calculated to be 68.23% and 59.68%, respectively. After 30 min of irradiation, the signal

389

intensity of TEMPO was reduced by 96.44% and 90.34% for Ce-WS2 and Ae-WS2,

390

respectively. The above results suggested that Ce-WS2 produced more electrons than

391

Ae-WS2 under simulated sunlight.

392

Photogenerated electrons are related to the band gaps of semiconductors. Low band

393

gaps are facile for the photogeneration of electrons.30, 52, 53 As shown in Figure S9, the

394

band gap of Ce-WS2 (1.67 eV) was lower than that of Ae-WS2 (2.04 eV), indicating that

395

Ce-WS2 was more likely to generate electrons and holes. In addition, due to the presence

396

of the metallic 1T phase, Ce-WS2 presented higher electron conductivity than that of

397

Ae-WS2 54, 55, which substantially improved the charge transfer efficiency and separation

398

efficiency of electrons and holes, leading to an increase in ROS (mainly •OH) levels and

399

nanotoxicity.3, 55

400

Release of Dissolved Ions 20

ACS Paragon Plus Environment

Page 20 of 44

Page 21 of 44

Environmental Science & Technology

401

Toxic ions released from nanomaterials may contribute to nanotoxicity.56 Figure S10

402

shows that the amount of dissolved W from WS2 nanosheets was greater in BG-11

403

culture medium than in deionized water. In addition, ions were released much more

404

rapidly from Ce-WS2 than from Ae-WS2 in both deionized water and BG-11 culture

405

medium. Similar results were reported by Wang et al.,11 who demonstrated that the

406

oxidation dissolution of chemically exfoliated MoS2 nanosheets was faster than that of

407

ultrasonically exfoliated nanosheets due to the presence of the 1T phase in the former, as

408

opposed to the pure 2H phase in the latter. Shang et al.57 demonstrated that the

409

enhancement of WS2 nanoparticle (NP) dissolution by UV light was due to the

410

photocorrosion of NPs by the holes: WS2 + 6h+

411

4d, Ce-WS2 produced more electrons and holes than Ae-WS2, which explains why

412

Ce-WS2 released ions much more quickly than Ae-WS2 under irradiation. The dissolution

413

products of WS2 nanosheets usually contain tungstate, W(VI).11, 27, 57 Tungstate salt

414

(Na2WO4) was chosen as the control to study the toxic effects of the ions on algae. As

415

shown in Figure S11a, tungstate salt showed no statistically significant inhibitory effect

416

on cell growth at a concentration of 50 μM (corresponding to WS2 nanosheets at

417

approximately 12.4 mg/L). Intracellular ROS and cell membrane integrity were also not

418

affected by tungstate salt (Figure S11b and S11c), suggesting that the differences in

419

toxicity between Ce-WS2 and Ae-WS2 were derived from nanosheets rather than the

420

released ions.

ℎ𝑣

W6+ + 2S. As shown in Figure 4c and

21

ACS Paragon Plus Environment

Environmental Science & Technology

421

Metabolic Mechanisms of Biological Responses

422

The identified metabolites included amino acids, fatty acids, small-molecule acids,

423

carbohydrates, and other biomolecules. The relative levels of the metabolites are

424

presented using heat maps in Figure S12. Using hierarchical clustering (HCL) analysis,

425

Ce-WS2 induced more obvious metabolite alteration than Ae-WS2 relative to the control

426

treatment. The two-dimensional score plots of the principal component analysis (PCA)

427

and partial least squares-discriminant analysis (PLS-DA) models both demonstrated that

428

the Ce-WS2 and Ae-WS2 groups were separated from each other and from the control

429

group (Figure 5a and 5b), indicating that WS2 nanosheets affected cellular metabolism.

430

As shown in Figure S13, WS2 nanosheet exposure significantly upregulated the levels of

431

amino acids and fatty acids. Previous reports suggested that MoS2 nanosheets exposed to

432

human dermal fibroblasts upregulate intracellular amino acids via the degradation of

433

abnormal proteins.58 The high levels of free fatty acids are related to β-oxidation

434

processes, and the upregulation of fatty acids acts as an organismal adaptation to

435

membrane oxidative stress.59, 60 Herein, Ce-WS2 induced more obvious perturbation of

436

the algal metabolic profile of amino acids and fatty acids than Ae-WS2 (Figure S13),

437

which was consistent with the higher toxicity of Ce-WS2. Palmitic acid and stearic acid

438

were significantly upregulated. Previous reports showed that the upregulation of palmitic

