Bioaccumulation and Toxicity of 13C-Skeleton Labeled Graphene

Aug 3, 2017 - ... can be downloaded at http://elder.njjcbio.com/index_en.php. ..... characterization of unlabeled GO (Figure S1); identification of GO...
8 downloads 0 Views 2MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article 13

Bioaccumulation and Toxicity of C-skeleton Labeled Graphene Oxide in Wheat Lingyun Chen, Chenglong Wang, Hongliang Li, Xiulong Qu, Shengtao Yang, and Xue-Ling Chang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b00822 • Publication Date (Web): 03 Aug 2017 Downloaded from http://pubs.acs.org on August 8, 2017

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

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

Page 1 of 29

Environmental Science & Technology

1

Revised No. es-2017-008228

2

Bioaccumulation and Toxicity of 13C-skeleton

3

Labeled Graphene Oxide in Wheat

4

Lingyun Chen1,§, Chenglong Wang2,§, Hongliang Li1,2, Xiulong Qu1, Sheng-Tao Yang1,*, and

5

Xue-Ling Chang2,*

6

1

7

Chengdu 610041, P. R. China;

8

2

9

High Energy Physics, Chinese Academy of Sciences, Beijing 100049, P. R. China.

College of Chemistry and Environment Protection Engineering, Southwest Minzu University,

CAS Key Laboratory for Biomedical Effects of Nanomaterials and Nanosafety, Institute of

10

§

11

KEYWORDS: Graphene oxide, Bioaccumulation, Toxicity, 13C-stable isotope, Oxidative stress

12

*Corresponding Author

13

Sheng-Tao

14

[email protected]

15

Xue-Ling

16

[email protected]

These two authors contributed equally.

Yang,

Chang,

Tel:

+86-28-85522269,

Fax:

+86-10-88236456,

Email

address:

Tel:

+86-10-88236709,

Fax:

+86-10-88236456,

Email

address:

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 29

17

ABSTRACT: Graphene nanomaterials have many diverse applications, but are considered to be

18

emerging environmental pollutants. Thus, their potential environmental risks and biosafety are

19

receiving increased attention. Bioaccumulation and toxicity evaluations in plants are essential for

20

biosafety assessment. In this study, 13C-stable isotope labeling of the carbon skeleton of

21

graphene oxide (GO) was applied to investigate the bioaccumulation and toxicity of GO in wheat.

22

Bioaccumulation of GO was accurately quantified according to the 13C/12C ratio. Wheat

23

seedlings were exposed to 13C-labeled GO at 1.0 mg/mL in nutrient solution for 15 d. 13C-GO

24

accumulated predominantly in the root with a content of 112 µg/g at day 15, hindered the

25

development and growth of wheat plants, disrupted root structure and cellular ultrastructure, and

26

promoted oxidative stress. The GO that accumulated in the root showed extremely limited

27

translocation to the stem and leaves. During the experimental period, GO was excreted slowly

28

from the root. GO inhibited the germination of wheat seeds at high concentrations (≥0.4 mg/mL).

29

The mechanism of GO toxicity to wheat may be associated with oxidative stress induced by GO

30

bioaccumulation, reflected by the changes of malondialdehyde concentration, catalase activity

31

and peroxidase activity. The results demonstrate that 13C labeling is a promising method to

32

investigate environmental impacts and fates of carbon nanomaterials in biological systems.

33

ACS Paragon Plus Environment

2

Page 3 of 29

34

Environmental Science & Technology

Table of Contents Graphic

35

36 37

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 29

38

INTRODUCTION

39

Graphene and its derivatives have attracted considerable recent interest owing to their unique

40

structure and remarkable physical and chemical properties.1,2 Of particular interest, graphene

41

oxide (GO) is a graphene sheet with carboxylic groups at the edges and phenol hydroxyl/epoxide

42

groups on the basal plane, and exhibits excellent water dispersion and amphiphilic characteristics.

43

Graphene materials have diverse applications in a wide variety of fields, including electronics,3

44

energy,4 mechanics,5 advanced materials,6 biomedicine,7 environmental remediation8,

45

biosensing,9 and agriculture.10 For example, graphene-based touch screen for cell phones is

46

produced in Chongqing, China. Graphene rechargeable batteries are manufactured in several

47

countries. Theranostic applications of GO are reported, such as drug/gene delivery, biosensing,

48

bioimaging and a scaffold for cell growth.7 Graphene adsorbents have been developed for the

49

remediation of polluted water.8,11,12 Graphene quantum dots stimulated the growth of coriander

50

and garlic plants.10 On the basis of these innovations, there is growing demand for graphene

51

products, which has stimulated the large-scale production of graphene for industrial applications.

52

Several production lines are operational with an annual production capacity of several hundred

53

tonnes. Generally, graphene nanomaterials are readily released into the environment and may

54

lead to potential nanotoxicity and environmental risks.13-15 Potential exposure to graphene

55

products continues to increase and the hazards of exposure to graphene require thorough

56

investigation.16–18

57

Graphene nanomaterials are considered to be emerging environmental pollutants, and their

58

adverse effects on biota have received increasing attention.19–23 Recent studies reveal that

59

graphene induces growth inhibition, cell death, oxidative stress and morphological changes in

60

plants.24 Begum et al. observed that GO induces cell death in Arabidopsis thaliana, and water-

ACS Paragon Plus Environment

4

Page 5 of 29

Environmental Science & Technology

61

soluble graphene is phytotoxic to cabbage (Brassica oleracea var. capitata), tomato

62

(Lycopersicon esculentum), red spinach (Amaranthus tricolor L.), and lettuce (Lactuca sativa) at

63

0.5~2.0 mg/mL.25,26 GO does not influence the germination and development of A. thaliana at

64

concentrations lower than 1.0 mg/L.27 Graphene inhibits the germination and growth of faba

65

bean (Vicia faba L.) via induction of oxidative damage at 0.1~1.6 mg/mL.28 In tomato, graphene

66

penetrates the seed coat to accelerate germination and stimulates stem and root elongation of

67

seedlings, but inhibits biomass accumulation at 0.04 mg/mL.29 GO inhibits root elongation of

68

Brassica napus and regulates the concentrations of abscisic acid (ABA) and indole-3-acetic acid

69

(IAA) at 0.025~0.1 mg/mL.30 Under co-exposure to graphene and pollutants, GO amplifies the

70

toxicity of arsenic in wheat (Triticum aestivum) at GO concentrations of 0.1~10.0 mg/L.31

71

Bioaccumulation of nanomaterials in plants is essential to fully understand their biological

72

behavior and ecotoxicity.32 Bioaccumulation data guide nanotoxicity evaluations on plant tissues

73

and shed light on the toxicological mechanisms. The bioaccumulation of diverse nanomaterials

74

in plants has been widely investigated.33-35 The exposure concentration, particle size, surface

75

charge, particle dissolution, surfactant/dissolved organic matter (DOM), and plant species are

76

important factors that affect the bioaccumulation and translocation of nanomaterials in plant

77

tissues.36-40 Nevertheless, quantification of nanomaterials in vivo remains a challenge.

