Graphene in the Aquatic Environment: Adsorption ... - ACS Publications

Aug 14, 2014 - Environmental Science & Technology 2016 50 (21), 11627-11636 .... Effects of Solution Chemistry on Adsorption of Selected Pharmaceutica...
0 downloads 4 Views 1MB Size
Subscriber access provided by GAZI UNIV

Critical Review

Graphene in the Aquatic Environment: Adsorption, Dispersion, Toxicity and Transformation Jian Zhao, Zhenyu Wang, Jason C. White, and Baoshan Xing Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es5022679 • Publication Date (Web): 14 Aug 2014 Downloaded from http://pubs.acs.org on August 21, 2014

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 61

Environmental Science & Technology

1

Graphene in the Aquatic Environment: Adsorption, Dispersion,

2

Toxicity and Transformation

3

Jian Zhao,†, ‡ Zhenyu Wang,†,‡ Jason C. White,§ and Baoshan Xing*,‡

4 5



6

Education, Ocean University of China, Qingdao 266100, China

7



8

USA

9

§

10

College of Environmental Science and Engineering, Key Laboratory of Ministry of

Stockbridge School of Agriculture, University of Massachusetts, Amherst, MA 01003,

Department of Analytical Chemistry, The Connecticut Agricultural Experiment Station,

New Haven, CT 06504, USA

11 12 13 14 15

*Corresponding author

16

Tel.: +1 413 545 5212

17

E-mail address: [email protected] (Prof. Baoshan Xing)

18 19

1

ACS Paragon Plus Environment

Environmental Science & Technology

20

Abstract:

21

Graphene-family nanomaterials (GFNs) including pristine graphene, reduced graphene

22

oxide (rGO) and graphene oxide (GO) offer great application potential, leading to the

23

possibility of their release into aquatic environments. Upon exposure, graphene/rGO and

24

GO exhibit different adsorption properties towards environmental adsorbates, and the

25

molecular interactions at the GFN-water interface are discussed. After solute adsorption,

26

the dispersion/aggregation behaviors of GFNs can be altered by solution chemistry, as

27

well as by the presence of colloidal particles and biocolloids. GO has different dispersion

28

performance from pristine graphene and rGO, which is further demonstrated from surface

29

properties. Upon exposure in aquatic environments, GFNs have adverse impacts on

30

aquatic organisms (e.g., bacteria, algae, plants, invertebrates and fish). The mechanisms

31

of GFNs toxicity at the cellular level are reviewed and the remaining unclear points on

32

toxic mechanisms such as membrane damage are presented. Moreover, we highlight the

33

transformation routes of GO to rGO. The degradation of GFNs upon exposure to UV

34

irradiation and/or biota is also reviewed. In view of the unanswered questions, future

35

research should include comprehensive characterization of GFNs, new approaches for

36

explaining GFNs aggregation, environmental behaviors of metastable GO, and the

37

relationship between dispersion of GFNs and the related adsorption properties.

38

2

ACS Paragon Plus Environment

Page 2 of 61

Page 3 of 61

39

Environmental Science & Technology

TOC art:

40 41

3

ACS Paragon Plus Environment

Environmental Science & Technology

42

1. Introduction

43

Graphene, due to its exceptional mechanical, electronic, optical and catalytic

44

properties,1,2,3 has attracted great attention in the scientific community. Ideal graphene

45

(Figure 1A) is a single layer of sp2-hybridized carbon atoms joined by covalent bonds.4,5

46

Because of the difficulty in isolating single layers of graphene, “few-layer” graphene

47

(2~5 layers), multilayer graphene (2~10 layers) and graphite nanoplates (2D graphite

48

material with a thickness and/or lateral dimension less than 100 nm) are all considered as

49

graphene-family nanomaterials (GFNs).6 Another type of GFNs is graphene derivatives,

50

such as graphene oxide (GO) (Figure 1B). GO is an intermediate product during synthesis

51

of reduced graphene oxide (rGO, Figure 1C) and prepared by oxidative exfoliation of

52

graphite.7 The synthesis and chemistry of rGO and GO were comprehensively reviewed

53

by Loh et al.8 and Dreyer et al.,9 respectively.

54

Currently, GFNs are in commercial production.10 With projected rapid increases in

55

production and application,11 GFNs will likely be released into the environment at

56

significant levels. For example, GO-polymer nanocomposites could release GO particles

57

during their life cycle because of photodegradation upon UV exposure.12 Notably,

58

considerable GFN release could occur during environmental applications such as

59

adsorbents for wastewater and drinking water treatment,13,14,15 materials for solid-phase

60

extraction,16 membranes for desalination,17 catalysts for aqueous organic pollutant

61

oxidation and degradation,18 and coating materials for filtration.19 In addition, GFNs

62

could be introduced to the environment during the waste disposal of GFN-containing

63

products. The released GFNs may have significant adverse environmental impacts.

64

Consequently, significant research attention has recently been focused on the

4

ACS Paragon Plus Environment

Page 4 of 61

Page 5 of 61

65

Environmental Science & Technology

environmental behavior of GFNs.12,20,21

66

However, the existing literature on environmental behavior of GFNs remains limited.

67

In aquatic environments, GFNs would first interact with system components such as

68

inorganic ions and natural organic matter (NOM), both of which are able to modify the

69

surface of GFNs. After adsorption, dispersion-aggregation behaviors could also be altered.

70

In addition, toxicity to aquatic organisms may be an important concern after

71

environmental GFN release. Upon interaction with environmental media and biota, GFNs

72

have been shown to transform and degrade.21 This thorough review will focus mainly on

73

these four aspects of GFN fate: adsorption, dispersion, toxicity and transformation. Also,

74

this review will identify key knowledge gaps and provide valuable insight into the

75

transport, fate, and risk posed from GFNs, as well as their potential use for select

76

environmental applications.

77

2. Adsorption properties of GFNs in aquatic environments

78

Adsorption on GFNs is a critical physiochemical process at the GFN-water interface

79

due to large surface area and surface-active properties of the sorbents. This process can (1)

80

affect the mobility and fate of both organic matter (e.g., humic acid, amino acids, proteins)

81

and xenobiotic pollutants (e.g., heavy metals, hydrophobic organic compounds); (2) alter

82

the surface characteristics of GFNs through coating and modification; and (3) be of

83

importance for potential adsorbent applications in water treatment technologies.

84

Adsorption of organic and inorganic compounds on GFNs has been investigated. In

85

reviewing the current literature (Table S1), molecular interactions and adsorption

86

capacity on the GFN surface will be addressed with regard to the structural properties of

87

both GFNs and adsorbates, as well as important environmental factors.