439

acid and stearic acid results in cell membrane damage,61 and damage to cell plasma

440

membranes is observed in Figure 3c. Moreover, increases in palmitic acid and stearic 22

ACS Paragon Plus Environment

Page 22 of 44

Page 23 of 44

Environmental Science & Technology

441

acid levels inhibit electron transport in photosynthesis and disintegrate phycobilin from

442

the thylakoid membrane,61, 62 which is consistent with the decrease in maximum

443

photosynthesis induced by WS2 nanosheets (Figure S5). The differentially regulated

444

metabolites were further defined according to a p value < 0.05 and false discovery rate

445

(FDR) < 0.05. There were 28 and 16 specific metabolites that distinguished Ce-WS2 and

446

Ae-WS2, respectively, from the control group (Figure S14). Next, the differentially

447

regulated metabolites were mapped onto their corresponding metabolic pathways. As

448

shown in Figure 5c and 5d, both Ce-WS2 and Ae-WS2 upregulated the pathways of

449

alanine, aspartate and glutamate metabolism, glycine, serine and threonine metabolism,

450

arginine and proline metabolism and glutathione metabolism. Metabolic pathways such

451

as glycine, serine, threonine metabolism and glutamate metabolism are involved in GSH

452

biosynthesis and metabolism.63,64 However, the results in Figure S3c show a decrease in

453

intracellular GSH levels. ROS generated by WS2 nanosheets depleted GSH and then

454

stimulated cells to promote GSH synthesis by compensation. In addition, Ce-WS2

455

downregulated starch and sucrose metabolism and glycerolipid metabolism, while

456

Ae-WS2 downregulated only glycerolipid metabolism (Figure 5c and 5d). Starch and

457

sucrose are important sources of energy storage in algae. The downregulation of starch

458

and sucrose metabolism may be associated with the reduced photosynthetic activities

459

under WS2 nanosheet stress.65 The above results showed that metabolic analysis can

460

provide new insights into the differences in toxicity between Ce-WS2 and Ae-WS2. 23

ACS Paragon Plus Environment

Environmental Science & Technology

461

The phase of a nanomaterial is a critical determinant of its applications and

462

environmental fate, such as pollutant adsorption, catalysis, antibacterial activity and

463

degradation.2, 9, 66 However, the biological responses to different phases of WS2

464

nanosheets remain largely unknown, which limits their environmental risk evaluation and

465

their suitability for environmentally friendly design. The results of the present study

466

demonstrated that the phases of WS2 nanosheets affected oxidative stress, membrane

467

damage, lipid peroxidation, uptake and cytotoxicity in algae. The metallic 1T phase

468

showed higher toxicity than the semiconducting 2H phase. Noticeably, WS2 nanosheets

469

at 0.1 mg/L did not trigger remarkable toxicity to algae and seemed safe at predicted

470

environmental concentrations (e.g., μg/L).67, 68 However, new nanomaterials, especially

471

2D nanomaterials, are being developed quickly, and investigations into these novel

472

nanomaterials might reveal materials of higher environmental concern.69 Compared with

473

the well-studied effects of size, shape and surface modifications of nanomaterials on

474

nanotoxicology, little is known about the effects of phase. The results of the present study

475

highlight the importance of considering the phase of a nanomaterial when evaluating its

476

environmental risk.

477 478

ASSOCIATED CONTENT

479

Supporting Information Available

480

Methods are provided for WS2 nanosheet characterization, electron microscopic 24

ACS Paragon Plus Environment

Page 24 of 44

Page 25 of 44

Environmental Science & Technology

481

observations, oxidative stress, the comet assay and metabolomic analysis. Figures S1-S14

482

show the results of WS2 nanosheet characterization, growth inhibition, ROS levels, GSH

483

content, correlation analysis, photosynthetic efficiency, abiotic oxidation, the comet

484

assay, ESR spectra, band gap, cytotoxicity and metabolomic analysis.

485 486

AUTHOR INFORMATION

487

Corresponding author:

488

*Email: [email protected] (X.H.).

489

Tel.: 86-22-23507800; fax: 86-22-23507800.

490 491

NOTES

492

The authors declare no competing financial interests.

493 494

ACKNOWLEDGMENTS

495

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

496

China (grant nos. 21577070 and 21307061).

497 498

REFERENCES

499

1.

500

metal dichalcogenides. Nat. Rev. Mater. 2017, 2 (8).

Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A., 2D transition

25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 44

501

2.