78

Quantification of graphene is more difficult than that of metal-containing nanoparticles, where

79

the total element present can be digested and quantified by atomic emission spectrometry or

80

inductively coupled plasma–mass spectrometry. Plant tissues contain high background quantities

81

of carbon that interfere with direct quantification of graphene in biological samples.41 Several

82

reports have used transmission electron microscopy (TEM) and Raman spectroscopy to detect

83

the presence/absence of graphene in plant tissues. For example, Zhang et al. used TEM and

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 29

84

Raman spectroscopy to confirm that graphene penetrates the seed coat and seedling root tip

85

cells.29 Black spots in the TEM images were assigned as GO, and the D-band and G-band signals

86

were observed in the seeds according to Raman spectra. Zhao et al. used TEM to investigate the

87

accumulation and translocation of GO.27 GO was recognized as black spots in root, leaf and

88

cotyledon cells.

89

13

C-labeling of the carbon skeleton is a well-established approach for quantification of

90

carbon nanomaterials in biological samples by analyzing the 13C/12C ratio with isotope ratio mass

91

spectrometry (IRMS).41–45 The 13C-labeling of the carbon skeleton does not damage the stability

92

and intrinsic structure of carbon nanomaterials 46–48 and enables the properties of carbon

93

nanomaterials to be traced in biological systems. The technique has the advantage of overcoming

94

the drawbacks of radioactive labeling, such as ready detachment, generation of radioactive waste,

95

and requirement for specific experimental approval.42 Herein, we prepared 13C-labeled GO for

96

bioaccumulation analysis and evaluated the toxicity of GO to wheat with reference to the

97

bioaccumulation data. The bioaccumulation of 13C-GO was quantified by IRMS. The effects of

98

GO on germination, growth, root elongation and oxidative stress of wheat were determined. The

99

implications for quantification of graphene in biological systems and evaluation of the safety of

100

graphene are discussed.

101 102 103

MATERIALS AND METHODS Preparation of 13C-Labeled GO and Unlabeled GO. 13C-Labeled graphite was prepared

104

by arc discharge method and oxidized by the modified Hummers method to produce 13C-GO.

105

The details of preparation protocols were given in the Supporting Information. Unlabeled GO for

ACS Paragon Plus Environment

6

Page 7 of 29

Environmental Science & Technology

106

toxicity evaluations was prepared by the modified Hummers method following our previous

107

reports.21 The 13C content of 13C-GO was determined by IRMS (Delta V Advantage, Thermo,

108

Bremen, Germany). Plant morphology was investigated by TEM (JEM-200CX, JEOL, Tokyo,

109

Japan) and atomic force microscopy (AFM; SPM-9600, Shimadazu, Kyoto, Japan). The

110

chemical states of elements were analyzed by X-ray photoelectron spectroscopy (XPS; Kratos,

111

Manchester, UK). The functional groups were identified by infrared spectroscopy (IR; Nicolet

112

Avatar 370, Thermo, Madison, WI, USA). The crystallinity was characterized by X-ray

113

diffraction spectroscopy (XRD, D/MAX 2000, Rigaku, Tokyo, Japan). The hydrodynamic radii

114

were measured on a nanosizer (Zetasizer 3000 HS, Malvern, Malvern, UK).

115

Plant Cultivation and GO Exposure. Seeds of transgenic wheat ‘Lunxuan 987’ were

116

obtained from the Institute of Crop Sciences, Chinese Academy of Agricultural Sciences, Beijing.

117

The cultivar ‘Lunxuan 987’ was chosen because it is high yielding, nutritious, and resistant to

118

biotic/abiotic stresses. Seedlings were cultured in modified Hoagland nutrient solution as

119

described in our previous report.42 The same protocol was adopted for plant culture here, because

120

it allowed the direct comparison between fullerenol and GO and could reflect the regulative

121

effects of size, shape and oxidation degree on the nano-biosafety. Wheat seeds were soaked in 15%

122

NaCl for 30 min, then twice soaked for 15 min in water. For the germination assay, 30 seeds

123

were placed on filter paper in a Petri dish (diameter 9 cm). Hoagland nutrient solution containing

124

GO (0–2.0 mg/mL) was introduced to the Petri dish to moisten the filter paper. The seeds were

125

incubated in an incubator in the dark at 25 °C and 80% relative humidity for 7 d. The germination

126

frequency was determined daily. Germination potential was calculated with Equation 1.

127

Germination potential=

(germinated seeds at day n+1)-(germinated seeds at day n) total seeds

(1)

ACS Paragon Plus Environment

7

Environmental Science & Technology

128

After germination, sets of three seedlings of uniform size were transferred to 100 mL

129

beakers containing GO at 0, 0.04, 0.2, 0.4, 0.8, 2.0 mg/mL (three replicate beakers per

130

concentration). The seedlings were cultivated under a day/night cycle of 12 h/12 h with

131

temperatures of 25/20°C, illumination of 24 000 lx during the day cycle, and 80% relative

132

humidity. Fresh Hoagland nutrient solution was added as necessary to maintain the volume of

133

100 mL. The seedlings were harvested at 15 d for toxicity evaluation.