5

ACS Paragon Plus Environment

Environmental Science & Technology

88

2.1. Interactions at the GFN-water interface. For environmental adsorbates such as

89

organic molecules, inorganic ions, macromolecules and mineral particles (Table S1), five

90

molecular interactions are summarized: π bonding interaction, hydrophobic effect;

91

electrostatic interaction (Coulomb force), hydrogen bonding and Lewis acid-base

92

interaction. The four types of π bonding interaction are π–π, n–π, cation–π and anion–π

93

bonding. π–π bonding occurs between C=C double bonds or benzene rings of adsorbed

94

organic molecules and benzene rings on GFN surface via π–π coupling. π–π interaction is

95

suggested as one of most important mechanisms contributing to GFN interaction with

96

aromatic compounds (e.g., PAHs, biphenyl, phenol),22,23 macromolecules (e.g., humic

97

acid, DNA, protein),24,25,26 and graphitic particles (e.g., carbon nanotubes (CNTs)).27 n–π

98

interactions occur between GFN electron-depleted sites and n-electron donors of organic

99

compounds which contain oxygen (e.g., 1-naphthol) or nitrogen (e.g., 1-naphthylamine)

100

with lone electron pairs.28 Cation–π bonds form between the GFN π–electrons and easily

101

protonated amino groups (e.g., tetracycline)29 or metal ions (e.g., Pb (II)).30 Conversely,

102

the GFN surface could act as an electron acceptor for anion (e.g., F–, Cl–, Br–) adsorption;

103

the anion–π bonding was unexpectedly strong using a computational method based on the

104

density functional theory.31 Hydrophobic effects will occur with hydrophobic organic

105

compounds (HOCs) on graphene and rGO, both of which have water hydrophobic

106

external surfaces. The strength of binding is directly related to the hydrophobicity of the

107

organic chemical. The hexadecane–water distribution coefficient (KHW) is a preferred

108

parameter over the octanol-water distribution coefficient (KOW) to correlate adsorption on

109

GFNs because octanol contains an –OH group that may possibly interact with organic

110

chemicals.32 For GO, the hydrophobic interaction is generally weak due to oxygen-

6

ACS Paragon Plus Environment

Page 6 of 61

Page 7 of 61

Environmental Science & Technology

111

containing functional groups (e.g., epoxy and hydroxyl groups) on the edge but still

112

occurs because of the hydrophobic basal plane.26 Electrostatic interaction depends on the

113

charge nature of both GFNs and adsorbates. GO sheets are strongly negatively charged,

114

with a zero point of charge (ZPC) lower than pH 2.0.33,34 Thus, the GO surface charge is

115

negative in aquatic environments (pH 5-9), showing electrostatic attraction towards both

116

organic and inorganic cations.35,36,37 A weaker attraction would exist on rGO surfaces due

117

to the residual oxygen functional groups after the reduction process.38 Hydrogen bonding

118

has been used to explain the interaction of GO functional groups with the adsorbed

119

organic compounds,39 humic acid24 and mineral particles24 that contain oxygen-functional

120

groups. In aquatic environments, this interaction could be weakened by water molecules

121

that form hydrogen bonds with GO. Lewis acid-base interactions, including ligand

122

exchange, metal ion exchange and complexation, play an important role in contributing to

123

sorption on GFNs, especially for GO.40 Metal ion exchange and complexation are

124

primary mechanisms for the adsorption of metal ions (e.g., Cu (II), Pb (II)) on GO.37,41,42

125

The specific interactions for each type of adsorbate are shown in Table S2. The potential

126

interactions that have not been reported in the literature but are possible based on

127

theoretical analysis are also listed in Table S2. For example, although there are no current

128

reports on the formation of covalent bonds, this interaction may occur between GO and

129

organic compounds containing functional groups such as –OH, –COOH and –NH2 as the

130

formation of covalent bonds have been reported on organic compounds (e.g., L-

131

phenylalanine) and CNTs.43

132

The strength of GFN interactions with adsorbates has been studied by both density

133

functional theory (DFT) calculations and by batch experiments. DFT calculation has been

7

ACS Paragon Plus Environment

Environmental Science & Technology

134

applied to simulate the binding energy and adsorption configuration of both inorganic

135

(Li+, Cl-)31,44,45 and organic adsorbates (benzene and naphthalene)46 on the graphene

136

surface. Batch experiments with the aid of isotherm modeling could also be used to

137

evaluate interaction strength. Among the models (Table S3), the Polanyi theory based

138

Dubinin-Ashtakhov (DA) model has parameters E and b, which describe the strength of

139

interaction forces.47 For most adsorbates, multiple interactions operate simultaneously at

140

the GFN-water interface (Table S2). The individual interactions can be identified by the

141

properties of both GFN and adsorbates. For example, π−π bond can be estimated by

142

π−electron polarizability (π*). A detailed explanation of specific interactions has been

143

summarized previously by our group.32,47 Currently, the contribution of individual

144

interactions is hypothesized in most studies. How one can accurately calculate/quantify

145

the contribution of a specific interaction is challenging. It is worth mentioning that Xia et

146

al. reported a biological surface adsorption index (BSAI) approach based on the

147

experimental obtained adsorption coefficients (K) and adsorbate properties such as excess

148

molar refraction, effective dipolarity and polarizability, hydrogen-bond acidity and

149

basicity, and McGowan characteristic volume.48 The BSAI model can predict the K value

150

of a specific organic adsorbate on CNTs and quantify the contribution of a specific

151

interaction without experimental adsorption data. For GFNs, a similar model may be

152

established after more adsorption data (at least twenty compounds) become available in

153

the near future.

154

2.2. Adsorption capacity. For single-layer graphene, both sides of the graphitic

155

surface provide adsorption sites. However, few-layer and multilayer graphene sheets

156

comprised of stacked graphene layers have a much greater likelihood of entering aquatic

8

ACS Paragon Plus Environment

Page 8 of 61

Page 9 of 61

Environmental Science & Technology

157

environments due to the complicated and difficult process of isolating single-layer

158

graphene.4 Single layers in multilayer graphene are stacked as in graphite,49 and the

159

interlayer space will not be occupied by environmental adsorbates such as organic

160

molecules because of the low adsorption capacity of graphite.40,50 Therefore, only the

161

exterior surfaces of multilayer graphene are available for adsorption. GFNs adsorption

162

capacity has mainly been obtained from isotherm modeling. Among the commonly used

163

isotherm models (Table S3), Langmuir and Polanyi theory-based DA models have a

164

maximum adsorption capacity parameter (Q0). The Langmuir model has been widely

165

used for inorganic and organic molecule adsorption fitting, whereas the DA model is

166

currently only applicable for an organic compound whose water solubility is available at

167

a specific pH and temperature. Figure 2 presents the relationship between the Brunauer-

168

Emmer-Teller (BET) surface area of GFNs and Langmuir-fitted adsorption capacity

169

(mmol/g) for adsorbates. There is a positive correlation for the adsorption of metal ions

170

and organic molecules on GO and rGO, respectively, showing that surface area is the

171

primary factor determining the magnitude of adsorption onto GFNs. The low R2 values of

172

the correlation can be explained as follows: (1) BET surface area cannot completely

173

reflect the actual surface area because of GFNs agglomeration and aggregation in

174

aqueous solution;51 (2) other parameters such as micropore volume could also contribute

175

the high-energy adsorption sites,52 but limited work has characterized this phenomena; (3)

176

oxygen content, number of functional groups, and structural defects which highly depend

177

on the oxidation/reduction procedures could also affect the adsorption capacities of both

178

GO and rGO;9,53 and (4) the occupied adsorption space by individual molecules depends

179

on molecular size and interface conformation that differ among the reported compounds.

