Subscriber access provided by - Access paid by the | UCSF Library
Article
Ambient water and visible-light irradiation drive changes in graphene morphology, structure, surface chemistry, aggregation and toxicity Xiangang Hu, Ming Zhou, and Qixing Zhou Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es503003y • Publication Date (Web): 16 Feb 2015 Downloaded from http://pubs.acs.org on February 18, 2015
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
ACS Paragon Plus Environment
Environmental Science & Technology
1
Ambient water and visible-light irradiation drive changes in graphene
2
morphology, structure, surface chemistry, aggregation and toxicity
3 4
Xiangang Hu, Ming Zhou, Qixing Zhou*
5 6
Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education),
7
Tianjin Key Laboratory of Environmental Remediation and Pollution Control, College of
8
Environmental Science and Engineering, Nankai University, Tianjin 300071, China.
9 10
ABSTRACT
11
The environmental behaviors and risks of graphene have attracted considerable
12
attention. However, the fundamental effects of ambient water and visible-light
13
irradiation on the properties and toxicity of graphene remain unknown. This work
14
revealed that hydration and irradiation result in the transformation of large-sheet
15
graphene to long-ribbon graphene. The thickness of the treated graphene decreased,
16
and oxides were formed through the generation of singlet oxygen. In addition,
17
hydration and irradiation resulted in greater disorder in the graphene structure and in
18
the expansion of the d-spacing of the structure due to the introduction of water
19
molecules and modifications of the functional groups. Oxidative modifications with
20
two-stage (fast and low) kinetics enhanced the number of negative surface charges on
21
the graphene and enhanced graphene aggregation. The above property alterations
22
reduced the nanotoxicity of graphene to algal cells by reducing the generation of
23
reactive oxygen species, diminishing protein carbonylation and decreasing tail DNA.
24
A comparative study using graphene oxide suggested that oxidative modifications
25
could play an important role in inhibiting toxicological activity. This study provides a
26
preliminary approach for understanding the environmental behaviors of graphene and
27
avoids overestimating the risks of graphene in the natural environment.
28
Keywords: Nanotoxicity, Graphene, Visible light, Transformation, Algae
1
ACS Paragon Plus Environment
Page 2 of 29
Page 3 of 29
Environmental Science & Technology
29
INTRODUCTION
30
Graphene-related research has rapidly grown in a wide range of disciplines because of
31
its broad applications in physics, chemistry, biology, medicine, environmental
32
protection and manufacturing.1-4 Given the potential of environmental and human
33
exposure to graphene due to its versatile applications, scientists are directing more
34
attention towards its environmental fate and biosafety.5-8 Generally, the
35
physicochemical properties of graphene determine its environmental fate and risks.9,10
36
A fundamental understanding of how graphene properties are altered under
37
environmentally relevant conditions is important for scientifically evaluating
38
ecological risks and designing safe graphene products.8,11 However, the effects of
39
ambient water and visible-light irradiation on the morphology, structure, surface
40
chemistry and toxicity of graphene remain unknown. This lack of information can
41
lead to over- or under-estimations of graphene risks in the natural environment.
42
Water is ubiquitous in the natural environment. For example, surface water,
43
underground water and atmospheric water (humidity) exist in the environment. In
44
addition, water is commonly used in scientific research and industry. Therefore, the
45
interactions of nanomaterials with water (hydration) comprise a fundamental issue
46
that links various fields of research related to nanomaterials and their applications.
47
During long-term stirring and trace levels of ozone in water, sufficient C60O was
48
produced at the surfaces of the nC60 particles to allow the formation of a stable
49
suspension in water.12 At room temperature, water has been observed to
50
spontaneously induce phase and morphology transformations of nanotubes.13
51
However, the interactions of water with two-dimensional graphene under
52
environmentally relevant conditions remain largely unknown.
53
Sunlight-catalyzed redox reactions (photooxidation and photoreduction) may be
54
critical transformation processes that affect nanomaterial coatings, oxidation states,
55
the generation of reactive oxygen species (ROS), and persistence.14-17 The oxidation
56
and mineralization of fullerenes dispersed in water by natural sunlight may result in
57
the attenuation of carbon-based nanomaterials.14 Many nanomaterials are innately
58
photoactive (e.g., fullerenes and carbon nanotubes), potentially producing ROS when 2
ACS Paragon Plus Environment
Environmental Science & Technology
59
exposed to sunlight.15,16 In addition, oxidation and carboxylation of carbon nanotubes
60
by OH radicals have been observed.17 Nanotube oxidation increases the surface
61
charge of the carbon nanotubes and their stability against aggregation while
62
decreasing the hydrophobicity.17 The effects of natural visible-light irradiation on the
63
morphology, structure, surface chemistry and aggregation of graphene are unclear and
64
worthy of additional study.
65
The physico-chemical features (e.g., morphology, structure, functional group, size
66
distribution) of nanomaterials are relevant for their toxicity. Coatings and
67
functionalization reduced the in vivo toxicity of carbon nanotubes.18 The
68
shape-dependent toxicity of carbon nanotubes was reported, which affects their
69
bioavailability.19 The surface defects of carbon nanotubes can result in the generation
70
of ROS and induce toxic effects in vivo or in vitro.20 In addition, the aggregation state
71
can affect the shape and surface area of carbon nanotubes and is related to the
72
inhibition of algal growth.21 However, the influences of physico-chemical features on
73
graphene toxicity are poorly understood. Compacted graphene sheets are more
74
damaging to mammalian fibroblasts than less densely packed graphene oxides,
75
depending on the nanomaterial aggregation.22 When hydration or visible-light
76
irradiation alters the properties of graphene, the corresponding nanotoxicity likely
77
differs from that of the pristine graphene.
78
The primary objectives of this study were to determine how water and
79
visible-light irradiation alter the morphology, structure, surface chemistry,
80
aggregation and toxicity of graphene. Specifically, we determined the following: (i)
81
the morphology, structure, surface chemistry and aggregation of graphene before and
82
after treatment using various optical and microscopic techniques; (ii) their
83
nanotoxicities (oxidative stress, protein carbonylation, tail DNA, development
84
inhibition and ultrastructure damage) to a model species (C. Vulgaris) before and after
85
these treatments; and (iii) the correlations between the graphene properties and
86
nanotoxicity.