134

Bioaccumulation of 13C-GO. To quantify the bioaccumulation of GO in wheat, wheat

135

seedlings (3 d post-germination) were exposed to 13C-GO (1.0 mg/mL) in Hoagland nutrient

136

solution. The seedlings were cultured by the aforementioned procedure and harvested at 7, 11,

137

and 15 d for IRMS measurements. The sampled seedlings were divided into the roots, stems

138

(here, stems referring to the whole seedlings except roots and leaves), and leaves, carefully

139

washed with deionized water for three times, lyophilized and ground into powder. Before

140

lyophilization, the root samples were examined with a scanning electron microscope (SEM, S-

141

4800, Hitachi, Japan) and Raman spectrometer (inVia, Renishaw, Wotton-under-Edge, UK) to

142

ensure the full removal of adsorbed 13C-GO. The 13C content in wheat samples was determined

143

by IRMS and expressed as δ values relative to the 13C content of the Vienna Pee Dee Belemnite

144

(VPDB) standard. The δ value was converted to GO concentration and percentage of exposed

145

dose per gram (%ED/g) following our previous report.42

146

Page 8 of 29

Toxicity of GO to Wheat. During the toxicity evaluations, wheats were exposed to GO at

147

concentrations of 0, 0.04, 0.2, 0.4, 0.8, 2 mg/mL, which covered the low and high concentrations

148

used in the literature.25-31 After harvesting at 15 d, the root, stem, and leaf lengths were measured

149

with a vernier caliper. The root, stem, and leaf fresh weights were recorded. After drying for 12 h

150

at 90°C, the dry weight of the samples was recorded.

ACS Paragon Plus Environment

8

Page 9 of 29

151

Environmental Science & Technology

Separately, fresh root samples harvested at 15 d were cut into small sections and fixed with

152

formaldehyde-acetate-alcohol solution. The samples were embedded in paraffin and sections of

153

10 µm thickness were stained with safranin and fast green. Images of the wheat paraffin sections

154

were captured under a microscope (CAB-30PC, Cabontek Co., Chengdu, China). Additional sets

155

of fresh root samples were embedded in tissue-freezing medium, frozen at −80 °C, cut into

156

sections of 20–30 µm thickness with a cryomicrotome (Leica CM 1850, Nussloch, Germany).

157

The sections were transferred onto cover glasses, stored at −20 °C until they were coated with

158

gold for 5 s using a sputter coater (E-1045, Hitachi, Tokyo, Japan), and observed with a SEM (S-

159

4800, Hitachi, Tokyo, Japan). A third set of samples was fixed with 3% glutaraldehyde, post-

160

fixed in 1% osmium tetroxide, dehydrated in a graded alcohol series, and embedded in epoxy

161

resin. Sections were cut with an ultramicrotome and post-stained with uranyl acetate and lead

162

citrate for TEM examination.

163

For oxidative stress assays, all kits were obtained from the Nanjing Jiancheng

164

Bioengineering Institute, Nanjing, China. Root samples harvested at 15 d were homogenized in

165

water (0.1 g tissue/1 mL water). The samples were centrifuged at 3000 rpm for 5 min to remove

166

the residues. The protein concentration in the supernatant was determined by staining with

167

Coomassie brilliant blue. The malondialdehyde (MDA) concentration and activities of

168

peroxidase (POD) and catalase (CAT) were analyzed following the manufacturer’s instructions

169

with an ultraviolet-visible spectrophotometer (UV-1800, Mapada, China). The protocols could

170

be downloaded at http://elder.njjcbio.com/index_en.php.

171 172

RESULTS AND DISCUSSION

ACS Paragon Plus Environment

9

Environmental Science & Technology

173

Page 10 of 29

Characterization of 13C-Labeled GO. The AFM image of 13C-GO is shown in Figure 1A.

174

The two-dimensional structure of the graphene sheets was identified and the height of the 13C-

175

GO layer was about 0.9 nm, consistent with the typical layer height of GO. The TEM image

176

further confirmed the sheet-like structure (Figure 1B). During the sampling, the 13C-GO sheets

177

were slightly folded and stacked, and thus their appearance differed from that under AFM.

178

According to the dynamic light scattering measurement, the hydrodynamic radius was 30 nm.

179

The IR spectrum confirmed the abundance of oxygen-containing groups (Figure 1C). The peak at

180

3382 cm−1 indicated the presence of –OH/-COOH groups. The C=O bonds were reflected by the

181

peak at 1731 cm−1. C-O bonds were indicated by the peak at 1053 cm−1. The typical C=C signal

182

was detected at 1623 cm−1. The XPS analysis indicated that the elemental contents of 13C-GO

183

were 68.1 at% for C and 31.9 at% for O (Figure 1D). By analyzing the C1s XPS, the carbon

184

atoms were divided into three components, namely C-C (27.8%), C-O (35.1%) and C=O (5.2%).

185

The only 2θ angle of 10.4° for 13C-GO indicated the large layer distance of graphene layers and

186

similar crystiline structure to 12C-GO, which had a 2θ angle of 10.5°. On the basis of IRMS

187

analyses, about 7.1 at% of C atoms were 13C atoms, which was much higher than that observed

188

in normal 12C-GO (1.1 at%), indicating the successful incorporation of 13C into the carbon

189

skeleton of GO. Except for the elevated 13C content, the remaining characterization data were

190

similar to those of normal 12C-GO sheets (Figure S1), indicating that 13C-GO was successfully

191

prepared.

ACS Paragon Plus Environment

10

Page 11 of 29

Environmental Science & Technology

192 193

Figure 1. Characterization of 13C-labeled GO. (A) AFM image; (B) TEM image; (C) IR

194

spectrum; (D) C1s XPS spectrum. The inset in (B) shows the magnified edge of 13C-GO.

195 196

Bioaccumulation of 13C-GO in wheat. The 13C contents of root, stem, and leaf samples

197

were determined by IRMS after exposure of the wheat seedlings to 13C-GO during the growth

198

period. Meaningful 13C abundance increases were observed in root samples comparing to the

199

root samples of the control group. Instead, no meaningful increases were found in stems and

200

leaves. The root uptake of 13C-GO was also supported by the SEM and TEM observations of

201

xenobiotic materials in roots (Figure S2). In particular, the root samples were carefully washed to

202

remove adsorbed GO before quantification. The SEM images and Raman spectra confirmed the

203

complete removal of adhered sheets (Figure S3). These results suggested that 13C-GO was

204

readily absorbed by the roots rather than attaching on the root surface, but the absorbed 13C-GO

205

showed limited translocation to the stem and leaves. The 13C-GO concentration in the root at day

206

7 was 283 µg/g, equivalent to 0.71% ED/g. The bioaccumulation of 13C-GO was much lower

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 29

207

than that of fullerenol (C60-OH), which accumulated at ~7% ED/g in wheat roots. This difference

208

might be due to that the large size of GO sheets hindered the penetration across biological

209

barriers. The small size of the C60-OH molecule may have promoted its uptake and translocation

210

in wheat. In addition, the high oxidation degree of GO might contributed to the root uptake.