9

ACS Paragon Plus Environment

Environmental Science & Technology

180

These differences could be overcome by using molecular dynamics (MD) simulation that

181

has been employed on peptide (arginine, glutamine, and asparagine) adsorption,54

182

surfactants (sodium cholate),55 fatty acids and cellulose.56 The last two factors most

183

certainly are responsible for the poor correlation between surface area of GO and the

184

adsorption capacity for organic molecules (Figure 2, Table S4). There are very few data

185

on adsorption of metal ions by rGO, probably because the functional groups on rGO are

186

not abundant enough for ion complexation and as such, has not been considered as an

187

attractive material for metal ion adsorption. In addition, adsorption capacity is one of the

188

main factors driving the increasing application of GFNs in various technologies. The

189

approaches on the enhancement of adsorption capacity are further summarized in the

190

Supporting Information.

191

2.3. Implications of adsorption on GFNs. In aquatic environments, two or more

192

different types of adsorbates are often co-existing. These adsorbates are likely to be co-

193

adsorbed on GFN sheets. Inorganic ions are able to produce a screening or shading effect

194

on the GFNs surface charge, thereby promoting the adsorption of negatively charged

195

organic molecules (e.g., Bisphenol A at pH 6.0)38 and the suppression of positively

196

charged organic molecules (e.g., tetracycline at pH 3.6).29 In addition, the increase in

197

ionic strength (I.S.) could also cause an increased “salting-out” effect by decreasing the

198

solubility of organic molecules, thus altering their adsorption onto GFNs. In comparison

199

with ionized organic molecules, the influence of I.S. on the non-ionized molecules is

200

minimal.57 Metal ions will preferentially adsorb onto GO rather than rGO because of the

201

abundant surface functional groups.13,36,58 Unexpectedly, a screening effect and

202

competitive adsorption from I.S. have minor influences on metal ion adsorption.13,59

10

ACS Paragon Plus Environment

Page 10 of 61

Page 11 of 61

203

Environmental Science & Technology

Instead, solution pH was shown to be a dominant factor in metal adsorption onto GO.13,42

204

NOM is widely present in natural aquatic environments, and its effect on adsorption is

205

more complex. NOM molecules have been reported to adsorb on GO via hydrogen bonds,

206

Lewis acid-base and π-π interactions;24 the maximum adsorption of humic acid by

207

graphite oxide could be as high as 190 mg/g.60 Although there are no reports on the

208

interaction of NOM with rGO, strong adsorption is expected via hydrophobic and π-π

209

interactions. During co-adsorption with other adsorbates such as organic molecules and

210

metal ions, NOM could (1) bind/solubilize organic molecules and metal ions; (2) compete

211

with organic molecules and metal ions; and (3) indirectly adsorb organic molecules and

212

metal ions on GFN surface. The integrated effect of NOM on organic molecule

213

adsorption was reported as suppression by Apul et al.22 This suppression was less obvious

214

for biphenyl, with a flexible molecular structure, than for phenanthrene, with a planar and

215

rigid molecular structure.22 Furthermore, adsorption suppression by NOM was lower on

216

GFNs (both graphene and GO) than CNTs and activated carbon, both of which had

217

higher micropore volumes. For metal ions, Zhao et al. reported a reduction of adsorption

218

at pH < 8.0 and an insignificant effect at pH > 8.0 for Co(II) on GO in the presence of

219

humic acid, while adsorption reduction was observed for Cd(II) across the whole pH

220

range (2-10).13 The reduction of adsorption was attributed to stronger complexation

221

activity of GO in comparison with bound humic acid on the GO surface. The effect of

222

NOM on metal ion adsorption by rGO has not been evaluated. Considering the low

223

complexation activity of rGO in the absence of humic acid, adsorption enhancement is

224

expected. This hypothesis could be supported by the data of Lin et al.,61 where Pb(II)

225

sorption on CNTs was greatly enhanced by the oxygen-containing functional groups and

11

ACS Paragon Plus Environment

Environmental Science & Technology

226

negative charges on CNT surfaces coated with humic acid.

227

Desorption hysteresis on GFNs is an important phenomenon impacting the mobility

228

and fate of adsorbates in aquatic environments. Desorption hysteresis may be artificial,

229

and could be caused by experimental artifacts such as change of adsorption temperature,

230

unachieved equilibrium, inadequate solid-liquid separation or degradation of the

231

adsorbate.47 True hysteresis for carbon nanomaterials such as CNTs and C60 is explained

232

by deformation-rearrangement of aggregates62 and impurities (e.g., disorderly amorphous

233

carbon).63 The formation of covalent bonds between the functional groups on carbon

234

nanomaterials and organic molecules is another possible reason for desorption hysteresis,

235

although this has not been directly shown. For GFNs, the hysteresis phenomenon of

236

organic molecules has been observed in nearly all current desorption studies.50,64,65 Zhang

237

et al. attributed this hysteresis on GO to entrapment of tetrabromobisphenol A molecules

238

within the highly condensed matrices.64 In our recent study, we observed two states of

239

graphene sheets in aqueous solution which were responsible for the hysteresis of

240

phenanthrene.65 One state was the curling and folding of graphene sheets and the other

241

was surrounding the edge of a sheet that was enclosed by other sheets. Both states could

242

form closed interstitial spaces and entrap phenanthrene molecules after rearrangement of

243

graphene sheets in water. In addition, the sheet-sheet aggregation of GFNs could be

244

induced by adsorbate molecules, and these molecules inside this sandwich-like structure

245

may not be able to diffuse out through the gallery space. Although no report on this

246

hypothesis has been published, it may be another reason for the observed hysteresis. Due

247

to desorption hysteresis, pollutants in aquatic environments would be locked or

248

sequestered within GFNs aggregates, thus decreasing contaminant bioavailability and

12

ACS Paragon Plus Environment

Page 12 of 61

Page 13 of 61

Environmental Science & Technology

249

environmental risk, assuming there is not particle ingestion by biota. However, the risk

250

could be enhanced if the graphene-pollutant complexes are transported to different

251

aquatic environments such as estuaries, or are accumulated by animals or human beings.

252

GFN sheets are able to adsorb both organic molecules (e.g., amino acids, proteins) and

253

inorganic ions (K+, NO3-, Co2+, and Cu2+) (Table S1). An important study by Creighton et

254

al. showed a significant depletion of micronutrients such as folic acid, pyridoxine HCl

255

and niacinamide upon graphene exposure,66 indicating an indirect “starvation” toxicity

256

mechanism in human hepatoma cells. Similar nutrient depletion is expected to occur in

257

toxicological studies with aquatic organisms, although no findings have been reported. In

258

addition, GFNs with high adsorption capacity might cause artifacts during in-vitro

259

toxicology assays, including preventing molecular probes from interacting with their

260

targets, quenching fluorescent probes in fluorescent tests, and/or reducing/oxidizing

261

probes.66,67,68 Great care needs to be taken to avoid these adsorption-related artifacts and

262

false positive results during risk assessment assays.69

263

3. Dispersion/aggregation behaviors of GFNs in aquatic environments

264

Graphene and rGO sheets with a hydrophobic graphitic lattice tend to undergo layer-

265

to-layer aggregation in water due to hydrophobic forces, whereas GO sheets with

266

carboxyl, hydroxyl, and epoxy groups on the surface are able to form relatively stable

267

suspensions.33 Upon release into aquatic environments, GFNs are likely to interact with

268

inorganic ions, surface active molecules, NOM, colloidal particles, as well as biocolloids.