87 88
MATERIALS and METHODS 3
ACS Paragon Plus Environment
Page 4 of 29
Page 5 of 29
Environmental Science & Technology
89
Water and visible-light irradiation treatments
90
Graphene nanosheets were obtained from the Nanjing XFNANO Materials Tech Co.,
91
Ltd., China. Single-layer graphene was prepared using thermal exfoliation and
92
hydrogen reduction. Similar to other carbon nanomaterials, one of the principal
93
challenges associated with testing the environmental fate or ecotoxicological effects
94
of graphene is that carbon nanomaterials are not readily dispersed at detectable
95
concentrations without adding a surfactant and/or a dispersion process. Sonication is a
96
well-reported and relatively acceptable method for dispersing nanomaterials.23
97
Compared with probe sonication, ice-bath sonication at low temperatures and less
98
than 150 W has been widely used for graphene, carbon nanotubes and fullerene.23-25
99
In this study, graphene (0.02 g) was suspended in 200 mL of pure water (18.2 Ω/cm)
100
and subjected to ice-bath sonication for 40 min at 120 W. The same sonication
101
treatment was used for all treatments to minimize the occurrence of artifacts between
102
the treatments due to sonication. To prepare the irradiated graphene, the suspended
103
graphene was placed in a visible-light incubator maintained with an irradiation of 34
104
W/m2, a temperature of 24°C, and a relative humidity of 80%. To prepare the hydrated
105
graphene, the graphene suspension was covered with aluminum foil and placed in the
106
same incubator. Hydration and irradiation treatments were imposed for 120 days.
107
During treatment, pure water was added to maintain a constant volume of 150–200
108
mL. To stimulate perturbation, the treated samples were gently stirred for 10 min
109
twice each day at 150 rpm. To collect a sufficient amount of the treated graphene, ten
110
identical samples (100 mg/L, 200 mL per sample) were prepared. After hydration and
111
irradiation, the graphene suspensions were filtered through 0.1 µm polyether sulfone
112
resin membranes, and the materials on the membranes were lyophilized to form
113
powders for X-ray powder diffraction (XRD), X-ray photoelectron spectroscopy (XPS)
114
and other analyses.
115 116
Characterization of graphene before and after treatment (see Supporting
117
Information)
118 4
ACS Paragon Plus Environment
Environmental Science & Technology
119
Generation of free radicals and modification kinetics
120
The concentrations of furfuryl alcohol (FFA), which is used as an 1O2 trapping reagent,
121
were determined using high performance liquid chromatography (Waters, 2695, USA)
122
with ultraviolet detection (Waters, 2487, USA), as previously reported.14 The kinetics
123
of graphene photomodification were measured using the time profile of damage to the
124
pristine sp2 structure. The sp2 structure was indirectly quantized using the specific
125
adsorption of graphene at 270 nm and detected using a TU-1901 spectrophotometer
126
with UVWin5 software.
127 128
C. vulgaris cultivation
129
C. vulgaris and its culture medium (BG-11) were purchased from the Freshwater
130
Algae Culture Collection at the Institute of Hydrobiology in China. All materials that
131
came into contact with the microalgae were sterilized before use. The current
132
environmental concentrations of graphene remain unknown. To directly compare the
133
reported doses of carbon nanomaterials for the toxicity tests,26,27 graphene suspensions
134
were prepared using several different concentrations (0.1, 1.0 and 10 mg/L) in BG-11
135
culture medium adjusted to pH = 7.0. To avoid the effects of nanomaterial
136
aggregation and prepare an exact concentration, a 10 mg/L stock solution was
137
dispersed in BG-11 culture medium under ice-bath sonication for 40 min at 120 W
138
before immediately diluting to 1 mg/L. Similarly, the prepared 1 mg/L suspensions
139
were dispersed using sonication before immediately diluting to 0.1 mg/L. The
140
concentration of algal cells at the beginning of the test was 0.5 × 106 cells/mL. The
141
suspensions were shaken once every 8 h (150 rmp for 10 min per time) and placed in
142
a light incubator under 34 W/m2 irradiation and 80% humidity for 5 days at 24°C.
143 144
Reactive oxygen species (ROS)
145
To determine the generation of ROS in cells, an intracellular ROS indicator,
146
2′,7′-dichlorofluorescin diacetate (DCFH-DA), was used. After 5 days of exposure to
147
graphene, the algal cells were washed twice with medium and then incubated in 2 mL
148
of 10 mM DCFH-DA at 37°C for 30 min. The cells were subsequently washed three 5
ACS Paragon Plus Environment
Page 6 of 29
Page 7 of 29
Environmental Science & Technology
149
times with PBS buffer at pH = 7.4, and their fluorescence was determined at an
150
excitation of 485 nm and emission of 520 nm using a fluorescence spectrometer
151
(PerkinElmer LS 55, USA).
152 153
Cellular development and malondialdehyde content
154
The number and diameter of cells were counted using a cell analyzer
155
CASY-TT (Innovatis, Germany). The amounts of chlorophyll a and chlorophyll b
156
were quantified to determine algal photosynthesis. The lipid peroxide concentration
157
was determined in terms of the malondialdehyde (MDA) content, and the chlorophyll
158
a, chlorophyll b and MDA contents were analyzed using a TU-1901
159
spectrophotometer, as described elsewhere.28
160 161
Comet assay
162
The damage level of the DNA strands in the algal cells was determined using a comet
163
assay and the method described in our previous study.29 After electrophoresis, the
164
slides were washed three times with 0.5 M Tris buffer (pH 7.5) and the DNA was
165
stained using ethidium bromide. The stained slides were examined using a
166
fluorescence microscope (Zeiss, Axio Imager Z1, Germany) equipped with a CCD
167
camera. Three slides were prepared per group, and 50 randomly selected and
168
non-overlapping cells were analyzed on each slide. The images were analyzed using
169
the CASP software, and the percentage of tail DNA (% DNA) was measured as an
170
indicator of DNA damage.
171 172
Protein carbonyls and cellular ultrastructure measurements (see Supporting
173
Information)
174 175
Statistical analysis
176
All treatments included three replicates, and the error bars on the results represented
177
the mean ± SD (standard deviation). Differences were regarded as statistically
178
significant at P < 0.05. The data were analyzed using a one-way analysis of variance 6
ACS Paragon Plus Environment
Environmental Science & Technology
179
(ANOVA) and were compared using a post hoc Tukey’s test. All analyses were
180
performed using the IBM Statistics SPSS 19 software.
181 182
RESULTS and DISCUSSION
183
Morphological alterations
184
In this study, atomic force microscopy (AFM) and field-emission transmission
185
electron microscopy (FETEM) were conducted to study the morphology of the
186
graphene before and after treatment. As shown in Figure 1a, the pristine graphene had
187
a nanosheet morphology and was 0.872 nm thick (average 0.812 ± 0.071 nm, n = 6),
188
which was consistent with the 0.5–1.0 nm thickness reported for single-layer
189
graphene.30 Moreover, abnormal pores and dentate edges were observed on and
190
around the nanosheets, which represented the inherent defects that formed during their
191
synthesis.31 After treatment, the large-sheet morphology transformed into a
192
long-ribbon morphology. After treatment, the thicknesses of the hydrated and
193
visible-light-irradiated graphene decreased to 0.672 (average 0.652 ± 0.057 nm, n = 6)
194
and 0.356 nm (average 0.366 ± 0.028 nm, n = 6), respectively. Raised dots were
195
distributed on the treated graphene, as indicated by the black arrows shown in Figure
196
1a. The thicknesses of the raised dots reached approximately 12 and 4 µm for the
197
hydrated and irradiated graphene, respectively. As shown in Figure 1b, the EDX
198
analyses revealed that the chemical compositions of the dots included oxides, nitrides
199
and sulfides, which originated from the atmosphere during treatment. During
200
long-term hydration and irradiation, the carbon atoms located in the pores and on the
201
edges of the graphene defects possessed dangling bonds with unpaired electrons that
202
could drive the formation of oxides, nitrides and sulfides.32 Furthermore, transmission
203
electron microscopy (TEM) showed the morphological alterations of the graphene
204
before and after hydration and irradiation treatments. As shown in the left-hand
205
images of Figure 2, the treated graphene exhibited more wrinkles than the pristine
206
graphene. Graphene is known to exhibit severely wrinkled or folded structures on its
207
edges and planes due to a large number of defects and a high surface-to-volume ratio.