211

Larue et al. quantified the lowly oxidized 14C-carbon nanotubes (CNTs; containing only 1.64 at%

212

oxygen) in wheat.49 Less than 0.0005% of applied CNTs were absorbed by the wheat root, and

213

about 200 µg/kg of gum Arabic-suspended CNTs and 43 µg/kg of humic acid-suspended CNTs

214

were detected in wheat leaves.

215

Over the time-course of the experiment, a trend for decreasing 13C-GO concentration in the

216

root was observed (112 µg/g at day 15), whereas the 13C-GO concentrations in the stem and

217

leaves increased (Please note that the differences between exposed and control groups were not

218

statistically significant; Figure 2). The mechanism for the decrease in 13C-GO concentration in

219

the root was unclear. We speculated that there might be three possibilities. The first mechanism

220

might be excretion directly via the root. As xenobiotic materials, the root might excrete 13C-GO

221

in root exudates. The second mechanism may be translocation to the stem and leaf, which is

222

predicted to be extremely slow on the basis of the present 13C measurements. The third

223

possibility is that 13C-GO was metabolized into CO2 and released to the atmosphere. Degradation

224

of GO is reported in the literature. For example, Lalwani et al. used lignin peroxidase to catalyze

225

the decomposition of H2O2 for degradation of graphene.50 Similarly, a myeloperoxidase–H2O2

226

system may also decompose GO.51 Girish et al. observed the in vivo degradation of graphene

227

over a period of 3 months, in which a macrophage played an important role in the degradation.52

228

Nevertheless, the degradation of GO in a biological system is expected to be slow, thus the total

229

contribution of degradation to the decreased concentration of 13C-GO is likely to be small.

ACS Paragon Plus Environment

12

Page 13 of 29

Environmental Science & Technology

230 231

Figure 2. Bioaccumulation of 13C-GO in wheat seedlings exposed to 13C-GO (1.0 mg/mL) in

232

Hoagland nutrient solution for 15 days. * p < 0.05 compared with the control group (n=3).

233 234

Toxicity of GO to wheat. In our study, GO showed toxicity to wheat seeds during the

235

germination, reflected by the decreased germination frequency and germinability (Figure

236

S4&S5). For normally germinated seedlings, GO also adversely affected biomass accumulation

237

and elongation in wheat seedlings (Table S1). The gain in fresh and dry biomass of wheat

238

seedlings was not significantly influenced at GO concentrations of 0.4 mg/mL and lower (Figure

239

3 A&B). The dry weight was suppressed by GO to a greater degree than was fresh weight,

240

implying that the water content of wheat seedlings was enhanced upon exposure to a high

241

concentration of GO. Interestingly, seedling length showed a trend to increase with GO

242

concentration from 0.04 to 0.4 mg/mL (although the differences were not statistically significant),

243

but decreased at higher concentrations of GO (Figure 3C). Consequently, the seedling was longer

244

and thinner with increasing GO concentration (Figure S6). Similarly, Zhang et al. reported that

245

graphene enhanced the stem length of tomato seedlings, but fresh biomass accumulation was

246

inhibited.29 These authors assigned the toxicity to penetration of vacuoles in root cells by

247

graphene. Zhang et al. observed that root and leaf elongation in wheat was stimulated by

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 29

248

graphene at concentrations from 250 to 1500 mg/L, but root hair development was suppressed.53

249

After 30 d exposure, root elongation was still significantly enhanced, whereas shoot length was

250

not. The shoot fresh weight was decreased in response to exposure to graphene for 30 d. The

251

toxicity of GO is concentration dependent and also species dependent. Zhao et al. reported that

252

the growth of A. thaliana was not influenced by GO at 1.0 mg/L and lower concentrations.27

253

However, GO inhibits root and shoot growth of cabbage, tomato and red spinach.26

254 255

Figure 3. Influence of GO on fresh weight (A), dry weight (B), seedling length (C) and root

256

length (D) of wheat seedlings. * p < 0.05 compared with the control group (n=10).

257

Damage to wheat root. The root number decreased (Figure S7) and the root elongation was

258

inhibited upon exposure to GO (Figure 3D). Inhibition of root development might be associated

259

with IAA and ABA.30 Cheng et al. reported that GO inhibited root growth of Brassica napus via

ACS Paragon Plus Environment

14

Page 15 of 29

Environmental Science & Technology

260

reduction in IAA concentration, which is the best-characterized molecular signal for root system

261

architecture and growth.30 Up-regulation of ABA also contributed, because a high concentration

262

of ABA inhibited seedling growth. The inhibition of root development obviously would hinder

263

the absorption of inorganic salts, water and other nutrients, and thereby affect seedling growth.

264

Sections of wheat roots were examined to reveal structural changes. The intact structure of

265

the root was observed in the control group. The epidermis, cortex, endodermis and xylem vessels

266

were clearly observed (Figure S8 A&B). GO induced damage to the cortex at low and high

267

concentrations (Figure S8 C–F). Aqtocytolysis occurred in the cortex to form aerenchyma-like

268

tissue, which increased the porosity of the cortex. The GO-induced structural damage to the

269

wheat root may render transportation of air to stems and leaves more difficult, thus increased

270

porosity was required to maintain the breath of the roots.