269

Their dispersion/aggregation behaviors can be very complex. Moreover, the transport and

270

fate of GFNs is governed by their suspension stability and this may well be altered upon

271

interactions with natural system constituents.

13

ACS Paragon Plus Environment

Environmental Science & Technology

272

3.1. Dispersion/aggregation of GO. The stability of GO sheets in water is affected by

273

the solution chemistry. The possible dispersion/aggregation processes are illustrated in

274

Figure 3A. Electrostatic repulsion is a major driving force preventing the GO sheets from

275

aggregating.70 Therefore, suspension stability of GO is pH-tunable owing to different

276

ionization degrees of the functional groups on the surface. Negatively charged colloids

277

with zeta potentials more than –30 mV are considered to be electrostatically stable,34

278

while ASTM defines colloids more negative than –40 mV as having “good stability”.71

279

Hence, GO sheets are generally stable in neutral solutions where the zeta potentials of

280

GOs were reported to be –40 mV,34 –45 mV,72 and –50 mV.73 As shown in Process I of

281

Figure 3A, at higher pHs, greater amounts of functional groups on the surface are ionized,

282

and the dispersion becomes more electrostatically stable. At lower pHs, the functional

283

groups are protonated, and individual GO sheets form layer-to-layer aggregates as

284

observed from MD simulations.74 Recently, the colloidal stability of GO under

285

environmentally relevant pH range was studied by Lanphere et al. (pH 5-9)75 and

286

Chowdhury et al. (pH 4-10);33 findings suggest a minor influence of pH on the dispersion

287

of GO in most aquatic environments.

288

Due to ionized functional groups on GO surfaces, an electrical double layer (EDL) is

289

formed in aquatic solutions. The presence of electrolytes is able to compress the EDL and

290

screen the surface charge of GO, thus resulting in irreversible homoaggregation76

291

(Process II in Figure 3A). This is supported by the results of Yoon et al.,73 in which the

292

net charge on GO nanoplatelets was significantly decreased with increasing salinity

293

(NaCl) at a specific pH. Lanphere et al. also reported a great decrease of net

294

electrophoretic mobility and an increase of hydrodynamic diameter when the I.S. of KCl

14

ACS Paragon Plus Environment

Page 14 of 61

Page 15 of 61

Environmental Science & Technology



1.5

M.75 Moreover, divalent or polyvalent cations may bridge the

295

was higher than 10

296

functional groups on two individual GO sheets via complexation77 as illustrated in

297

Process II (Figure 3A), thus inducing stronger aggregation. In addition, GO aggregation

298

in electrolytes likely follows Derjaguin-Landau-Verwey-Overbeek (DLVO) theory. The

299

critical coagulation concentrations (CCC) of GO were reported to be 44 and 0.9 mM for

300

NaCl and CaCl2, respectively.33 However, in another report, GO nanoplatelets were stable

301

and homogeneous over salinity (NaCl) concentrations up to 5 wt% (i.e., I.S. at 0.85 M),73

302

indicating that this type of GFNs may still have a good stability in seawater (~3.5 wt%).

303

The different GO dispersion performances could be explained by the material size and

304

oxidation degree. GO sheets with smaller size and lower C/O ratio will exhibit better

305

dispersion due to lower van der Waals attraction and higher electrostatic repulsion.72,73

306

The interaction of GO sheets with macromolecules such as humic acids and proteins is

307

shown as Process III (a,b) in Figure 3A. After adsorption on GO via π-π stacking,

308

hydrogen-bonding, and Lewis acid-base interactions as described earlier, humic acid is

309

able to improve the stability of GO sheets.33 Electrostatic and steric repulsions are

310

responsible for the stability of colloidal particles.76 Because the coating of humic acid did

311

not significantly influence GO surface charge, steric repulsion would be the major

312

determinant in dispersing GO. It should be noted that this dispersion enhancement by the

313

macromolecules could also be influenced by solution pH. Another possible process

314

during GO-NOM interaction is bridging. Bridging has been observed between negatively

315

charged TiO2 nanoparticles (NPs) and fulvic acid.78 Our previous study also found that

316

hydrodynamic diameter of CuO NPs was increased in the presence of fulvic acid.79 For

317

GO, bridging caused flocculation with humic acid, subsequently settling out from

15

ACS Paragon Plus Environment

Environmental Science & Technology

318

water.24 The destabilization of GO due to bridging may also occur in the presence of

319

protein, which was supported by the observed GO aggregation by fetal bovine serum in

320

the culture medium.76 In addition, Zinchenko et al. reported that NPs (e.g., C60(OH)n and

321

CNTs) could be removed from water by entrapment within long-chain DNA-chitosan

322

complexes, indicating that GO sheets may be also assembled/wrapped by long-chain

323

natural macromolecules such as alginic acid and DNA.80

324

Aquatic environments contain a great number of natural particles,81 and as such,

325

heteroaggregation is an important process for GO sheets (Process (IV) in Figure 3A).

326

Heteroaggregation of GO with two mineral clays, hematite and kaolin, was investigated

327

recently.24 For positively charged hematite, an electrostatical patching effect governed

328

aggregation with GO, and resulted in a flat configuration. For negatively charged kaolin,

329

the heteroaggregation with GO was through Lewis acid-base and hydrogen-bonding

330

interactions after charge repulsions were overcome. There have been no reports on the

331

interaction of GO with other types of minerals such as montmorillonite and goethite.

332

Interestingly, GO has been considered as an amphiphilic material with hydrophilic

333

domains on the edges and hydrophobic domains on the basal plane. The GO sheets can

334

disperse insoluble materials such as graphite, multi-wall CNT (MWCNT) and single-wall

335

CNT (SWCNT) to form stable suspensions through strong supramolecular π-π

336

stacking;27,72,82 the obtained suspensions were stable over two years.82 In aquatic

337

environments, GO sheets are likely to interact with other solid particles which have π-

338

conjugated structure (e.g., black carbon). However, GO is unlikely to disperse these

339

materials in aquatic environments unless GO concentration is extremely high (25-1000

340

mg/L).72,82

16

ACS Paragon Plus Environment

Page 16 of 61

Page 17 of 61

Environmental Science & Technology

341

Importantly, dispersion/aggregation behaviors of GO sheets in aquatic environments

342

are governed by the synergistic effects of individual environmental factors as shown in

343

Figure 3. There is only one report focusing on natural water and natural systems

344

directly.33 In this study, a short-term (1 hour) investigation showed that aggregation of

345

GO was negligible in almost all types of water (surface water, ground water and

346

wastewater included) except for a wastewater from North Oconee and a synthetic

347

groundwater. After long-term (28 days) testing, GO sheets remained stable only in a

348

natural surface water sample from Calls Creek and several synthetic wastewaters. The

349

aggregation of GO suspension in these waters was mainly induced by I.S. and ionic

350

species (e.g., Ca2+, Mg2+) while NOM (e.g., HA) partially suppressed this destabilizing

351

effect.