208
Scrolls and multiple folds can result in a number of dark lines, even in monolayer 7
ACS Paragon Plus Environment
Page 8 of 29
Page 9 of 29
Environmental Science & Technology
209
graphene (as observed on the graphene shown in the images on the right side of
210
Figure 2).33 The scrolls and multiple folds of pristine graphene exhibited an ordered
211
morphology. However, this ordered morphology became disordered in the hydrated
212
and irradiated graphene, as indicated by the red arrows in Figure 2. This result
213
suggested that the initial π-π stacking interactions were degraded due to hydration and
214
irradiation.
215 216
Structural alterations
217
Next, Raman spectrometry (RS) was used to analyze the structure of the graphene
218
before and after the hydration and irradiation treatments, as shown in Figure 3a.
219
Typical D and G bands of graphene were detected at approximately 1350 and 1594
220
cm-1. These bands reflect the disordered structure and ordered sp2 carbon system of
221
graphene, respectively.34 Other features, such as the 2D band and the combined-mode
222
D + G bands located at approximately 2689 and 2937 cm-1, respectively,35 were also
223
observed in the RS spectra. No obvious shifts in any of the aforementioned peaks
224
were observed after the treatments. The ratios of the D-band intensity to the G-band
225
intensity were 1.08, 1.39, and 1.24 for the graphene, hydrated graphene and irradiated
226
graphene, respectively, indicating that hydration and visible-light irradiation enhance
227
the disordered structure of graphene. These results are consistent with the TEM
228
images presented in Figure 2.
229
The structural alterations of the graphene were further investigated using XRD, as
230
shown in Figure 3b. The widths of the peaks were very large, and the half-peak widths
231
and crystalline grain sizes of the XRD peaks were 9.1° and 0.9 nm, respectively, in all
232
samples. The XRD patterns of the graphene, hydrated graphene and
233
visible-light-irradiated graphene contained peaks at 2θ = 22.9°, 20.2°, and 19.4°,
234
respectively. In contrast with the low-concentration single-layer graphene on the mica
235
plates of AFM or the copper meshes of TEM, graphene nanosheets (graphene
236
powders) interact with one another through intrinsic π electrons in XRD patterns.
237
Based on Braggs law, the d-spacing of graphene should be slightly larger than that
238
(0.34 nm) of graphite.36 The d-spacings of graphene, hydrated graphene and 8
ACS Paragon Plus Environment
Environmental Science & Technology
239
visible-light-irradiated graphene were calculated to be 0.39, 0.43 and 0.46 nm,
240
respectively. These data suggest that hydration and visible-light irradiation increased
241
the d-spacing of the graphene. According to a previous report, the relative humidity
242
and stacking configuration of graphene sheets are the primary factors that govern the
243
d-spacings of graphene-based nanomaterials.13 The hydrodynamic diameter of a water
244
molecule is 0.27 nm, and the introduction of a water monolayer results in 0.22–0.25
245
nm of lattice expansion,37 which is significantly larger than the difference between the
246
pristine and treated graphene. However, the wide peaks could be differentiated into
247
more than one small peak. For example, the peak from the pristine graphene consisted
248
of three small peaks at 2θ = 17.3°, 22.9° and 29.5°, with calculated d-spacings of 0.51,
249
0.39 and 0.30 nm, respectively. The peak from the hydrated graphene consisted of
250
three small peaks at 2θ = 12.2°, 17.0°, and 26.0°, with calculated d-spacings of 0.73,
251
0.52, and 0.34 nm, respectively. Given the changes in the d-spacings of the
252
differentiated peaks, a few water molecules most likely became incorporated between
253
the carbon atoms of the graphene. However, water molecules were not the only
254
factors that affected the lattice expansion. The lattice defects and the oxygen/nitrogen
255
functional groups reduced the d-spacings of the crystalline peaks.38,39 Therefore, the
256
crystal structures were a systemic result of the incorporation of water-molecules and
257
the introduction of defects and functional groups.
258 259
Alterations in surface chemistry
260
The Fourier-transform infrared (FTIR) spectra of the graphene before and after the
261
hydration and irradiation treatments are presented in Figure S1. In the pristine
262
graphene spectrum, the C=C band at 1600-1700 cm-1 was assigned to the skeletal
263
vibrations of un-oxidized graphene. This peak disappeared in the spectra of the treated
264
graphene. In addition, C=O, C-OH and C-O-C vibrations all contributed to the broad
265
peaks at 600-1100 cm-1,38,40 which suggested that the graphene was oxygenated after
266
the treatments. The peaks at approximately 1380 and 3290 cm-1 were assigned to C-N
267
and N-H vibrations,38,39 which indicated that the nitrogen groups were modified.
268
Changes in the surface chemistry of the graphene were further confirmed by XPS. 9
ACS Paragon Plus Environment
Page 10 of 29
Page 11 of 29
Environmental Science & Technology
269
The full spectra of the graphene before and after treatments are presented in Figure S2.
270
The pristine graphene spectrum consisted of 94.0% C1s and 6.0% O1s. The small
271
amount of oxygen in the pristine graphene resulted from the synthesis procedure.41
272
The spectrum of the hydrated graphene contained 92.7% C1s, 5.8% O1s, 1.3% N1s
273
and 0.2% S2p, and the spectrum of the irradiated graphene contained 90.2% C1s,
274
8.9% O1s, 0.7% N1s and 0.2% S2p. The EDX results reflected the presence of a small
275
amount of nitrogen and sulfur in the raised dots of the hydrated and irradiated
276
graphene. Considering that nitrogen and sulfur were not observed in the pristine
277
graphene, both elements most likely resulted from the ambient atmosphere. The
278
remarkable increase of O1s in the irradiated graphene implied that graphene can be
279
gradually oxidized using visible-light irradiation under an ambient atmosphere. The
280
specific compositions of the C1s and O1s signals were determined, as shown in
281
Figure S3. The C1s signal consisted of C=O and C-C, which was consistent with the
282
FTIR results presented in Figure S2. Unlike the O1s for the other graphene samples,
283
the O1s signal in the hydrated graphene had multiple components, including -OH
284
(37.7%), C-O (8.9%) and C-O-C (53.4%).