502

W.; Asefa, T.; Shenoy, V. B.; Eda, G.; Chhowalla, M., Enhanced catalytic activity in

503

strained chemically exfoliated WS2 nanosheets for hydrogen evolution. Nat. Mater. 2013,

504

12 (9), 850-855.

505

3.

506

Eda, G.; Chhowalla, M., Conducting MoS2 nanosheets as catalysts for hydrogen

507

evolution reaction. Nano Lett. 2013, 13 (12), 6222-6227.

508

4.

509

Y., High supercapacitor and adsorption behaviors of flower-like MoS2 nanostructures. J.

510

Mater. Chem. A 2014, 2 (38), 15958-15963.

511

5.

512

Tao, W.; Cai, T.; Li, Y. J.; Gan, T.; Barrett, A.; Bharwani, Z.; Chen, H. B.; Farokhzad, O.

513

C.,

514

anti-exocytosis-enhanced synergistic cancer therapy. ACS Nano 2018, 12 (3), 2922-2938.

515

6.

516

Liu, G.; Xing, H. Y.; Bu, W. B.; Sun, B. Q.; Liu, Z., PEGylated WS2 nanosheets as a

517

multifunctional theranostic agent for in vivo dual-modal CT/photoacoustic imaging

518

guided photothermal therapy. Adv. Mater. 2014, 26 (12), 1886-1893.

519

7.

520

of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013,

Voiry, D.; Yamaguchi, H.; Li, J. W.; Silva, R.; Alves, D. C. B.; Fujita, T.; Chen, M.

Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M. W.; Asefa, T.; Shenoy, V. B.;

Wang, X. H.; Ding, J. J.; Yao, S. W.; Wu, X. X.; Feng, Q. Q.; Wang, Z. H.; Geng, B.

Zhu, X. B.; Ji, X. Y.; Kong, N.; Chen, Y. H.; Mahmoudi, M.; Xu, X. D.; Ding, L.;

Intracellular

mechanistic

understanding

of

2D

MoS2

nanosheets

for

Cheng, L.; Liu, J. J.; Gu, X.; Gong, H.; Shi, X. Z.; Liu, T.; Wang, C.; Wang, X. Y.;

Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L. J.; Loh, K. P.; Zhang, H., The chemistry

26

ACS Paragon Plus Environment

Page 27 of 44

Environmental Science & Technology

521

5 (4), 263-275.

522

8.

523

Group 6 transition metal dichalcogenide nanomaterials: synthesis, applications and future

524

perspectives. Nanoscale Horiz. 2018, 3 (2), 90-204.

525

9.

526

Q.; Chu, W. S.; Wu, X. J.; Yang, J. L.; Wang, C. M.; Xiong, Y. J.; Jin, C. H.; Ajayan, P.

527

M.; Song, L., Stable metallic 1T-WS2 nanoribbons intercalated with ammonia ions: the

528

correlation between structure and electrical/optical properties. Adv. Mater. 2015, 27 (33),

529

4837-4844.

530

10. Mahler, B.; Hoepfner, V.; Liao, K.; Ozin, G. A., Colloidal synthesis of 1T-WS2 and

531

2H-WS2 nanosheets: applications for photocatalytic hydrogen evolution. J. Am. Chem.

532

Soc. 2014, 136 (40), 14121-14127.

533

11. Wang, Z. Y.; von dem Bussche, A.; Qiu, Y.; Valentin, T. M.; Gion, K.; Kane, A. B.;

534

Hurt, R. H., Chemical dissolution pathways of MoS2 nanosheets in biological and

535

environmental media. Environ. Sci. Technol. 2016, 50 (13), 7208-7217.

536

12. Amorim, M. J. B.; Lin, S. J.; Schlich, K.; Navas, J. M.; Brunelli, A.; Neubauer, N.;

537

Vilsmeier, K.; Costa, A. L.; Gondikas, A.; Xia, T.; Galbis, L.; Badetti, E.; Marcomini, A.;

538

Hristozov, D.; von der Karnmer, F.; Hund-Rinke, K.; Scott-Fordsmand, J. J.; Nel, A.;

539

Wohlleben, W., Environmental impacts by fragments released from nanoenabled

540

products: a multiassay, multimaterial exploration by the SUN approach. Environ. Sci.

Samadi, M.; Sarikhani, N.; Zirak, M.; Zhang, H.; Zhang, H. L.; Moshfegh, A. Z.,

Liu, Q.; Li, X. L.; Xiao, Z. R.; Zhou, Y.; Chen, H. P.; Khalil, A.; Xiang, T.; Xu, J.