352

3.2. Dispersion/aggregation of graphene and rGO. Different from GO, graphene and

353

rGO sheets are unlikely to be stable in water due to their hydrophobicity. Li and

354

coworkers developed an approach to produce aqueous rGO suspension through

355

electrostatic repulsion by increasing solution pH (pH 10) and minimizing electrolyte

356

concentration.34 However, with the addition of electrolytes (NaCl), the suspension

357

became destabilized and immediately coagulated (Process V in Figure 3B). Therefore,

358

rGO sheets in natural waters would most likely settle out as large agglomerates. Another

359

approach to enhance the stabilization is noncovalent modification through dispersant

360

amendment. Environmental dispersants such as surfactants (Process VI (c,d)) and

361

macromolecules (Process VII) could prevent rGO aggregation by disrupting the inter-

362

sheet van der Waals attractions.83

363

There are numerous studies that investigated the dispersion of graphene during

17

ACS Paragon Plus Environment

Environmental Science & Technology

364

surfactant exfoliation of graphite.84-88 Surfactants adsorbed on graphene sheets can be in

365

the form of random encapsulation, micelles or hemimicelles.89 Guardia and coworkers

366

found that nonionic surfactants (especially for Tween 80 and P-123) were more effective

367

than ionic counterparts at stabilizing graphene sheets.85 They further concluded that steric

368

repulsion (contributed by nonionic surfactants) is more efficient than electrostatic

369

repulsion (contributed by ionic surfactants). However, Lin et al. explored the aggregation

370

kinetics of graphene stabilized by the ionic surfactant sodium cholate through MD

371

simulation and theoretical modeling.55 Their results revealed that steric hindrance was

372

more important than electrostatic interactions at graphene dispersion. These conflicting

373

reports on ionic surfactants suggest that the stabilization mechanism is complex and may

374

also relate to other parameters, including maximum adsorption capacity, surfactant chain

375

length and critical micelle concentration (CMC). Other surface-active molecules such as

376

pyrene derivatives,90 amino acids91 and polyphenol92 are also able to disperse graphene

377

sheets. For all dispersants, a well-dispersed graphene suspensions still requires sonication

378

to provide the activation energy to weaken layer-to-layer van der Waals interactions

379

between the layered sheets. However, long-term sonication appears to reduce flake size

380

and produce defects and oxygen-containing functional groups.93

381

Compared to surfactants, very few dispersion studies focusing on macromolecules are

382

available. Recently, DNA molecules (oligodeoxynucleotides) were used to disperse

383

graphene sheets through MD simulations.83 DNA prevented graphene aggregation

384

through disruption of interlayer van der Waals interactions, with the dispersion being

385

dependent on DNA size and molarity. It is reported that CNT aggregation could be

386

reduced by macromolecules such as humic acid, bovine serum albumin (BSA) and

18

ACS Paragon Plus Environment

Page 18 of 61

Page 19 of 61

Environmental Science & Technology

387

alginate.94 Similarly, tannic acid also exhibited good dispersion performance towards

388

CNTs through steric repulsion.95 The structural similarity of graphene/rGO with CNTs

389

suggests that stable graphene/rGO dispersion by the above mentioned macromolecules is

390

likely. Additional potential dispersants in aquatic environments include fulvic acid, lipids,

391

and aromatic macromolecules.

392

4. Toxicity of GFNs to aquatic organisms

393

The risk of GFNs to aquatic environments has received increasing attention, with these

394

materials now considered to be potential environmental pollutants of emerging concern.

395

Here, we address the following questions: (1) Which organisms in aquatic environments

396

are the most sensitive to GFNs? (2) How toxic is GFN in comparison to other carbon

397

materials (e.g., CNTs, black carbon)? (3) What are the main mechanisms responsible for

398

GFN-induced toxicity at the cellular level? (4) How would environmental factors

399

influence the toxicity? The recently published studies on GFNs toxicity to aquatic

400

organisms are presented in Table S5.

401

4.1. Toxicity to different organisms. The antimicrobial activity of GFNs have

402

attracted significant scientific interest as summarized in Table S6. Minimum inhibition

403

concentrations (MICs) of rGO against four types of bacteria were much lower than the

404

antibiotic kanamycin, with MIC values being 1-8 and 64-128 µg/mL, respectively.96

405

After exposure to GO and rGO (40 mg/L) for 2 h, gram-negative E. coli had viability

406

reductions of 69.3% and 47.4%, respectively.97 For another gram-negative bacterium, P.

407

aeruginosa, 2 h median lethal concentrations (LC50) of both GO and rGO were within the

408

range of 75-100 µg/mL.98 Krishnamoorthy et al. reported that gram-negative bacteria (E.

409

coli and S. typhimuirum) were more susceptible to GFNs (rGO) than gram-positive

19

ACS Paragon Plus Environment

Environmental Science & Technology

410

bacteria (B. subtilis and E. faecalis).96 The authors attributed the higher susceptibility to

411

the thinner peptidoglycan layer of gram-negative species (7-8 nm) as compared to gram-

412

positive microbes (20-80 nm). However, the opposite result was observed by Akhavan

413

and Ghaderi,99 where increased resistance of E. coli was attributed to the gram negative

414

outer membrane that is lacking in gram-positive organisms such as S. aureus. Therefore,

415

the role of cell wall structure in minimizing GFN toxicity is unclear. In another study on

416

carbon materials, B. subtilis was resistant to both MWCNT and graphite. However, the

417

gram-positive S. epidermis had similar or lower resistance as compared to the gram-

418

negatives E. coli and P. aeruginosa.100 This indicates that antibacterial activity of GFNs

419

is bacterial species-dependent rather than gram-dependent. Moreover, these two studies

420

employed different test approaches (broth microdilution96 vs. agar dilution99), which is

421

likely another reason for the different results.

422

In-vivo antibacterial activity of graphene (graphite nanoplatelet) was reported by

423

employing a unique bacterium-nematode infection model.101 In this study, the host C.

424

elegans was first infected with the pathogenic P. aeruginosa which expressed the green

425

fluorescent protein (gfp). After infection, the bacteria could be observed in nematodes by

426

fluorescence microscopy. Interestingly, bacteria in nematode bodies cannot be detected

427

after graphene exposure, presumably due to complete inactivation of P. aeruginosa by

428

the GFN. Fourier transform infrared spectroscopy (FTIR) mapping showed that graphene

429

sheets were distributed along the nematode body, including embryos, without causing any

430

adverse effects on the animal. Similarly, insignificant toxicity of GO to C. elegans was

431

observed.102 For crustaceans, the only toxicological report was on A. amphitrite larvae,

432

with a 48 h LC50 of 560 µg/mL.103 The main overt symptoms of toxicity were swimming

20

ACS Paragon Plus Environment

Page 20 of 61

Page 21 of 61

Environmental Science & Technology

433

speed reduction and settlement inhibition upon GO exposure. However, there were no

434

further investigations to elucidate the mechanisms of toxicity. Zebrafish, a model fish

435

species, exhibited good resistance to GFNs (GO and multi-functional graphene) in spite

436

of complete GFN distribution from head to tail.104,105 Notably, NPs (e.g., CuO)

437

accumulated in fish muscle,106 potentially raising food safety concerns but no information

438

is available on this point. Additional research efforts are needed to focus on sub-acute and

439

chronic toxic effects of GFNs to fish and other biota after longer-term exposure.