271

The ultrastructure of wheat roots was investigated by TEM to further examine the damage

272

to root cells induced by GO. The control group showed typical cellular ultrastructure. The cell

273

membrane was tightly appressed to the cell wall. The nucleus had a distinct nuclear membrane

274

and nucleolus. Several empty vacuoles were observed. Upon exposure to 0.4 mg/mL GO, many

275

vacuoles containing xenobiotic materials were observed, which we speculated to be GO or a GO-

276

biomolecule complex. Slight detachment of the cell membrane and cell wall was observed, but

277

the cell membrane and cell wall remained almost intact (Figure 4 A&B). These observations

278

suggested that the uptake of GO at 0.4 mg/mL in the 13C quantification experiment was likely the

279

result of endocytosis, rather than penetration via a damaged cell membrane and cell wall. Severe

280

ultrastructural damage was observed in response to exposure to GO at 2.0 mg/mL. The typical

281

nucleus structure was lost and detachment of the cell membrane and cell wall was obvious,

282

indicating shrinkage of cells (Figure 4 C&D). Under higher magnification, many vacuoles were

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 29

283

evident in the cell and it was impossible to distinguish the disrupted cellular components and GO

284

particles. The cell membrane was also severely damaged. In one cell we observed chromatin

285

condensation. The nuclear membrane was diffuse and the cytoplasm was divided into multiple

286

components by membrane-like structures. Again, we could not distinguish GO particles from the

287

disrupted cellular organelles. The structural damages induced by GO at high concentrations

288

indicated the toxicity of GO to root cells. The structural changes of plant root induced by

289

graphene were also reported in the literature.54

ACS Paragon Plus Environment

16

Page 17 of 29

Environmental Science & Technology

290 291

Figure 4. Ultrastructural changes in wheat root upon exposure to GO. (A) Control group; (B)

292

wheat root exposed to 0.4 mg/mL GO; (C, D) wheat root exposed to 2.0 mg/mL GO. The boxes

293

represent higher magnification images of portions of images C and D.

ACS Paragon Plus Environment

17

Environmental Science & Technology

294

Page 18 of 29

The possible mechanism of GO toxicity to roots might involve oxidative stress. The MDA

295

concentration significantly decreased at GO concentrations of 0.04 and 0.2 mg/mL, indicating

296

the absence of oxidative stress (Figure 5A). However, the MDA concentration markedly

297

increased at GO concentration of 0.8 mg/mL, suggesting the oxidative stress in wheat root. The

298

decline in MDA concentration at the GO concentration of 2.0 mg/mL may be attributable to

299

acute toxicity and destruction of cellular function. Activity of CAT was significantly increased

300

only with exposure to GO at 1.0 mg/mL (Figure 5B). Activity of POD was indicated to be more

301

sensitive, with significantly increased activity observed at 0.04 and 0.2 mg/mL and decreasing

302

activity at 0.4 mg/mL and higher concentrations (Figure 5C). Oxidative stress is widely

303

acknowledged to be a toxicological mechanism for nanomaterials in diverse biological systems.

304

Oxidative stress in plants in response to exposure to graphene has been observed previously. Ren

305

et al. attributed the toxicity of sulfonated graphene to oxidative damage.54 GO induces an

306

increase in reactive oxygen species concentration in leaves of cabbage, tomato and red spinach at

307

0.5–2.0 mg/mL.26 Consistent with these changes, accumulation of H2O2 is observed in leaves.

308

Oxidative damage leads to the death of root cells and membrane leakage in leaf cells. Anjum et

309

al. reported that GO induced oxidative stress in faba bean at 0.2–1.6 mg/mL.28 Cheng et al.

310

observed oxidative stress in B. napus exposed to GO at 0.025–0.1 mg/mL.30 No significant

311

increase in MDA concentration was observed, but activities of superoxide dismutase (SOD),

312

POD and CAT increased, indicating that exposure to GO represents a form of stress for plants.

313

Zhang et al. reported oxidative stress in wheat induced by exposure to GO at 0.25–1.5 mg/mL

314

for 48 h, with O2.− and MDA concentrations and SOD activity significantly increased. 53 Some

315

exceptions have been reported. Zhao et al. did not observe generation of reactive oxygen species

316

in A. thaliana seedlings, and the MDA and H2O2 concentrations and SOD and CAT activities

ACS Paragon Plus Environment

18

Page 19 of 29

Environmental Science & Technology

317

remained unchanged, in response to GO exposure at 10–1000 µg/L.27 Hu et al. reported that 0.2

318

mg/mL of hydrated graphene ribbon suppressed the oxidative stress of wheat seedlings.55

319

However, overall, oxidative damage might be an important mechanism for GO-induced toxicity

320

to wheat roots.

321 322

Figure 5. Oxidative stress in wheat roots exposed to GO. (a) Malondialdehyde concentration; (b)

323

catalase activity; (c) peroxidase activity. * p < 0.05 compared with the control group (n=10).

324 325

In summary, we utilized stable isotope labeling for quantification of graphene in biological

326

system, and demonstrated that 13C-GO accumulated in wheat roots at low concentrations. The

327

large size of 13C-GO hinders its translocation to stem and leaves. The direct contact and

328

bioaccumulation of GO in the root inhibited development of the root system, altered the root

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 29

329

structure and cellular ultrastructure, and led to inhibition of seedling growth. The mechanism of

330

GO toxicity to wheat roots may be oxidative stress. The present study provides a novel method

331

for quantitatively tracing of GO in biosafety assessments and presents systematic information on

332

the toxicity of GO to wheat. In future, this labeling technology could be used to investigate the

333

nanotoxicity, the mechanisms, metabolism and transmission of graphene and its derivatives to

334

evaluate long-term environmental health impacts. The quantifications and toxicity evaluations of

335

different carbon nanomaterials using similar protocols would benefit the understanding of the

336

unique nano-bioeffects, such as the shape effect, size effect and surface effect. We envisage that

337

the results will benefit the the ongoing environmental risk assessments of graphene

338

nanomaterials.

339 340

ASSOCIATED CONTENT

341

Supporting Information

342

The Supporting Information includes the preparation protocol of 13C-GO, the characterization of

343

unlabeled GO (Figure S1), the identification of GO in root by SEM and TEM (Figure S2), SEM

344

images and Raman spectra of washed root surface (Figure S3), the influence of GO on

345

germination (Figure S4), the photographs of germinated wheat seeds (Figure S5), the

346

photographs of wheat seedlings (Figure S6), the root numbers of wheat seedlings (Figure S7), the

347

structural changes of wheat roots (Figure S8), the detailed data of Figure 3 (Table S1).The

348

Supporting Information is available free of charge on the ACS Publications website at DOI: .