285
The UV-vis absorption spectra of the graphene before and after the hydration and
286
irradiation treatments are presented in Figure S4. The peak in the spectrum of the
287
pristine graphene at approximately 270 nm reflected the concentration of π electrons
288
and a structural ordering that was consistent with the presence of sp2 carbon.42 In
289
addition, the pristine graphene resulted in a peak at 230 nm due to the presence of
290
6.0% oxygen, as shown by the XPS results. After all treatments, the peak at 270 nm
291
shifted to approximately 230 nm, which is characteristic of the π–π* transition of the
292
C=C in graphene oxide.43 These results are consistent with graphene oxygenation
293
during treatment. The UV-vis absorption decreased remarkably between 250 and 800
294
nm in the treated graphene, which suggested that the electronic conjugation within the
295
pristine graphene sheets was disrupted.44 In addition, the modifications of oxygen and
296
nitrogen altered the d-spacings in the graphene.45 The d-spacing expanded in the
297
treated graphene, as shown in Figure 3b, and was linked to alterations in the surface
298
chemistry, as shown in Figure S3. In addition, the disordered structures evident in the 10
ACS Paragon Plus Environment
Environmental Science & Technology
299
FETEM images and RS spectra were consistent with the introduction of oxygen and
300
nitrogen groups. These results demonstrated that graphene can be oxidized, and its
301
aromatic structure can be rearranged in the natural environment.
302 303
Alteration of the dispersion state
304
The initial aggregation period was defined as the period between experimental
305
initiation and when the hydrodynamic diameter (Dh) was 1.5-fold less than its initial
306
value (1.5Dih).46 In this study, the initial aggregation rate was calculated for the period
307
between 0 and 300 s because the Dh of graphene exceeds 1.5Dih after 300 s. The initial
308
aggregation rates were obtained from Dh data fitting using a one-dimensional linear
309
regression equation and Origin 9.0 software.46 As shown in Figure S5, graphene was
310
unstable, and its size increased with time, yielding an initial aggregation rate constant
311
of 1.43 nm/s. The hydrated and irradiated graphene exhibited lower initial aggregation
312
rates (0.53 and 0.38 nm/s, respectively). During the nanomaterial agglomeration
313
period, the nanomaterials undergo rapid agglomeration and moderate agglomeration
314
before remaining relatively stable.47 However, the initial aggregation rate of graphene
315
was fast. Given that graphene suspensions were exposed to C. vulgaris for 5 days
316
(shaken once every 8 h at 150 rmp for 10 min) in the toxicological experiments, the
317
Dh values of the materials were detected on the 5th day under the shaking conditions
318
described above. The final Dh values of the graphene, hydrated graphene, and
319
irradiated graphene, were 2132, 1245 and 836 nm on the 5th day, respectively. The
320
aggregation results demonstrated that irradiation enhanced graphene aggregation,
321
which agreed with the increasing concentrations of oxygen-containing groups.
322
Furthermore, the ζ-potential results confirmed that irradiation improved the dispersion
323
of graphene. As shown in Figure S6, the ζ-potential became negative as the pH
324
increased, which suggested that the surfaces of the three materials exhibited negative
325
charges. Irradiation caused the graphene surface to become more negative, with
326
ζ-potentials ranging from −27 to −38 mV at pH 7−10.
327 328
Generation of free radicals and the kinetics of modification 11
ACS Paragon Plus Environment
Page 12 of 29
Page 13 of 29
Environmental Science & Technology
329
Previous research showed that aqueous colloidal dispersions of carbon nanotubes and
330
nC60 can generate ROS intermediates, including singlet oxygen (1O2), when exposed
331
to solar light.48 In theory, pristine graphene is chemically inert due to the giant
332
π-conjugation system. However, defects and impurities could occur during fabrication,
333
including the formation of nanopores and oxygen-containing groups, as shown in
334
Figures 1 and S3. These defects and impurities are highly photochemical active, likely
335
form free radicals and result in the photomodification of graphene.49 In this study, the
336
generation of 1O2 was investigated using the probe compound, FFA, as shown in
337
Figure S7. A remarkably higher FFA decay rate was obtained for the graphene under
338
irradiation relative to the graphene that was not irradiated. The FFA decay rate
339
constants from hydrated (without light) graphene and irradiated graphene were
340
0.0019/min and 0.022/min, respectively, according to common pseudo first-order
341
kinetics fitting for 1O2,48 which suggested that the irradiation triggered the generation
342
of 1O2.
343
Direct photoreactions generally exhibit first order kinetics.50 However, the kinetics
344
of photomodification exhibit two different stages before and after 24 h, fast and slow
345
photomodification, as shown in Figure S8. The two stages were fit using first order
346
kinetics and resulted in R2 values greater than 0.9. The fast and slow kinetics
347
constants were 6.9×10-3/h and 5.6×10-4/h for hydrated graphene, respectively. For
348
irradiated graphene, the fast and slow kinetic rate constants were 22.3×10-3/h and
349
13.5×10-4/h, respectively. These values implied the presence of at least two
350
components with different photoreactivities. Furthermore, XPS was used to detect
351
alterations of the chemical components during the fast and slow stages. The ratios of
352
C=O to C-OH during the fast (12 h irradiation) and slow stages (120 h irradiation)
353
were 0.71 and 0.12, respectively. These results demonstrated that the rate of graphene
354
photomodification decreased as the ratio of C=O to C-OH decreased.
355 356
Reduction of oxidative stress and DNA damage
357
Excessive ROS generation is a general phenomenon and the primary mechanism of
358
nanotoxicity.51-53 Excessive ROS generation will alter cellular membrane properties, 12
ACS Paragon Plus Environment
Environmental Science & Technology
359
damage DNA and reduce enzyme activity.54,55 As shown in Figure 4a, pristine
360
graphene resulted in the generation of more ROS than was observed in the control
361
without graphene exposure. However, hydration and irradiation inhibited the
362
generation of ROS by graphene. The concentrations of MDA reflected the lipid
363
peroxide concentrations.56 Hydration and irradiation reduced the enhancements of the
364
MDA content that were caused by graphene. The carbonyl-protein content is another
365
important indicator of oxidative stress.55 As illustrated in Figure 4, pristine graphene
366
increased the concentration of carbonyl proteins, and hydration and irradiation
367
reduced the protein carbonylation. The proportion of tail DNA reflects the level of
368
DNA damage. The pristine graphene triggered a significant increase in tail DNA
369
relative to the control, and the hydrated and irradiated graphene resulted in less tail
370
DNA. These results demonstrated that ambient hydration and visible-light irradiation
371
reduce the oxidative stress and DNA damage induced by graphene in algal cells.