27

ACS Paragon Plus Environment

Environmental Science & Technology

541

Technol. 2018, 52 (3), 1514-1524.

542

13. Hu, X.; Li, D.; Gao, Y.; Mu, L.; Zhou, Q., Knowledge gaps between

543

nanotoxicological research and nanomaterial safety. Environ. Int. 2016, 94, 8-23.

544

14. Wang, W. Y.; Sedykh, A.; Sun, H. N.; Zhao, L. L.; Russo, D. P.; Zhou, H. Y.; Yan,

545

B.; Zhu, H., Predicting nano-bio interactions by integrating nanoparticle libraries and

546

quantitative nanostructure activity relationship modeling. ACS Nano 2017, 11 (12),

547

12641-12649.

548

15. Akhavan, O.; Ghaderi, E.; Hashemi, E.; Akbari, E., Dose-dependent effects of

549

nanoscale graphene oxide on reproduction capability of mammals. Carbon 2015, 95,

550

309-317.

551

16. Ren, M. X.; Zhao, L.; Ding, X. C.; Krasteva, N.; Rui, Q.; Wang, D. Y.,

552

Developmental basis for intestinal barrier against the toxicity of graphene oxide. Part.

553

Fibre Toxicol. 2018, 15.

554

17. Reddy Pullagurala, V. L.; Adisa, I. O.; Rawat, S.; Kim, B.; Barrios, A. C.;

555

Medina-Velo, I. A.; Hernandez-Viezcas, J. A.; Peralta-Videa, J. R.; Gardea-Torresdey, J.

556

L., Finding the conditions for the beneficial use of ZnO nanoparticles towards plants-A

557

review. Environ. Pollut. 2018, 241, 1175-1181.

558

18. Hanna, S. K.; Bustos, A. R. M.; Peterson, A. W.; Reipa, V.; Scanlan, L. D.; Coskun,

559

S. H.; Cho, T. J.; Johnson, M. E.; Hackley, V. A.; Nelson, B. C.; Winchester, M. R.;

560

Elliott, J. T.; Petersen, E. J., Agglomeration of Escherichia coli with positively charged 28

ACS Paragon Plus Environment

Page 28 of 44

Page 29 of 44

Environmental Science & Technology

561

nanoparticles can lead to artifacts in a standard caenorhabditis elegans toxicity assay.

562

Environ. Sci. Technol. 2018, 52 (10), 5968-5978.

563

19. Mortimer, M.; Devarajan, N.; Li, D.; Holden, P. A., Multiwall carbon nanotubes

564

induce more pronounced transcriptomic responses in Pseudomonas aeruginosa PG201

565

than graphene, exfoliated boron nitride, or carbon black. ACS Nano 2018, 12 (3),

566

2728-2740.

567

20. Spielman-Sun, E.; Lombi, E.; Donner, E.; Howard, D.; Unrine, J. M.; Lowry, G. V.,

568

Impact of surface charge on cerium oxide nanoparticle uptake and translocation by wheat

569

(Triticum aestivum). Environ. Sci. Technol. 2017, 51 (13), 7361-7368.

570

21. Li, X.; Shan, J. Y.; Zhang, W. Z.; Su, S.; Yuwen, L. H.; Wang, L. H., Recent

571

advances in synthesis and biomedical applications of two-dimensional transition metal

572

dichalcogenide nanosheets. Small 2017, 13 (5).

573

22. Ouyang, S.; Hu, X.; Zhou, Q.; Li, X.; Miao, X.; Zhou, R., Nanocolloids in natural

574

water: isolation, characterization, and toxicity. Environ. Sci. Technol. 2018, 52 (8),

575

4850-4860.

576

23. Brunetti, A. E.; Neto, F. C.; Vera, M. C.; Taboada, C.; Pavarini, D. P.; Bauermeister,

577

A.; Lopes, N. P., An integrative omics perspective for the analysis of chemical signals in

578

ecological interactions. Chem. Soc. Rev. 2018, 47 (5), 1574-1591.

579

24. Gunsolus, I. L.; Haynes, C. L., Analytical aspects of nanotoxicology. Anal. Chem.

580

2016, 88 (1), 451-479. 29

ACS Paragon Plus Environment

Environmental Science & Technology

581

25. Teo, W. Z.; Chng, E. L. K.; Sofer, Z.; Pumera, M., Cytotoxicity of exfoliated

582

transition-metal dichalcogenides (MoS2, WS2, and WSe2) is lower than that of graphene

583

and its analogues. Chem.-Eur. J. 2014, 20 (31), 9627-9632.