349

ACS Paragon Plus Environment

20

Page 21 of 29

Environmental Science & Technology

350 351

AUTHOR INFORMATION

352

Corresponding Authors

353

* Telephone: +86-28-85522269, Fax: +86-028-85524382, Email: [email protected] (Prof. S.-T.

354

Yang)

355

* Telephone: +86-10-88236709, Fax: +86-10-88236456, Email: [email protected] (Prof. X.-L.

356

Chang)

357

Notes

358

The authors declare no competing financial interest.

359 360 361

ACKNOWLEDGMENTS We thank Prof. Xian Zhang at Key Lab of Urban Environment and Health, Institute of Urban

362

Environment, CAS for IRMS analyses. This work was supported by the National Program for

363

Support of Top-notch Young Professionals, the National Natural Science Foundation of China

364

(Nos. 11475194 and 11675189), Beijing Natural Science Foundation (No. 2152038), the

365

National Key Research and Development Program of China (2016YFA0201603), and the

366

Fundamental Research Funds for the Central Universities, Southwest Minzu University (No.

367

2016NZDFH01).

368 369

REFERENCES

370

(1) Marchesan, S.; Melchionna, M.; Prato, M. Wire up on carbon nanostructures! How to play a

371

winning game. ASC Nano. 2015, 9(10), 9441–9450.

ACS Paragon Plus Environment

21

Environmental Science & Technology

372

(2) Zheng, L.; Zheng, L.; Sun, H. Y.; Gao, C. Superstructured assembly of nanocarbons:

373

Fullerenes, nanotubes, and graphene. Chem. Rev. 2015, 115(15), 7046−7117.

Page 22 of 29

374

(3) Li, N.; Yang, G. Z.; Sun, Y.; Song, H. W.; Cui, H.; Yang, G. W.; Wang, C. X. Free-standing

375

and transparent graphene membrane of polyhedron box-shaped basic building units directly

376

grown using a NaCl template for flexible transparent and stretchable solid-state

377

supercapacitors. Nano. Lett. 2015, 15(5), 3195−3203.

378

(4) Cui, S. M.; Mao, S.; Lu, G. H.; Chen, J. H. Graphene coupled with nanocrystals:

379

Opportunities and challenges for energy and sensing applications. J. Phys. Chem. Lett.

380

2013, 4(15), 2441−2454.

381 382 383 384 385 386

(5) Young, R. J.; Kinloch, I. A.; Gong, L.; Novoselov, K. S. The mechanics of graphene nanocomposites: A review. Compos. Sci Technol. 2012, 72(12), 1459–1476. (6) Wang, M.; Duan, X. D.; Xu, W. X.; Duan X. F. Functional three-dimensional graphene/polymer composites. ACS. Nano 2016, 10(8), 7231−7247. (7) Chung, C.; Kim, Y. K.; Shin, D.; Ryoo, S. R.; Hong, B. H.; Min, D. H. Biomedical applications of graphene and graphene oxide. Acc. Chem. Res. 2013, 46(10), 2211–2224.

387

(8) Zhao, L. Q.; Yu, B. W.; Xue, F. M.; Xie, J. R.; Zhang, X. L.; Wu, R. H.; Wang, R. J.; Hu, Z.

388

Y.; Yang, S. T.; Luo, J. B. Facile hydrothermal preparation of recyclable S-doped graphene

389

sponge for Cu2+ adsorption. J. Hazard. Mater. 2015, 286, 449–456.

ACS Paragon Plus Environment

22

Page 23 of 29

Environmental Science & Technology

390

(9) He, S. J.; Song, B.; Li, D.; Zhu, C. F.; Qi, W. P.; Wen, Y. Q.; Wang, L. H.; Song, S. P.; Fang,

391

H. P.; Fan, C. H. A graphene nanoprobe for rapid, sensitive, and multicolor fluorescent

392

DNA analysis. Adv. Funct. Mater. 2015, 20(3), 453–459.

393

(10) Chakravarty, D.; Erande, M. B.; Late, D. J. Graphene quantum dots as enhanced plant

394

growth regulators: Effects on coriander and garlic plants. J. Sci. Food. Agric. 2015, 95(13),

395

2772–2778.

396

(11) Zhao, L. Q.; Dong, P. J.; Xie, J. R.; Li, J. Y.; Wu, L. X.; Yang, S. T.; Luo, J. B. Porous

397

graphene oxide-chitosan aerogel for tetracycline removal. Mater. Res. Express. 2014, 1(1),

398

015601.

399

(12) Wu, R. H.; Yu, R. H.; Liu, X. Y.; Li, H. L.; Wang, W. X.; Chen, L. Y.; Bai, Y. T.; Ming, Z.;

400

Yang, S. T. One-pot hydrothermal preparation of graphene sponge for the removal of oils

401

and organic solvents. Appl. Surf. Sci. 2016, 362, 56–62.

402 403 404 405 406 407 408 409

(13) Teo, W. Z.; Sofer, Z.; Sembera, F.; Janousek, Z.; Pumera, M. Cytotoxicity of fluorographene. RSC Adv. 2015, 5, 107158–107165. (14) Teo, W. Z.; Chng, E. L. K.; Sofer, Z.; Pumera, M. Cytotoxicity of halogenated graphenes. Nanoscale 2014, 6, 1173–1180. (15) Chng, E. L. K.; Sofer, Z.; Pumera, M. Cytotoxicity profile of highly hydrogenated graphene. Chem. Eur. J. 2014, 20, 6366–6373. (16) Hu, X. G.; Zhou, Q. X. Health and ecosystem risks of graphene. Chem. Rev. 2013, 113(5), 3815−3835.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 29

410

(17) Kang, S.; Mauter, M. S.; Elimelech, M. Microbial cytotoxicity of carbon-based

411

nanomaterials: Implications for river water and wastewater effluent. Environ. Sci. Technol.

412

2009, 43(7), 2648–2653.

413 414

(18) Seabra, A. B.; Paula, A. j.; Lima, R. D.; Alves, O. L.; Duran, N. Nanotoxicity of graphene and graphene oxide. Chem. Res. Toxicol. 2014, 27(2), 159−168.