372 373
Reductions of reproduction inhibition and damage to cellular structures
374
The effects of the hydration and irradiation treatments on algal reproduction and
375
development were investigated, as shown in Figure S9. Compared with the control,
376
the pristine graphene significantly reduced the reproduction of algal cells and the
377
concentrations of chlorophyll a and b, which are indicators of photosynthesis. The
378
hydration and irradiation of graphene reversed these adverse effects. Neither the
379
pristine graphene nor the treated graphene induced any changes in the diameters of the
380
algal cells, which indirectly indicated the agglomeration of the cells, as shown in
381
Figure S9. To explore the roles of chemical compositions on retarding toxicological
382
activities, a comparative study using surface modified graphene (graphene oxide) was
383
conducted. The properties of used graphene oxide are presented in detail in our
384
previous work.57 As shown in Figure S9, graphene oxides remarkably reduced the
385
adverse effects on the number of cells, cell diameter and chlorophyll content,
386
suggesting that oxidative modification may play an important role in decreasing
387
toxicological activity. 13
ACS Paragon Plus Environment
Page 14 of 29
Page 15 of 29
Environmental Science & Technology
388
The TEM images of the algal cells confirmed the positive effects of hydration and
389
irradiation, as shown in Figure S10. In the control cells, the cytoplasm membrane was
390
near the intact cell wall. A layered structure of thylakoids was visible in the
391
chloroplasts. In the cells that were exposed to pristine graphene, plasmolysis occurred
392
and the cell wall was ruptured, as indicated by double and single red arrows,
393
respectively. The structure of thylakoids was obscured, as indicated by the pink
394
arrows. In the cells that were exposed to the treated graphene, the cell walls remained
395
intact, and the occurrence of plasmolysis was reduced, especially in the case of the
396
irradiated graphene. Given that the effective diameter of the pores in the cell walls of
397
the algae did not exceed 3 nm, the large graphene nanosheets could not directly cross
398
the cell walls.58 However, the adsorption of graphene onto cell walls or its
399
incorporation into cell walls was observed, as indicated by the yellow arrows. The
400
observed alterations in the cell walls and chloroplasts are consistent with the
401
aforementioned results related to oxidative stress, chlorophyll content and cellular
402
reproduction. These data demonstrate that ambient hydration and visible-light
403
irradiation reduce the graphene-induced inhibition of reproduction and
404
cellular-structure damage of algae.
405 406
Correlations between nanomaterial properties and nanotoxicity
407
The structural defects of carbon nanotubes play a major role in causing toxicity, even
408
genotoxicity,20,59 largely because of their high activity. For pristine graphene,
409
numerous defects (pores and dentate edges) were observed on the nanosheets, as
410
displayed in the AFM image presented in Figure 1a. Generally, the graphene edges
411
are sharp with dangling bonds and are highly reactive to guest atoms or molecules.54
412
Reportedly, these sharp edges can directly rupture the membranes of cells, bacteria
413
and viruses, thereby inducing physical damage to living organisms.55,60,61 Thus, the
414
sharp edges of the pristine graphene caused damage to the cell walls, as illustrated in
415
Figure S10. By contrast, for the treated graphene, the sharp and dentate edges
416
disappeared, and the cell walls that were exposed to the treated graphene remained
417
intact. 14
ACS Paragon Plus Environment
Environmental Science & Technology
418
Morphology is another factor that affects graphene nanotoxicity.62 The nanosheets
419
observed in the pristine graphene were transformed into nanoribbons in the treated
420
graphene, as described in Figure 1a. Previous studies showed that graphene with a
421
nanoribbon morphology exhibits a lower phytotoxicity to wheat relative to graphene
422
nanosheets.28 Compared with the thickness of graphene, the thicknesses of the
423
hydrated and irradiated graphene decreased by 19.7% and 56.2%, respectively. This
424
decreased thickness reduced the stiffness and rigidity of the graphene, and the thinner
425
materials were readily deformed by weak forces, such as the water surface tension.33
426
The pathological response to fibers and carbon nanotubes increased with their
427
stiffness,63 and a similar phenomenon was observed for plate-like graphene in this
428
study. The UV-vis absorption spectra reflected the effects of light irradiation, and the
429
reduced visible-light absorption (as shown in Figure S4) suggested that the
430
photoreactivity of the treated graphene was diminished. These results suggest that
431
hydration and irradiation reduced the structural defects, rigidity, photoreactivity and
432
nanotoxicity of the graphene.
433
The irradiated graphene presented more surface negative charges and was more
434
stable than graphene and hydrated graphene, as shown in Figure S5. The negative
435
charges reduce the direct interactions with algal cells because the surfaces of the algal
436
cells are charged.64 In addition, irradiation reduced the aggregation of nanomaterials
437
due to electrostatic repulsion. The aggregation rate of the irradiated graphene was
438
smaller than that of the non-irradiated graphene. The aggregation state of graphene
439
can affect the shape, size and surface area of carbon nanomaterials. This effect is
440
more pronounced for highly agglomerated carbon nanotubes, which results in a
441
greater decrease in the overall DNA content than the better-dispersed carbon nanotube
442
bundles.65 Similarly, the irradiated graphene exhibited a lower toxicity than the
443
graphene that was not irradiated. One possible explanation for this result is that the
444
larger agglomerated materials were stiffer and more rigid, which could result in
445
greater physical damage to organisms.66
446 447
The inhibition of the growth of algal cells exposed to carbon nanotubes was strongly correlated with the shading of the nanomaterials,26 and similar indirect 15
ACS Paragon Plus Environment
Page 16 of 29
Page 17 of 29
Environmental Science & Technology
448
effects were observed for the cells exposed to graphene in this study. Different
449
graphene suspensions (10 mg/L) were placed in a constant-temperature incubator at
450
24°C for 5 days and were stirred once every 8 h. Black layers of graphene formed on
451
the surfaces of the pristine and hydrated graphene suspensions, as indicated by the
452
yellow arrows shown in Figure S11. The suspension of irradiated graphene was more
453
transparent than the other graphene suspensions, and no black layer was observed on
454
its surface. The transparency and presence or absence of a graphene layers on the
455
surfaces of each suspension are consistent with the hypothesis that the observed
456
variations in reproduction and chlorophyll concentrations could be attributed to the
457
shading effects of the graphene, as shown in Figure S9. Considering that transparency
458
was dependent on the level of graphene oxidation,67 the oxygen groups provided a
459
connection between the transparency of the graphene suspension and the inhibition of
460
algal reproduction. Notably, the various physicochemical properties (morphology,
461
defects, structure, surface chemistry, etc.) of nanomaterials were interdependent for
462
determining the nanotoxicity of such materials.8 It is important to consider the
463
systematic effects of these properties when analyzing the nanotoxicity of graphene.
464 465
ASSOCIATED CONTENT
466
Supporting Information Available
467
Additional descriptions of the graphene characterization, protein carbonyls and cellular
468
ultrastructure measurements, figures regarding the characteristics of graphene before
469
and after treatments, and the effects of hydration and irradiation on cellular number,
470
diameter and chlorophyll contents. This information is available free of charge via the
471
Internet at http://pubs.acs.org/.
472 473
AUTHOR INFORMATION
474
Corresponding author
475
* E-mail:
[email protected] (Q.Z). Phone: +86-022-23507800; fax:
476
+86-022-66229562.
477 16
ACS Paragon Plus Environment
Environmental Science & Technology
478
Notes
479
The authors declare no competing financial interest.