584

26. Beddow, J.; Johnson, R. J.; Lawson, T.; Breckels, M. N.; Webster, R. J.; Smith, B.

585

E.; Rowland, S. J.; Whitby, C., The effect of oil sands process -affected water and model

586

naphthenic acids on photosynthesis and growth in Emiliania huxleyi and Chlorella

587

vulgaris. Chemosphere 2016, 145, 416-423.

588

27. Zou, W.; Zhou, Q. X.; Zhang, X. L.; Hu, X. G., Environmental transformations and

589

algal toxicity of single-layer molybdenum disulfide regulated by humic acid. Environ.

590

Sci. Technol. 2018, 52 (5), 2638-2648.

591

28. Miller, R. J.; Muller, E. B.; Cole, B.; Martin, T.; Nisbet, R.; Bielmyer-Fraser, G. K.;

592

Jarvis, T. A.; Keller, A. A.; Cherr, G.; Lenihan, H. S., Photosynthetic efficiency predicts

593

toxic effects of metal nanomaterials in phytoplankton. Aquat. Toxicol. 2017, 183, 85-93.

594

29. Li, R. B.; Guiney, L. M.; Chang, C. H.; Mansukhani, N. D.; Ji, Z. X.; Wang, X.;

595

Liao, Y. P.; Jiang, W.; Sun, B. B.; Hersam, M. C.; Nel, A. E.; Xia, T., Surface oxidation

596

of graphene oxide determines membrane damage, lipid peroxidation, and cytotoxicity in

597

macrophages in a pulmonary toxicity model. ACS Nano 2018, 12 (2), 1390-1402.

598

30. He, W. W.; Kim, H. K.; Warner, W. G.; Melka, D.; Callahan, J. H.; Yin, J. J.,

599

Photogenerated charge carriers and reactive oxygen species in ZnO/Au hybrid

600

nanostructures with enhanced photocatalytic and antibacterial activity. J. Am. Chem. Soc. 30

ACS Paragon Plus Environment

Page 30 of 44

Page 31 of 44

Environmental Science & Technology

601

2014, 136 (2), 750-757.

602

31. Wang, X.; Mansukhani, N. D.; Guiney, L. M.; Ji, Z. X.; Chang, C. H.; Wang, M. Y.;

603

Liao, Y. P.; Song, T. B.; Sun, B. B.; Li, R. B.; Xia, T.; Hersam, M. C.; Nel, A. E.,

604

Differences in the toxicological potential of 2D versus aggregated molybdenum disulfide

605

in the lung. Small 2015, 11 (38), 5079-5087.

606

32. Reid, M. S.; Kedzior, S. A.; Villalobos, M.; Cranston, E. D., Effect of ionic strength

607

and surface charge density on the kinetics of cellulose nanocrystal thin film swelling.

608

Langmuir 2017, 33 (30), 7403-7411.

609

33. Kamrani, S.; Rezaei, M.; Kord, M.; Baalousha, M., Transport and retention of carbon

610

dots (CDs) in saturated and unsaturated porous media: Role of ionic strength, pH, and

611

collector grain size. Water Res. 2018, 133, 338-347.

612

34. Acerce, M.; Voiry, D.; Chhowalla, M., Metallic 1T phase MoS2 nanosheets as

613

supercapacitor electrode materials. Nat. Nanotechnol. 2015, 10 (4), 313-318.

614

35. Ouyang, S.; Hu, X.; Zhou, Q., Envelopment-internalization synergistic effects and

615

metabolic mechanisms of graphene oxide on single-cell chlorella vulgaris are dependent

616

on the nanomaterial particle size. ACS Appl. Mater. Interfaces 2015, 7 (32),

617

18104-18112.

618

36. Liao, K. H.; Lin, Y. S.; Macosko, C. W.; Haynes, C. L., Cytotoxicity of graphene

619

oxide and graphene in human erythrocytes and skin fibroblasts. ACS Appl. Mater.

620

Interfaces 2011, 3 (7), 2607-2615. 31

ACS Paragon Plus Environment

Environmental Science & Technology

621

37. Akhavan, O.; Ghaderi, E.; Esfandiar, A., Wrapping bacteria by graphene nanosheets

622

for isolation from environment, reactivation by sonication, and inactivation by

623

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

624

38. Mu, L.; Gao, Y.; Hu, X. G., L-Cysteine: A biocompatible, breathable and beneficial

625

coating for graphene oxide. Biomaterials 2015, 52, 301-311.