415

(19) Mao, L.; Liu, C. L.; Lu, K.; Su, Y.; Gu, C.; Huang, Q. G.; Petersen, E. J. Exposure of few

416

layer graphene to limnodrilus hoffmeisteri modifies the graphene and changes its

417

bioaccumulation by other organisms. Carbon 2016, 109, 566−574.

418

(20) Xie, J. R.; Ming, Z.; Li, H. L.; Yang, H.; Yu, B. W.; Wu, R. H.; Liu, X. Y.; Bai, Y. T.;

419

Yang, S. T. Toxicity of graphene oxide to white rot fungus Phanerochaete chrysosporium.

420

Chemosphere 2016, 151, 324–331.

421

(21) Chang, Y. L.; Yang, S. T.; Liu, J. H.; Dong, E. Y.; Wang, Y. W.; Cao, A. N.; Liu, Y. F.;

422

Wang, H. F. In vitro toxicity evaluation of graphene oxide on A549 cells. Toxicol. Lett.

423

2011, 200(3), 201–210.

424 425 426 427

(22) Zou, X. F.; Zhang, L.; Wang, Z. W.; Luo, Y. Mechanisms of the antimicrobial activities of graphene materials. J. Am. Chem. Soc. 2016, 138(7), 2064−2077. (23) Hu, W. B.; Peng, C.; Luo, W. J.; Lv, M.; Li, X. M.; Li, D.; Huang, Q.; Fan, C. H. Graphenebased antibacterial paper. ACS Nano 2010, 4(7), 4317–4323.

ACS Paragon Plus Environment

24

Page 25 of 29

Environmental Science & Technology

428

(24) Lahiani, M. H.; Dervishi, E.; Ivanov, I.; Chen, J. H.; Khodakovskaya, M. Comparative study

429

of plant responses to carbon-based nanomaterials with different morphologies.

430

Nanotechnology 2016, 27(26), 0957–4484.

431

(25) Begum, P.; Fugetsu, B. Induction of cell death by graphene in Arabidopsis thaliana

432

(Columbia Ecotype) T87 cell suspensions. J. Hazard. Mater. 2013, 260(18), 1032–1041.

433

(26) Begum, P.; Ikhtiari, R.; Fugetsu, B. Graphene phytotoxicity in the seedling stage of

434 435 436

cabbage, tomato, red spinach, and lettuce. Carbon 2011, 49(12), 3907–3919. (27) Zhao, S. Q.; Wang, Q. Q.; Zhao, Y. L.; Rui, Q.; Wang, D. Y. Toxicity and translocation of graphene oxide in Arabidopsis thaliana. Environ. Toxicol. Pharm. 2015, 39(1), 145–156.

437

(28) Anjum, N. A.; Singh, N.; Singh, M. K.; Sayeed, I.; Duarte, A. C.; Pereira, E.; Ahmad, I.

438

Single-bilayer graphene oxide sheet impacts and underlying potential mechanism

439

assessment in germinating faba bean (Vicia faba L.). Sci. Total. Environ. 2014, 472(4), 834–

440

841.

441 442

(29) Zhang, M.; Gao, B.; Chen, J. J.; Li, Y. C. Effects of graphene on seed germination and seedling growth. J. Nanopart. Res. 2015, 17(2), 1–8.

443

(30) Cheng, F.; Liu, Y. F.; Lu, G. Y.; Zhang, Y. K.; Xie, L. L.; Yuan, Y. F.; Xu, B. B.

444

Graphene oxide modulates root growth of Brassica napus L. and regulates ABA and IAA

445

Concentration. J. Plant. Physiol. 2016, 193, 57–63.

446 447

(31) Hu, X. G.; Kang, J. Lu, K. H.; Zhou, R. R.; Mu, L.; Zhou, Q. X. Graphene oxide amplifies the phytotoxicity of arsenic in wheat. Sci. Rep. 2014, 4, 6122–6122.

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 29

448

(32) Zhao, J.; Wang, Z. Y.; White, J. C.; Xing, B. S. Graphene in the aquatic environment:

449

Adsorption, dispersion, toxicity and transformation. Environ. Sci. Technol. 2014, 48(17),

450

9995−10009.

451 452

(33) Miralles, P.; Church, T. L.; Harris, A. T. Toxicity, uptake, and translocation of engineered nanomaterials in vascular plants. Environ. Sci. Technol. 2012, 46(17), 9224−9239.

453

(34) Petersen, E. J.; Flores-Cervantes, D. X.; Bucheli, T. D.; Elliott, L. C. C.; Fagan, J. A.;

454

Gogos, A.; Hanna, S.; Kagi, R.; Mansfield, E.; Bustos, A. R. M.; Plata, D. L.; Reipa, V.;

455

Westerhoff, P.; Winchester, M. R. Quantification of carbon nanotubes in environmental

456

matrices: Current capabilities, case studies, and future prospects. Environ. Sci. Technol.

457

2016, 50 (9), 4587–4605.

458

(35) Tripathi, D. K.; Singh, S.; Singh, S.; Pandey, R.; Singh, V. P.; Sharma, N. C.; Prasadf, S.

459

M.; Dubeya, N. K.; Chauhan, D. K. An overview on manufactured nanoparticles in plants:

460

Uptake, translocation, accumulation and phytotoxicity. Plant Physiol. Biochem. 2017, 110,

461

2−12.

462

(36) Dimkpa, C. O.; McLean, J. E.; Martineau, N.; Britt, D. W.; Haverkamp, R.; Anderson, A. J.

463

Silver nanoparticles disrupt wheat (Triticum aestivum L.) growth in a sand matrix. Environ.

464

Sci. Technol. 2013, 47(2), 1082−1090.

465

(37) Zhai, G. S.; Gutowski, S. M.; Walters, K. S.; Yan, B.; Schnoor, J. L. Charge, size, and

466

cellular selectivity for multiwall carbon nanotubes by maize and soybean. Environ. Sci.

467

Technol. 2015, 49(12), 7380−7390.

ACS Paragon Plus Environment

26

Page 27 of 29

Environmental Science & Technology

468

(38) Xia, T.; Kovochich, M.; Liong, M.; Madler, L.; Gilbert, B.; Shi, H. B.; Yeh, J. I.; Zink, J. I.