440

Algae may be one of most vulnerable organisms to NPs in the aquatic food-chain).79,107

441

One study on algal toxicity reported EC50s (72 h) of 1.14 and 2.25 mg/L for monolayer

442

graphene and graphene nanopowder, respectively.108 To date, there are no reports on the

443

interaction of GFNs with other organisms such as ciliates, yeasts or aquatic plants (Table

444

S6). However, the toxicity of GFNs (GO) to a series of terrestrial plants (cabbage, tomato

445

and red spinach) has been reported.109 Observed visible symptoms of phytotoxicity

446

included necrotic lesions and membrane damage of leaf cells after GO (500 mg/L)

447

exposure for 20 days; notably, the damage was dose-dependent (500-2000 mg/L). Similar

448

membrane damage was recently reported from an in vitro study with Arabidopsis

449

thaliana T87 cells at a GO exposure concentration of 40 mg/L.110 The above results on

450

plants were conducted under hydroponic conditions, where GFN behavior approximates

451

aquatic environments. In comparison with terrestrial plants, aquatic plants have thinner

452

cuticles, larger air spaces inside the mesophyll layer and more stomata on the epidermal

453

surfaces.111 On the basis of these differences, aquatic plants may be more sensitive to NPs

454

such as GFNs as compared to terrestrial species. However, no comparative study has

455

been performed to verify this hypothesis.

21

ACS Paragon Plus Environment

Environmental Science & Technology

456

4.2. GO and rGO in comparison with other carbon materials. GFNs can be

457

compared to other carbon materials where sufficient toxicological data exists. GO and

458

rGO, with smaller size and thinner layers, have higher antibacterial activities than

459

graphite oxide and graphite.97 Thickness is likely a more important factor because smaller

460

lateral size did not contribute higher antibacterial activity in another study from this

461

group.112 The importance of thickness is further confirmed by an investigation with

462

crustaceans (A. amphitrite),103 in which nano-size black carbon had much lower toxicity

463

than single-layer graphene. Graphene, with sharp edges, can damage plasma membranes

464

and cause cell death.97,110 Moreover, the lower adverse effects of black carbon than GFNs

465

were also found in a cellular Arabidopsis thaliana investigation.110 Taken together, one

466

can conclude that graphene materials are more toxic than conventional carbon materials

467

such as black carbon and graphite. Conversely, fullerene and CNTs share a similar sp2

468

carbon structure with graphene, but differ in shape or conformation. Zebrafish (Danio

469

rerio) and nematodes (C. elegans) exhibited severe growth inhibition after exposure to

470

fullerols (25 mg/L)105 and MWCNTs (100 mg/L),101 respectively. However, graphene had

471

minor effects at the same concentrations. Therefore, it seems that the toxicity of graphene

472

to aquatic organisms may not be as great as carbon nanomaterials such as fullerenes and

473

CNTs. Further work needs to be done to reveal the dominant properties of GFNs that lead

474

to this lower toxicity than CNTs and fullerols.

475

It is interesting and necessary to explore the toxicity difference of graphene/rGO with

476

its derivatives such as GO and carboxylated graphene. To date, the comparative studies

477

have focused on bacteria and human cells (Table S5). rGO consistently shows more

478

pronounced adverse effects than GO in these studies, with the exception of the two

22

ACS Paragon Plus Environment

Page 22 of 61

Page 23 of 61

Environmental Science & Technology

479

studies on bacterium E. coli97 and human endothelial cells.113 In these two studies, the

480

authors attribute the stronger cytotoxicity of GO to its greater dispersibility, smaller

481

aggregate size and higher surface functionality; all of which increases direct contact with

482

cells and induces greater intracellular oxidative stress. However, this explanation fails in

483

several other studies in Table S5, where higher rGO toxicity was demonstrated.98,99 The

484

following explanations are proposed for the higher observed toxicity of rGO. (1) Thinner

485

monolayer sheets. The thickness of the rGO monolayer (~0.34 nm) is much smaller than

486

that of GO (~1 nm)114 due to the removal of oxygen-containing groups. Even after

487

aggregation, there would be sharp edges exposed on the surfaces because of various

488

lateral sizes, resulting in stronger and more negative interactions with cellular membranes

489

as evidenced by the efflux of cytoplasmic RNA.99 (2) Better electrical conductivity. After

490

reduction, the sp2 hybridized electron structure on the graphitized sheets is restored,

491

which can be visualized by the reduction of the ID/IG ratio from Raman spectra. The

492

electrical conductivity of rGO are therefore improved relative to GO. This improvement

493

could lead to the greater damage of cellular membranes through increasing oxidization of

494

intracellular glutathione and subsequent oxidative stress99 or by directly promoting

495

electron transfer between rGO edges and biological membranes.97 (3) Stronger

496

hydrophobicity. It is known that bacteria cells tend to associate with hydrophobic

497

surfaces as compared to hydrophilic materials.112 Thus, the hydrophobic rGO sheets

498

could more readily interact with bacteria and other cells. Subsequent nutrient absorption

499

and gas exchange across the membrane will then be hindered. The covering of GFN

500

sheets on the surface of cells/organisms was found in several studies on bacteria,97,98

501

yeast115 and plant roots.109

23

ACS Paragon Plus Environment

Environmental Science & Technology

502

4.3. Toxicity mechanisms at the cellular level. On the basis of the existing literature,

503

we propose several cytotoxic mechanisms for graphene (both GO and rGO) in Figure 4.

504

The primary mechanism will likely be the synergy between direct contact and indirect

505

oxidative stress. After exposure, monolayer or a few-layer GFN (GO and rGO) sheets are

506

able to cut and penetrate cell membranes (also the cell wall, if present), and result in

507

direct physical membrane damage.99 Another type of physical damage by GFNs is

508

destructive extraction of lipid molecules via hydrophobic effects, which has been

509

observed for E. coli by both transmission electron microscopy and computer

510

simulation.116 In addition, the covering of cells by GFNs could result in indirect growth

511

inhibition by restricting ion/gas exchange on membranes as mentioned above.

512

Single graphene sheets tend to aggregate due to van der Waals forces. Small GFN

513

aggregates (~500 nm) could be internalized through endocytosis while larger aggregates

514

(up to 5 µm) via phagocytosis. As shown in Figure 4, GFN (GO) sheets enclosed by

515

vesicles could be deposited into vacuoles after uptake.110 The internalized cytoplasmic

516

GFNs can also induce oxidative stress.98,109 In addition to the possible production of

517

reactive oxygen species (ROS) by GFNs themselves,117 ROS are mainly produced by

518

interacting of GFNs with mitochondria, chloroplasts (for plant and alga cells),

519

peroxisomes and oxidases.110 As an example, graphene may interrupt electron transfer in

520

the mitochondria by causing the overproduction of ROS. Normally, the antioxidant

521

system of cells can balance ROS production. However, excessive ROS could result in

522

subsequent cell death through a number of pathways (e.g., organelle dysfunction). Begum

523

and Fugetsu observed the dysfunction of mitochondria in plant cells (A. thaliana) but it is

524

unclear if the damage was caused directly by physical contact or indirectly induced by

24

ACS Paragon Plus Environment

Page 24 of 61

Page 25 of 61

Environmental Science & Technology

525

oxidative stress.110 Oxidative stress may also result in lipid peroxidation96 and the loss of

526

cell integrity. Another important toxicity mechanism is DNA damage, including nucleic

527

acid fragmentation and destruction. DNA damage has been detected in human cells after

528

incubation with GFNs,118,119 whereas no DNA damage was observed in plant cells.110

529

Similar to mitochondrial dysfunction, DNA damage may also be caused by either

530

physical damage and/or oxidative stress. GFNs internalized by endocytosis are first

531

enclosed in endosomes, but then may be released into the cytoplasm120 (Figure 4).