626

39. Herd, H.; Daum, N.; Jones, A. T.; Huwer, H.; Ghandehari, H.; Lehr, C. M.,

627

Nanoparticle geometry and surface orientation influence mode of cellular uptake. ACS

628

Nano 2013, 7 (3), 1961-1973.

629

40. Saikia, J.; Yazdimamaghani, M.; Moghaddam, S. P. H.; Ghandehari, H., Differential

630

protein adsorption and cellular uptake of silica nanoparticles based on size and porosity.

631

ACS Appl. Mater. Interfaces 2016, 8 (50), 34820-34832.

632

41. Yang, L. X.; Shang, L.; Nienhaus, G. U., Mechanistic aspects of fluorescent gold

633

nanocluster internalization by live HeLa cells. Nanoscale 2013, 5 (4), 1537-1543.

634

42. Chen, X. L.; O'Halloran, J.; Jansen, M. A. K., The toxicity of zinc oxide

635

nanoparticles to Lemna minor (L.) is predominantly caused by dissolved Zn. Aquat.

636

Toxicol. 2016, 174, 46-53.

637

43. Akhavan, O.; Ghaderi, E., Toxicity of graphene and graphene oxide nanowalls

638

against bacteria. ACS Nano 2010, 4 (10), 5731-5736.

639

44. Hu, W. B.; Peng, C.; Lv, M.; Li, X. M.; Zhang, Y. J.; Chen, N.; Fan, C. H.; Huang,

640

Q., Protein corona-mediated mitigation of cytotoxicity of graphene oxide. ACS Nano 32

ACS Paragon Plus Environment

Page 32 of 44

Page 33 of 44

Environmental Science & Technology

641

2011, 5 (5), 3693-3700.

642

45. Akhavan, O.; Ghaderi, E.; Akhavan, A., Size-dependent genotoxicity of graphene

643

nanoplatelets in human stem cells. Biomaterials 2012, 33 (32), 8017-8025.

644

46. Akhavan, O.; Ghaderi, E.; Emamy, H.; Akhavan, F., Genotoxicity of graphene

645

nanoribbons in human mesenchymal stem cells. Carbon 2013, 54, 419-431.

646

47. Mu, Q. X.; Jiang, G. B.; Chen, L. X.; Zhou, H. Y.; Fourches, D.; Tropsha, A.; Yan,

647

B., Chemical basis of interactions between engineered nanoparticles and biological

648

systems. Chem. Rev. 2014, 114 (15), 7740-7781.

649

48. Priester, J. H.; Moritz, S. C.; Espinosa, K.; Ge, Y.; Wang, Y.; Nisbet, R. M.;

650

Schimel, J. P.; Goggi, A. S.; Gardea-Torresdey, J. L.; Holden, P. A., Damage assessment

651

for soybean cultivated in soil with either CeO2 or ZnO manufactured nanomaterials. Sci.

652

Total Environ. 2017, 579, 1756-1768.

653

49. Elgrabli, D.; Dachraoui, W.; Menard-Moyon, C.; Liu, X. J.; Begin, D.; Begin-Colin,

654

S.; Bianco, A.; Gazeau, F.; Alloyeau, D., Carbon nanotube degradation in macrophages:

655

live nanoscale monitoring and understanding of biological pathway. ACS Nano 2015, 9

656

(10), 10113-10124.

657

50. Sang, Y. H.; Zhao, Z. H.; Zhao, M. W.; Hao, P.; Leng, Y. H.; Liu, H., From UV to

658

near-infrared, WS2 nanosheet: a novel photocatalyst for full solar light spectrum

659

photodegradation. Adv. Mater. 2015, 27 (2), 363-369.

660

51. Jia, H. M.; He, W. W.; Wamer, W. G.; Han, X. N.; Zhang, B. B.; Zhang, S.; Zheng, 33

ACS Paragon Plus Environment

Environmental Science & Technology

661

Z.; Xiang, Y.; Yin, J. J., Generation of reactive oxygen species, electrons/holes, and

662

photocatalytic degradation of rhodamine B by photoexcited CdS and Ag2S micro-nano

663

structures. J. Phys. Chem. C 2014, 118 (37), 21447-21456.

664

52. Zhao, H. X.; Chen, X. Y.; Li, X. T.; Shen, C.; Qu, B. C.; Gao, J. S.; Chen, J. W.;

665

Quan, X., Photoinduced formation of reactive oxygen species and electrons from metal

666

oxide-silica nanocomposite: An EPR spin-trapping study. Appl. Surf. Sci. 2017, 416,

667

281-287.