469

Nel, A. E. Comparison of the mechanism of toxicity of zinc oxide and cerium oxide

470

nanoparticles based on dissolution and oxidative stress properties. ACS Nano 2008, 2(10),

471

2121–2134.

472

(39) Wang, Z. Y.; Li, J.; Zhao, J.; Xing, B. S. Toxicity and internalization of CuO nanoparticles

473

to prokaryotic alga Microcystis aeruginosa as affected by dissolved organic matter.

474

Environ. Sci. Technol. 2011, 45(14), 6032–6040.

475

(40) Torre-Roche, R. D. L.; Hawthorne, J.; Deng, Y. Q.; Xing, B. S.; Cai, W. J.; Newman, L. A.;

476

Wang, Q.; Ma, X. M.; Hamdi, H.; White, J. C. Multiwalled carbon nanotubes and C60

477

fullerenes differentially impact the accumulation of weathered pesticides in four agricultural

478

plants. Environ. Sci. Technol. 2013, 47(21), 12539–12547.

479

(41) Du, M. M.; Zhang, H.; Li, J. X.; Yan, C. Z.; Zhang, X.; Chang, X.-L. Bioaccumulation,

480

depuration and transfer to offspring of 13C-labeled fullerenols by Daphnia magna. Environ.

481

Sci. Technol. 2016, 50(19), 10421–10427.

482

(42) Wang, C. L.; Zhang, H.; Ruan, L. F.; Chen, L. Y.; Li, H. L.; Chang, X. L.; Zhang, X.; Yang,

483

S. T. Bioaccumulation of 13C-fullerenol nanomaterials in wheat. Environ. Sci.: Nano 2016,

484

3(4), 799–805.

485

(43) Chang, X. L.; Ruan, L. F.; Yang, S. T.; Sun, B. Y.; Guo, C.; Zhou, L. J.; Dong, J. Q.; Yuan,

486

H.; Xing, G. M.; Zhao, Y. L.; Yang, M. Quantification of carbon nanomaterials in vivo:

487

Direct stable isotope labeling on the skeleton of fullerene C60. Environ. Sci.: Nano 2014,

488

1(1), 64–70.

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 29

489

(44) Yang, S. T.; Guo, W.; Lin, Y.; Deng, X. Y.; Wang, H. F.; Sun, H. F.; Liu, Y. F.; Wang, X.;

490

Wang, W.; Chen, M.; Huang, Y. P.; Sun, Y. P. Biodistribution of pristine single-walled

491

carbon nanotubes in vivo. J. Phys. Chem. C. 2007, 111(48), 17761–17764.

492

(45) Liu, J. H.; Yang, S. T.; Wang, X.; Wang, H. F.; Liu, Y. M.; Luo, P. G.; Liu, Y. F.; Sun, Y.

493

P. Carbon nanoparticles trapped in vivo-similar to carbon nanotubes in time-dependent

494

biodistribution. ACS Appl. Mater. Interfaces 2014, 6(16), 14672−14678.

495

(46) Wang, C. L.; Ruan, L. F.; Chang, X.-L.; Zhang, X. L; Yang, S.-T.; Guo, X. H.; Yuan, H.; 13

496

Guo, C. B.; Shi, W. Q.; Sun, B. Y.; Zhao, Y. L. The isotopic effects of

497

carbon cage (C70) fullerenes and their formation process. RSC Adv. 2015, 5(94), 76949–

498

76956.

499

(47) Ruan, L. F.; Chang, X. L.; Sun, B. Y.; Guo, C. B.; Dong, J. Q.; Yang, S. T.; Gao, X. F.;

500

Zhao, Y. L.; Yang, M. Preparation and spectra of

501

2014, 59(10), 905–912.

502 503

C-labeled large

13

C-enriched fullerene. Chin. Sci. Bull.

(48) Wang, Z. Z.; Chang, X. L.; Lu, Z. H.; Gu, M.; Zhao, Y. L.; Gao, X. F. A precision structural model for fullerenols. Chem. Sci. 2014, 5(8), 2940–2948.

504

(49) Larue, C.; Pinault, M.; Czarny, B.; Georgin, D.; Jaillard, D.; Bendiab, N.; Mayne-

505

L’Hermite, M.; Taran, F.; Dive, V.; Carrière, M. Quantitative evaluation of multi-walled

506

carbon nanotube uptake in wheat and rapeseed. J. Hazard. Mater. 2012, 227-228(43), 155–

507

163.

508

(50) Lalwani, G.; Xing, W. L.; Sitharaman, B. Enzymatic degradation of oxidized and reduced

509

graphene nanoribbons by lignin peroxidase. J. Mater. Chem. B. 2014, 2(37), 6354–6362.

ACS Paragon Plus Environment

28

Page 29 of 29

Environmental Science & Technology

510

(51) Kurapati, R.; Russier, J.; Squillaci, M. A.; Treossi, E.; Ménard-Moyon, C.; Rio-Castillo, A.

511

E. D.; Vazquez, E.; Samorì, P.; Palermo, V.; Bianco, A. Dispersibility-dependent

512

biodegradation of graphene oxide by myeloperoxidase. Small 2015, 11(32), 3985–3994.

513

(52) Girish, C. M.; Sasidharan, A.; G. Gowd, S.; Nair, S.; Koyakutty, M. Confocal raman

514

imaging study showing macrophage mediated biodegradation of graphene in vivo. Adv.

515

Healthcare Mater. 2013, 2(11), 1489–1500.

516

(53) Zhang, P.; Zhang, R. R.; Fang, X. Z.; Song, T. Q.; Cai, X. D.; Liu, X. J.; Du, X. T. Toxic

517

effects of graphene on the growth and nutritional levels of wheat (Triticum aestivum L.):

518

short- and long-term exposure studies. J. Hazard. Mater. 2016, 317, 543–551.

519

(54) Ren, W.; Chang, H. W.; Teng, Y. Sulfonated graphene-induced hormesis is mediated

520

through oxidative stress in the roots of maize seedlings. Sci. Total. Environ. 2016, 572,

521

926–934.

522 523

(55) Hu, X. G.; Zhou, Q. X. Novel hydrated graphene ribbon unexpectedly promotes aged seed germination and root differentiation. Sci. Rep. 2014, 4, 3782.

524

ACS Paragon Plus Environment

29