532

Therefore, physical damage could result from the GFN sheets which enter the cells by

533

direct penetration or from endosomal release.

534

At the cellular level, intercellular GFNs may also be subject to biodegradation and

535

exclusion after uptake (Figure 4). After internalized by macrophages, GFNs may be

536

degraded in late endosomes and lysosomes that possess digestive environments. It has

537

been reported that both in-vivo and in-vitro macrophages can degrade carboxyl

538

functionalized graphene;121 a detailed degradation mechanism will be illustrated in the

539

next section. To date, there is no literature on biodegradation of GFNs in aquatic

540

organisms. Exclusion is a type of self-defense of cells; we have quantified the exclusion

541

of CuO NPs from human cells (A549).122 Intercellular GFNs may be excluded from

542

aquatic organisms, although further investigation is needed to document this process.

543

4.4. Influence of environmental factors. To date, there is little information regarding

544

toxicity of GFNs under different environmental conditions. Current research on other

545

NPs suggests that the GFN toxicity will depend on both inherent material properties and

546

water chemistry.123 Our previous study found that DOM coating mitigated CuO NPs

547

toxicity to bacteria (E. coli) due to the hindrance of direct contact between particles and

25

ACS Paragon Plus Environment

Environmental Science & Technology

548

bacteria.79 Interestingly, the opposite result was obtained with algae (M. aeruginosa),

549

suggesting the influence of DOM may be species-dependent. GFNs are able to adsorb

550

DOM (Table S1), and this surface coating is expected to alter material toxicity to aquatic

551

organisms. In addition, I.S., ion species and pH of the water are likely to influence the

552

GFN toxicity by altering its dispersion state.

553

Co-existence of GFNs with pollutants in aquatic environments could alter the transport,

554

accumulation and toxicity of both constituents. Zhang et al. reported enhanced cadmium

555

accumulation in fish by co-addition of TiO2 NPs; the higher accumulation factor of TiO2

556

directly led to enhanced uptake of the heavy metal.124 In contrast, the bioavailable

557

fraction of a pollutant (e.g., HOCs) can be reduced if accumulation factor of the NPs

558

carrier in the test organism is lower than the adsorbed pollutant.125 Therefore, for GFNs,

559

the toxicity could be synergistic for both GFNs and the adsorbed pollutant, depending

560

both on GFN sorption capacity and the accumulation factors for both constituents in the

561

test organism. Interestingly, a recent study revealed a stress-induced toxicity of GO to

562

nematodes (C. elegans)102 by impairment of the inherent antioxidant defense system

563

under additional oxidative or thermal stress through a multiple-path mechanism. This

564

indicates that possible cooperative toxicity of GFNs may occur under other

565

environmental stresses, including co-contaminants. In addition, GO could be transformed

566

to rGO under UV irradiation126 or in the presence of reducing chemicals such as sulfur-

567

containing compounds,127 resulting in changes in the pattern of GO toxicity in aquatic

568

environments. The transformation process is reviewed in detail in the next section.

569

5. Transformation and Biodegradation of GFNs in aqueous environments

26

ACS Paragon Plus Environment

Page 26 of 61

Page 27 of 61

Environmental Science & Technology

570

GFNs may be transformed and degraded under ambient conditions after release into

571

aqueous environments or after uptake by biota. Until now, direct evidence on

572

transformation and degradation of GFNs in natural environments has not been reported.

573

However, there are a number of investigations under laboratory conditions (Table S7),

574

which imply the possible transformation and degradation of GO in aqueous environments.

575

5.1. Transformation of GO to rGO in aquatic solution. GO is considered as a

576

metastable material which undergoes spontaneous reduction.128,129 This reduction reaction

577

could be accelerated in the presence of UV irradiation, inorganic and organic reductants,

578

and biota (Figure 5A). Under UV irradiation, GO sheets were reduced to rGO in the

579

presence of TiO2 as a photocatalyst.130,131 This deoxygenation process was nearly

580

completed in only 2 h. In the absence of a photocatalyst, the reduction of GO aqueous

581

solution was still achieved, but only after longer irradiation times (5 h).126 However, this

582

does suggest that GO could be reduced in natural environments under longer-term UV-

583

rich irradiation.

584

Zero-valent metals such as Al132 and Fe133 were observed to be efficient reductants for

585

the deoxygenation of GO sheets in acidic solution. The possible mechanism can be

586

expressed as: GO + a·H+ + b·n·M → rGO + b·Mn+ + c·H2O. In neutral solution, the

587

reduction was not obvious after 2-days of observation. This phenomenon was explained

588

by the formation of metal oxide film on the surface of zero-valent metal, thereby

589

protecting against further oxidation.132

590

Another group of inorganic reductants important in anoxic environments such as

591

eutrophic waters are sulfur-containing compounds.134 All evaluated compounds (NaHSO3,

592

Na2SO3, Na2S2O3, Na2S, SOCl2, and SO2) reduced GO, with NaHSO3 being the most

27

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 61

593

effective.127 SOCl2, which could form HSO3- when reacting with water, also showed high

594

reducing activity towards GO as indicated by the C/O ratio of the resulting rGO. The

595

transformation of GO under nitrate-reducing conditions has not yet been reported.

596

A number of organic antioxidants have been used for GO reduction in the synthesis of

597

GFNs. Polyphenols and vitamins (C and E) are two common types of natural antioxidants,

598

and both have been investigated for GO reduction. Wang et al. reported the efficient

599

reduction of GO in tea solution with the assistance of heating; the polyphenol

600

components were the main contributors to the observed transformation.135 Moreover, the

601

reducing ability of polyphenols was enhanced with iron amendment because of electron

602

donation into GO sheets and promotion of proton dissociation from polyphenols.136 For

603

vitamins, L-ascorbic acid (vitamin C) was found to yield highly reduced rGO.91,137

604

However, the reducing activity of vitamin E remains unexplored. In addition, amino acids

605

(glycine,

606

polysaccharides (natural cellulose),141 and protein (BSA, lysozyme)142,143 have been

607

shown to remove oxygen moieties on GO sheets upon reduction.