668

53. Hailili, R.; Wang, Z. Q.; Li, Y. X.; Wang, Y. H.; Sharma, V. K.; Gong, X. Q.; Wang,

669

C. Y., Oxygen vacancies induced visible-light photocatalytic activities of CaCu3Ti4O12

670

with controllable morphologies for antibiotic degradation. Appl. Catal. B-Environ. 2018,

671

221, 422-432.

672

54. Xu, H.; Yi, J. J.; She, X. J.; Liu, Q.; Song, L.; Chen, S. M.; Yang, Y. C.; Song, Y. H.;

673

Vajtai, R.; Lou, J.; Li, H. M.; Yuan, S. Q.; Wu, J. J.; Ajayan, P. M., 2D heterostructure

674

comprised of metallic 1T-MoS2/Monolayer O-g-C3N4 towards efficient photocatalytic

675

hydrogen evolution. Appl. Catal. B-Environ. 2018, 220, 379-385.

676

55. Fan, J. J.; Li, Y. F.; Nguyen, H. N.; Yao, Y.; Rodrigues, D. F., Toxicity of

677

exfoliated-MoS2 and annealed exfoliated-MoS2 towards planktonic cells, biofilms, and

678

mammalian cells in the presence of electron donor. Environ.-Sci. Nano 2015, 2 (4),

679

370-379.

680

56. Djurisic, A. B.; Leung, Y. H.; Ng, A. M. C.; Xu, X. Y.; Lee, P. K. H.; Degger, N.; 34

ACS Paragon Plus Environment

Page 34 of 44

Page 35 of 44

Environmental Science & Technology

681

Wu, R. S. S., Toxicity of metal oxide nanoparticles: mechanisms, characterization, and

682

avoiding experimental artefacts. Small 2015, 11 (1), 26-44.

683

57. Shang, E. X.; Niu, J. F.; Li, Y.; Zhou, Y. J.; Crittenden, J. C., Comparative toxicity

684

of Cd, Mo, and W sulphide nanomaterials toward E.coli under UV irradiation. Environ.

685

Pollut. 2017, 224, 606-614.

686

58. Yu, Y. D.; Wu, N.; Li, Y. L.; Li, Y. Y.; Zhang, L.; Yang, Q.; Miao, W. J.; Ding, X.

687

F.; Jiang, L.; Huang, H., Dispersible MoS2 nanosheets activated TGF-β/smad Pathway

688

and perturbed the metabolome of human dermal fibroblasts. ACS Biomater. Sci. Eng.

689

2017, 3 (12), 3261-3272.

690

59. Sanchis, J.; Llorca, M.; Olmos, M.; Schirinzi, G. F.; Bosch-Orea, C.; Abad, E.;

691

Barcelo, D.; Farre, M., Metabolic responses of Mytilus galloprovincialis to Fullerenes in

692

Mesocosm Exposure Experiments. Environ. Sci. Technol. 2018, 52 (3), 1002-1013.

693

60. Fokina, N. N.; Ruokolainen, T. R.; Nemova, N. N.; Bakhmet, I. N., Changes of blue

694

mussels Mytilus edulis L. lipid composition under cadmium and copper toxic effect. Biol.

695

Trace Elem. Res. 2013, 154 (2), 217-225.

696

61. Wu, J. T.; Chiang, Y. R.; Huang, W. Y.; Jane, W. N., Cytotoxic effects of free fatty

697

acids on phytoplankton algae and cyanobacteria. Aquat. Toxicol. 2006, 80 (4), 338-345.

698

62. Venediktov, P. S.; Krivoshejeva, A. A., The mechanisms of fatty-acid inhibition of

699

electron transport in chloroplasts. Planta 1983, 159 (5), 411-414.

700

63. Lu, S. C., Glutathione synthesis. Biochim. Biophys. Acta-Gen. Subj. 2013, 1830 (5), 35

ACS Paragon Plus Environment

Environmental Science & Technology

701

3143-3153.

702

64. Elie, M. R.; Choi, J.; Nkrumah-Elie, Y. M.; Gonnerman, G. D.; Stevens, J. F.;

703

Tanguay, R. L., Metabolomic analysis to define and compare the effects of PAHs and

704

oxygenated PAHs in developing zebrafish. Environ. Res. 2015, 140, 502-510.

705

66. Zuo, X. W.; Zhang, H. G.; Zhu, Q.; Wang, W. F.; Feng, J.; Chen, X. G., A dual-color

706

fluorescent biosensing platform based on WS2 nanosheet for detection of Hg2+ and Ag+.