480 481
ACKNOWLEDGMENTS
482
This work was financially supported by the National Natural Science Foundation of
483
China as a general project (grant Nos. 31170473 and 21307061), a joint project (grant
484
No. U1133006) and a key project (grant No. 21037002), the Ministry of Education of
485
China as an innovative team project (grant no. IRT 13024), Tianjin Natural Science
486
Foundation (grant No. 14JCQNJC08900), the Specialized Research Fund for the
487
Doctoral Program of Higher Education of China (grant No. 2013003112016), the
488
Postdoctoral Science Foundation of China (grant No. 2014M550138) and the
489
Fundamental Research Funds for the Central Universities (grant No. 65121006).
490 491
REFERENCES
492
(1) Qu, X.; Brame, J.; Li, Q.; Alvarez, P. J. J. Nanotechnology for a safe and
493
sustainable water supply: Enabling integrated water treatment and reuse. Accounts
494
Chem. Res. 2013, 46, 834-843.
495
(2) Novoselov, K. S.; Falko, V. I.; Colombo, L.; Gellert, P. R.; Schwab, M. G.; Kim,
496
K. A roadmap for graphene. Nature 2012, 490, 192-200.
497
(3) Mao, H. Y.; Laurent, S.; Chen, W.; Akhavan, O.; Imani, M.; Ashkarran, A. A.;
498
Mahmoudi, M. Graphene: Promises, facts, opportunities, and challenges in
499
nanomedicine. Chem. Rev. 2013, 113, 3407-3424.
500
(4) Geim, A. K.; Novoselov, K. S. The rise of graphene. Nat. Mater. 2007, 6,
501
183-191.
502
(5) Gardea-Torresdey, J. L.; Rico, C. M.; White, J. C. Trophic transfer, transformation,
503
and impact of engineered nanomaterials in terrestrial environments. Environ. Sci.
504
Technol. 2014, 48, 2526-2540.
505
(6) Card, J. W.; Jonaitis, T. S.; Tafazoli, S.; Magnuson, B. A. An appraisal of the
506
published literature on the safety and toxicity of food-related nanomaterials. Crit. Rev.
507
Toxicol. 2011, 41, 20-49. 17
ACS Paragon Plus Environment
Page 18 of 29
Page 19 of 29
Environmental Science & Technology
508
(7) Sanchez, V. C.; Jachak, A.; Hurt, R. H.; Kane, A. B. Biological interactions of
509
graphene-family nanomaterials: An interdisciplinary review. Chem. Res. Toxicol.
510
2012, 25, 15-34.
511
(8) Hu, X. G.; Zhou, Q. X. Health and ecosystem risks of graphene. Chem. Rev. 2013,
512
113, 3815-3835.
513
(9) Xiu, Z. M.; Ma, J.; Alvarez, P. J. J. Differential effect of common ligands and
514
molecular oxygen on antimicrobial activity of silver nanoparticles versus silver ions.
515
Environ. Sci. Technol. 2011, 45, 9003-9008.
516
(10) Yang, Y.; Wang, J.; Xiu, Z. M.; Alvarez, P. J. J. Impacts of silver nanoparticles
517
on cellular and transcriptional activity of nitrogen-cycling bacteria. Environ. Toxicol.
518
Chem. 2013, 32, 1488-1494.
519
(11) Lowry, G. V.; Gregory, K. B.; Apte, S. C.; Lead, J. R. Transformations of
520
nanomaterials in the environment. Environ. Sci. Technol. 2012, 46, 6893-6899.
521
(12) Luu, X. C.; Yu, J.; Striolo, A. Nanoparticles adsorbed at the water/oil interface:
522
Coverage and composition effects on structure and diffusion. Langmuir 2013, 29,
523
7221-7228.
524
(13) Wang, D. A.; Liu, L. F.; Zhang, F. X.; Tao, K.; Pippel, E.; Domen, K.
525
Spontaneous phase and morphology transformations of anodized titania nanotubes
526
induced by water at room temperature. Nano Lett. 2011, 11, 3649-3655.
527
(14) Hou, W. C.; Jafvert, C. T. Photochemical transformation of aqueous C60 clusters
528
in sunlight. Environ. Sci. Technol. 2009, 43, 362-367.
529
(15) Chen, C. Y.; Jafvert, C. T. The role of surface functionalization in the solar
530
light-induced production of reactive oxygen species by single-walled carbon
531
nanotubes in water. Carbon 2011, 49, 5099-5106.
532
(16) Qu, X.; Alvarez, P. J. J.; Li, Q., Photochemical transformation of carboxylated
533
multiwalled carbon nanotubes: Role of reactive oxygen species. Environ. Sci. Technol.
534
2013, 47, 14080-14088.
535
(17) Allen, B. L.; Kichambare, P. D.; Gou, P.; Vlasova, II; Kapralov, A. A.; Konduru,
536
N.; Kagan, V. E.; Star, A. Biodegradation of singlewalled carbon nanotubes through
537
enzymatic catalysis. Nano Lett. 2008, 8, 3899-3903. 18
ACS Paragon Plus Environment
Environmental Science & Technology
538
(18) Sharifi, S.; Behzadi, S.; Laurent, S.; Forrest, M. L.; Stroeve, P.; Mahmoudi, M.
539
Toxicity of nanomaterials. Chem. Soc. Rev. 2012, 41, 2323-2343.
540
(19) Al Faraj, A.; Bessaad, A.; Cieslar, K.; Lacroix, G.; Canet-Soulas, E.; Cremillieux,
541
Y. Long-term follow-up of lung biodistribution and effect of instilled SWCNTs using
542
multiscale imaging techniques. Nanotechnology 2010, 21, 175103.
543
(20) Muller, J.; Huaux, F.; Fonseca, A.; Nagy, J. B.; Moreau, N.; Delos, M.;
544
Raymundo-Pinero, E.; Beguin, F.; Kirsch-Volders, M.; Fenoglio, I.; Fubini, B.; Lison,
545
D. Structural defects play a major role in the acute lung toxicity of multiwall carbon
546
nanotubes: toxicological aspects. Chem. Res. Toxicol. 2008, 21, 1698-1705.
547
(21) Long, Z.; Ji, J.; Yang, K.; Lin, D.; Wu, F. Systematic and quantitative
548
investigation of the mechanism of carbon nanotubes' toxicity toward algae. Environ.
549
Sci. Technol. 2012, 46, 8458-8466.
550
(22) Fubini, B.; Ghiazza, M.; Fenoglio, I. Physico-chemical features of engineered
551
nanoparticles relevant to their toxicity. Nanotoxicology 2010, 4, 347-363.
552
(23) Petersen, E. J.; Henry, T. B. Methodological considerations for testing the
553
ecotoxicity of carbon nanotubes and fullerenes: Review. Environ. Toxicol. Chem.
554
2012, 31, 60-72.