707

Biosens. Bioelectron. 2016, 85, 464-470.

708

67. Wang, Y.; Nowack, B., Environmental risk assessment of engineered nano-SiO2,

709

nano iron oxides, nano-CeO2, nano-Al2O3, and quantum dots. Environ. Toxicol. Chem.

710

2018, 37 (5), 1387-1395.

711

68. Black, M. N.; Henry, E. F.; Adams, O. A.; Bennett, J. C. F.; MacCormack, T. J.,

712

Environmentally relevant concentrations of amine-functionalized copper nanoparticles

713

exhibit different mechanisms of bioactivity in Fundulus Heteroclitus in fresh and

714

brackish water. Nanotoxicology 2017, 11 (8), 1070-1085.

715

69. Arvidsson, R., Risk assessments show engineered nanomaterials to be of low

716

environmental concern. Environ. Sci. Technol. 2018, 52 (5), 2436-2437.

717

36

ACS Paragon Plus Environment

Page 36 of 44

Page 37 of 44

Environmental Science & Technology

718

Figure Captions

719

Figure 1. Characterizations of Ce-WS2 and Ae-WS2. a, XRD patterns of Ce-WS2 and

720

Ae-WS2z compared with the peak line of bulk 2H-WS2 from the Joint Committee on

721

Powder Diffraction Standards (JCPDS) card; b, Raman spectra of Ce-WS2, Ae-WS2 and

722

bulk WS2; c, high-resolution TEM images of Ce-WS2; d, high-resolution TEM images of

723

Ae-WS2. To show the crystalline structure clearly, enlarged views of the basal planes are

724

inserted in c and d. e, XPS spectra and deconvolution analysis of Ce-WS2, Ae-WS2 and

725

bulk WS2.

726 727

Figure 2. Uptake of Ce-WS2 and Ae-WS2 and damage to the cellular structure. a-c, SEM

728

images; d-I, TEM images of algae exposed to WS2 nanosheets of different phases. j and

729

k, Cellular uptake rate and pathway of WS2 nanosheets. a, d and g, Control without

730

nanomaterial exposure; b, e, and h, Ce-WS2 at 10 mg/L; c, f and i, Ae-WS2 at 10 mg/L; j,

731

cellular nanomaterial content; k, effect of physical inhibition and pharmacological

732

inhibitors on the uptake of WS2 nanosheets in algal cells. “∗”, represents statistical

733

significance at p < 0.05.

734 735

Figure 3. Lipid peroxidation and membrane damage induced by WS2 nanosheets of

736

different phases in Chlorella vulgaris. A, MDA content; b, red blood cell hemolysis; c,

737

flow cytometry images of PI-dyed algal cells after 96 h of exposure. “∗” represents 37

ACS Paragon Plus Environment

Environmental Science & Technology

738

statistical significance at p < 0.05.

739 740

Figure 4. ESR spectra of ROS and photogenerated electrons. ESR spectra were obtained

741

from samples containing DMPO with Ce-WS2 (a) and DMPO with Ae-WS2 (b) with and

742

without the addition of DMSO. ESR spectra were obtained from Ce-WS2 with TEMPO

743

(c) and Ae-WS2 with TEMPO (d).

744 745

Figure 5. Metabolic analysis of Chlorella vulgaris exposed to WS2 nanosheets. PCA

746

score plot (a) and PLS-DA score plot (b) of metabolites in algal cells exposed to WS2

747

nanosheets. Significantly disturbed metabolic pathways in algal cells influenced by

748

Ce-WS2 (c) and Ae-WS2 (d) treatment. The red and green arrows indicate the up- and

749

downregulated metabolic pathways, respectively.

750 751 752

38

ACS Paragon Plus Environment

Page 38 of 44

Page 39 of 44

Environmental Science & Technology

753 754

Figure 1.

39

ACS Paragon Plus Environment

Environmental Science & Technology

755 756

Figure 2.

40

ACS Paragon Plus Environment

Page 40 of 44

Page 41 of 44

Environmental Science & Technology

757 758

Figure 3.

759 760

41

ACS Paragon Plus Environment

Environmental Science & Technology

761 762

Figure 4.

42

ACS Paragon Plus Environment

Page 42 of 44

Page 43 of 44

Environmental Science & Technology

763 764

Figure 5.

43

ACS Paragon Plus Environment

Environmental Science & Technology

ACS Paragon Plus Environment

Page 44 of 44