555
(24) Hou, W.-C.; BeigzadehMilani, S.; Jafvert, C. T.; Zepp, R. G. Photoreactivity of
556
unfunctionalized single-wall carbon nanotubes involving hydroxyl radical: Chiral
557
dependency and surface coating effect. Environ. Sci. Technol. 2014, 48, 3875-3882.
558
(25) Qu, X.; Alvarez, P. J. J.; Li, Q. Impact of Sunlight and humic acid on the
559
deposition kinetics of aqueous fullerene nanoparticles nC60. Environ. Sci. Technol.
560
2012, 46, 13455-13462.
561
(26) Schwab, F.; Bucheli, T. D.; Lukhele, L. P.; Magrez, A.; Nowack, B.; Sigg, L.;
562
Knauer, K. Are carbon nanotube effects on green algae caused by shading and
563
agglomeration? Environ. Sci. Technol. 2011, 45, 6136-6144.
564
(27) Wei, L. P.; Thakkar, M.; Chen, Y. H.; Ntim, S. A.; Mitra, S.; Zhang, X. Y.
565
Cytotoxicity effects of water dispersible oxidized multiwalled carbon nanotubes on
566
marine alga, Dunaliella tertiolecta. Aquat. Toxicol. 2010, 100, 194-201.
567
(28) Hu, X.; Zhou, Q. Novel hydrated graphene ribbon unexpectedly promotes aged 19
ACS Paragon Plus Environment
Page 20 of 29
Page 21 of 29
Environmental Science & Technology
568
seed germination and root differentiation. Sci. Rep. 2014, 4, 3782.
569
(29) Liu, Y.; Zhou, Q. X.; Xie, X. J.; Lin, D. S.; Dong, L. X. Oxidative stress and
570
DNA damage in the earthworm Eisenia fetida induced by toluene, ethylbenzene and
571
xylene. Ecotoxicology 2010, 19, 1551-1559.
572
(30) Fasolino, A.; Los, J. H.; Katsnelson, M. I. Intrinsic ripples in graphene. Nat.
573
Mater. 2007, 6, 858-861.
574
(31) Banhart, F.; Kotakoski, J.; Krasheninnikov, A. V. Structural defects in graphene.
575
ACS Nano 2011, 5, 26-41.
576
(32) Niu, T.; Zhou, M.; Zhang, J.; Feng, Y.; Chen, W. Growth intermediates for CVD
577
graphene on Cu (111): Carbon clusters and defective graphene. J. Am. Chem. Soc.
578
2013, 135, 8409-8414.
579
(33) Meyer, J. C.; Geim, A. K.; Katsnelson, M. I.; Novoselov, K. S.; Booth, T. J.;
580
Roth, S. The structure of suspended graphene sheets. Nature 2007, 446, 60-63.
581
(34) Yin, H. J.; Tang, H. J.; Wang, D.; Gao, Y.; Tang, Z. Y. Facile synthesis of
582
surfactant-free Au cluster/graphene hybrids for high-performance oxygen reduction
583
reaction. ACS Nano 2012, 6, 8288-8297.
584
(35) Graf, D.; Molitor, F.; Ensslin, K.; Stampfer, C.; Jungen, A.; Hierold, C.; Wirtz, L.
585
Spatially resolved raman spectroscopy of single- and few-layer graphene. Nano Lett.
586
2007, 7, 238-242.
587
(36) Huh, S. H. X-ray diffraction of multi-layer graphenes: Instant measurement and
588
determination of the number of layers. Carbon 2014, 78, 617-621.
589
(37) Talyzin, A. V.; Hausmaninger, T.; You, S. J.; Szabo, T. The structure of
590
graphene oxide membranes in liquid water, ethanol and water-ethanol mixtures.
591
Nanoscale 2014, 6, 272-281.
592
(38) Ali, M. A.; Reza, K. K.; Srivastava, S.; Agrawal, V. V.; John, R.; Malhotra, B. D.
593
Lipid lipid interactions in aminated reduced graphene oxide interface for biosensing
594
application. Langmuir 2014, 30, 4192-4201.
595
(39) Ramakrishnan, S.; Dhakshnamoorthy, M.; Jelmy, E. J.; Vasanthakumari, R.;
596
Kothurkar, N. K. Synthesis and characterization of graphene oxide-polyimide
597
nanofiber composites. Rsc. Adv. 2014, 4, 9743-9749. 20
ACS Paragon Plus Environment
Environmental Science & Technology
598
(40) Guo, H. L.; Wang, X. F.; Qian, Q. Y.; Wang, F. B.; Xia, X. H. A green approach
599
to the synthesis of graphene nanosheets. ACS Nano 2009, 3, 2653-2659.
600
(41) Tang, H. X.; Ehlert, G. J.; Lin, Y. R.; Sodano, H. A. Highly efficient synthesis of
601
graphene nanocomposites. Nano Lett. 2012, 12, 84-90.
602
(42) Chen, D. Z.; Li, L. D.; Guo, L. An environment-friendly preparation of reduced
603
graphene oxide nanosheets via amino acid. Nanotechnology 2011, 22, 325601.
604
(43) Luo, Z.; Lu, Y.; Somers, L. A.; Johnson, A. T. High yield preparation of
605
macroscopic graphene oxide membranes. J. Am. Chem. Soc. 2009, 131, 898-899.
606
(44) Zhu, C.; Guo, S.; Fang, Y.; Dong, S. Reducing sugar: New functional molecules
607
for the green synthesis of graphene nanosheets. ACS Nano 2010, 4, 2429-2437.
608
(45) Ali, M. A.; Kamil Reza, K.; Srivastava, S.; Agrawal, V. V.; John, R.; Malhotra,
609
B. D. Lipid–lipid interactions in aminated reduced graphene oxide interface for
610
biosensing application. Langmuir 2014, 30, 4192-4201.
611
(46) Bouchard, D.; Zhang, W.; Powell, T.; Rattanaudompol, U.-S. Aggregation
612
kinetics and transport of single-walled carbon nanotubes at low surfactant
613
concentrations. Environ. Sci. Technol. 2012, 46, 4458-4465.
614
(47) Lin, S.; Shih, C.-J.; Strano, M. S.; Blankschtein, D. Molecular insights into the
615
surface morphology, layering structure, and aggregation kinetics of
616
surfactant-stabilized graphene dispersions. J. Amer. Chem. Soc. 2011, 133,
617
12810-12823.
618
(48) Hou, W.-C.; Kong, L.; Wepasnick, K. A.; Zepp, R. G.; Fairbrother, D. H.; Jafvert,
619
C. T. Photochemistry of aqueous C60 clusters: Wavelength dependency and product
620
characterization. Environ. Sci. Technol. 2010, 44, 8121-812.
621
(49) Zhang, L.; Zhou, L.; Yang, M.; Liu, Z.; Xie, Q.; Peng, H. Photo-induced free
622
radical modification of graphene. Small 2013, 9, 1134-1143.
623
(50) Zepp, R. G.; Gumz, M. M.; Miller, W. L.; Gao, H. Photoreaction of
624
valerophenone in aqueous solution. J. Phys. Chem. A 1998, 102, 5716–5723.
625
(51) Yang, Y.; Mathieu, J. M.; Chattopadhyay, S.; Miller, J. T.; Wu, T.; Shibata, T.;
626
Guo, W.; Alvarez, P. J. Defense mechanisms of Pseudomonas aeruginosa PAO1
627
against quantum dots and their released heavy metals. ACS Nano 2012, 6, 6091-6098. 21
ACS Paragon Plus Environment
Page 22 of 29
Page 23 of 29
Environmental Science & Technology
628
(52) Xiu, Z.-M.; Ma, J.; Alvarez, P. J. Differential effect of common ligands and
629
molecular oxygen on antimicrobial activity of silver nanoparticles versus silver ions.
630
Environ. Sci. Technol. 2011, 45, 9003-9008.
631
(53) Yang, Y.; Wang, J.; Xiu, Z.; Alvarez, P. J. Impacts of silver nanoparticles on
632
cellular and transcriptional activity of nitrogen-cycling bacteria. Environ. Toxicol.
633
Chem. 2013, 32, 1488-1494.
634
(54) Yan, L.; Gu, Z.; Zhao, Y. Chemical mechanisms of the toxicological properties
635
of nanomaterials: generation of intracellular reactive oxygen species. Chem.-Asian J.
636
2013, 8, 2342-2353.
637
(55) Hu, X.; Mu, L.; Wen, J.; Zhou, Q. Covalently synthesized graphene
638
oxide-aptamer nanosheets for efficient visible-light photocatalysis of nucleic acids
639
and proteins of viruses. Carbon 2012, 50, 2772-2781.
640
(56) Tkalec, M.; Štefanić, P. P.; Cvjetko, P.; Šikić, S.; Pavlica, M.; Balen, B. The
641
effects of cadmium-zinc interactions on biochemical responses in tobacco seedlings
642
and adult plants. PloS one 2014, 9, 87582.
643
(57) Hu, X.; Kang, J.; Lu, K.; Zhou, R.; Mu, L.; Zhou, Q. Graphene oxide amplifies
644
the phytotoxicity of arsenic in wheat. Sci. Rep. 2014, 4, 6122.
645
(58) Berestovsky, G. N.; Ternovsky, V. I.; Kataev, A. A. Through pore diameter in
646
the cell wall of Chara corallina. J. Exp. Bot. 2001, 52, 1173-1177.
647
(59) Fenoglio, I.; Greco, G.; Tomatis, M.; Muller, J.; Raymundo-Pinero, E.; Béguin,
648
F.; Fonseca, A.; Nagy, J. B.; Lison, D.; Fubini, B. Structural defects play a major role
649
in the acute lung toxicity of multiwall carbon nanotubes: Physicochemical aspects.
650
Chem. Res. Toxicol. 2008, 21, 1690-1697.
651
(60) Akhavan, O.; Ghaderi, E. Toxicity of graphene and graphene oxide nanowalls
652
against bacteria. ACS Nano 2010, 4, 5731-5736.
653
(61) Liu, S.; Zeng, T. H.; Hofmann, M.; Burcombe, E.; Wei, J.; Jiang, R.; Kong, J.;
654
Chen, Y. Antibacterial activity of graphite, graphite oxide, graphene oxide, and
655
reduced graphene oxide: membrane and oxidative stress. ACS Nano 2011, 5,
656
6971-6980.
657
(62) Wang, Q.; Lee, S.; Choi, H. Aging study on the structure of Fe0-nanoparticles: 22
ACS Paragon Plus Environment
Environmental Science & Technology
658
stabilization, characterization, and reactivity. J. Phys. Chem. C 2010, 114, 2027-2033.
659
(63) Poland, C. A.; Duffin, R., Kinloch, I.; Maynard, A.; Wallace, W. A. H.; Seaton,
660
A.; Stone, V.; Brown, S.; MacNee, W.; Donaldson, K. Carbon nanotubes introduced
661
into the abdominal cavity of mice show asbestos-like pathogenicity in a pilot study.
662
Nat. Nanotechnol. 2008, 3, 423–428.
663
(64) Xie, Z.; Liu, Y.; Hua, C.; Chen, L.; Li, D. Relationships between the biomass of
664
algal crusts in fields and their compressive strength. Soil Biol. Biochem. 2007, 39,
665
567-572.
666
(65) Belyanskaya, L.; Weigel, S.; Hirsch, C.; Tobler, U.; Krug, H.; Wick, P. Effects of
667
carbon nanotubes on primary neurons and glial cells. Neurotoxicology 2009, 30,
668
702-711.
669
(66) Wick, P.; Manser, P.; Limbach, L.; Dettlaff-Weglikowska, U.; Krumeich, F.;
670
Roth, S.; Stark, W.; Bruinink, A. The degree and kind of agglomeration affect carbon
671
nanotube cytotoxicity. Toxicol. Lett. 2007, 168, 121-131.
672
(67) Krishnamoorthy, K.; Veerapandian, M.; Yun, K.; Kim, S.-J. The chemical and
673
structural analysis of graphene oxide with different degrees of oxidation. Carbon
674
2013, 53, 38-49.
675 676 677 678 679 680 681 682 683 684 685 686 687 23
ACS Paragon Plus Environment
Page 24 of 29
Page 25 of 29
Environmental Science & Technology
688
Figure Legends
689
Figure 1. Atomic force microscope images of graphene before and after treatment (a)
690
and the energy-dispersive spectra of the raised dots on the treated graphene (b).
691 692
Figure 2. Field-emission transmission electron microscopy images of pristine
693
graphene, hydrated graphene and visible-light-irradiated graphene. The scale bars of
694
the left- and right-hand images represent 0.2 µm and 10 nm, respectively.
695 696
Figure 3. Raman spectra (a) and X-ray powder diffraction patterns (b) of pristine
697
graphene, hydrated graphene and irradiated graphene. The labels D, G, 2D and D + G
698
denote the typical Raman spectral peaks of graphene. The main peaks in the X-ray
699
diffraction pattern can be subdivided into multiple individual subpeaks (d1, d2, d3 and
700
d4).
701 702
Figure 4. Effects of hydration and visible-light irradiation on the reduction of
703
oxidative stress and DNA damage induced by pristine graphene.
704 705 706 707 708 709 710 711 712 713 714 715 716 717 24
ACS Paragon Plus Environment
Environmental Science & Technology
Figures
718
719 720 721
Figure 1. ---------------------
722 723 724 725 726 727 728 729 730 731 732 25
ACS Paragon Plus Environment
Page 26 of 29
Page 27 of 29
Environmental Science & Technology
733 734 735
Figure 2. ------------------
736 737 738 739 740 741 742 743 744 745 26
ACS Paragon Plus Environment
Environmental Science & Technology
746 747 748
Figure 3. -----------------
749 750 751 752 753 754 755
27
ACS Paragon Plus Environment
Page 28 of 29
Page 29 of 29
Environmental Science & Technology
756 757 758
Figure 4. -----------------
759
28
ACS Paragon Plus Environment