L-cysteine),138,139

saccharides

(glucose,

fructose

and

sucrose),140

608

The transformation from GO to rGO may also take place under direct interaction with

609

aquatic organisms. In aqueous solution, peeled wild carrot root was able to reduce GO;

610

the mechanism of reduction was attributed to the presence of endophytic microorganisms,

611

as evidenced by the control experiments treated with microbial inhibitors.144 Therefore,

612

this reduction most likely occurred after accumulation within the plant. In addition,

613

microorganisms could also independently reduce GO without the participation of plant

614

host. Salas et al. reported the bacterially mediated reduction of GO for the first time using

615

the environmental microbe Shewanella;145 this was verified by both Wang et al.146 and

28

ACS Paragon Plus Environment

Page 29 of 61

Environmental Science & Technology

616

Jiao et al.147 More importantly, reduction by Escherichia coli, one of the most widely

617

present stains in aqueous environments, was recently observed.148 The mechanism of

618

bacterial transformation of GO involves reduction during bacterial respiration145 and

619

glycolysis.148 During these processes, electrons are generated and transferred from the

620

cell interior to GO, which acts as an external electron acceptor. Salas et al. found that

621

during respiration among the primary protein components, inner-membrane-anchored c-

622

type cytochrome (CymA) was found to be dispensable.145 This is inconsistent with the

623

findings of Jiao et al.,147 where CymA was as important as other proteins, including the

624

periplasmic (MtrA) and outer-membrane-anchored (MtrB) c-type cytochromes. There is

625

only one report on biotransformation of GO mediated by fungi.149 In this study, the epoxy

626

functionalities of GO were easily coupled with the amine groups of nicotinamide adenine

627

dinucleotide phosphate (NADPH) present in yeast, leading to the removal of oxygen

628

functionalities from the GO surface. The control group treated with yeast inhibitors was

629

conducted simultaneously, further confirming the reduction of GO sheets via yeast

630

activity. Last, poly(norepinephrine), a marine mussel-derived polymer, was shown to be

631

an effective agent for GO reduction.150

632

In order to achieve a reasonable reduction within a short time (1 min-3 days),

633

assistance by heating, sonication, specific solution pHs, and substrates (H2O2) is often

634

employed. In natural environments, reduction and transformation of GO could occur

635

spontaneously over time without amendment,128 but the reaction rates will likely be quite

636

slow. In addition to environmental media as discussed above, GO transformation is

637

expected to be promoted under other reducing conditions (e.g., sediments in a eutrophic

638

pond),134 or in the presence of other environmental reductants such as humic substances,

29

ACS Paragon Plus Environment

Environmental Science & Technology

639

flavonoids, ferulic acid, and carvacrol;151,152 however, no reports on these reducing

640

conditions have been published.

641

5.2. Degradation of GFNs in aqueous solution. Photodegradation of GO has been

642

observed in combination with TiO2 NPs in ethanol.153 In this study, GO sheets were

643

completely reduced to rGO after 2 h UV irradiation as indicated by the increasing

644

graphitized sp2 structure on graphene oxides, which is in agreement with the finding of

645

Williams et al.130 and Kim et al.131 It should be noted that the observed photodegradation

646

was only reported in ethanol; further research needs to be done on the photodegradation

647

of GFNs in aqueous solutions (indicated by the dotted arrow in Figure 5B).

648

We are aware of only two studies on biodegradation of GFNs in aqueous solution

649

(Figure 5B). Kotchey et al. reported the biodegradation of GO sheets to rGO via

650

enzymatic catalysis.21 The oxidative biodegradation by plant-derived horseradish

651

peroxidase (HRP) resulted in a number of defects and holes on the graphitic lattice of GO,

652

and the final product of oxidation was CO2, according to oxidation product analysis. The

653

active site for catalyzed biodegradation was the heme group of the HRP molecule;

654

molecular modeling suggests that the primary binding position on GO sheets was the

655

basal plane, as opposed to the rather the edge which lacks sufficient sites for stable GO-

656

HRP contacts. Furher biodegradation of rGOs was not observed in this study, which can

657

be explained as follows: (1) After binding, the distance between active heme group and

658

basal plane of rGO (11.5 Å) was longer than GO (8.7 Å), which led to unsuitable

659

interaction between Heme group and rGO sheets; (2) The functionalized groups (e.g.,

660

epoxide, hydroxyl, and carboxyl groups) on the basal plane of GFNs may act as the

661

originating point of oxidization and biodegradation by HRP, and the basal plane of rGO

30

ACS Paragon Plus Environment

Page 30 of 61

Page 31 of 61

Environmental Science & Technology

662

sheet lacks oxygen-containing groups (Figure 1). The lack of rGO degradation was

663

consistent with that of pristine CNTs, which are also not enzymatically degraded by

664

HRP.154,155,156 The CNT biodegradation studies did reveal that two human derived

665

enzymes, eosinophil peroxidase157 and neutrophil enzyme myeloperoxidase,158 degraded

666

carboxylated CNTs; however, the degradative capacity of these two enzymes has not

667

been verified for GFNs. Another important study revealed in-vivo biodegradation via

668

confocal Raman imaging of carboxylated graphene in tissue-bound macrophages after

669

uptake by mice.121 The structural disorder of graphene sheet was detected after 8 days,

670

and carboxylated graphene was degraded to amorphous carbon over 3 months. The in-

671

vitro study on macrophages reiterates the potential biodegradation of carboxylated

672

graphene. Interestingly, the degradation of carboxylated graphene began from the edges

673

in this study, rather than from basal plane as reported by Kotchey et al.21 This difference

674

on the primary action sites of graphene may result from the different biodegradation

675

systems (enzyme vs. macrophage); a topic that merits further investigation. In aquatic

676

environments, similar GFN biodegradation may also take place after accumulation by

677

select biota (e.g., fish) that possess cellular macrophages.

678

Thermal degradation has been shown to be effective for select polymer

679

nanocomposites.159 GO can be reduced by microwave irradiation for 1 min160 or heating

680

at 80-90 °C for 5 h in alkaline solution.161 However, the obtained rGO was resistant to

681

further degradation. Black carbon has a similar graphite-like structure as graphene;

682

residence times of black carbon in soil162 and marine waters163 were reported as 2000 and

683

2400-13900 years, respectively. If the size and surface area of these materials were not

31

ACS Paragon Plus Environment

Environmental Science & Technology

684

considered, the possible residence time of graphene and rGO may be expected to be

685

thousand years in the similar environments.

686

6. Challenges and Perspectives

687

We have reviewed four critical processes determining GFNs fate and disposition in

688

aquatic environments. GO preferentially adsorbs metal ions and positively charged

689

organic molecules, whereas graphene/rGO adsorb hydrophobic and aromatic molecules.

690

Because of their dissimilar surface properties, the materials exhibit distinct

691

dispersion/aggregation behaviors under environmentally relevant conditions. The toxicity

692

and transformation of GFNs has been summarized based on the currently available data.

693

However, research on the environmental behavior of GFNs is still at an early stage.

694

Considerable challenges limit the understanding the environmental fate, exposure and

695

risk of GFNs; four such challenges are identified as follows:

696

(1) Depending on synthesis routes,2,164 assisted processes165 and original graphite

697

materials;72 GFNs structural and surface properties can vary considerably. These

698

properties include lateral size, C/O ratio, and structural defects, all of which induce

699

different colloidal behaviors, adsorption capabilities and toxicities of graphene and

700

GO.22,66,72,112,165 Unfortunately, the existing literature often contains insufficient

701

information on synthesis conditions. Therefore, a more thorough and detailed description

702

on production procedures and material characterization are required.

703

(2) DLVO theory has been widely applied in colloidal stability studies. In traditional

704

DLVO theory, two colloid particles are assumed to be spherical, and the interactions are

705

dominated by summed van der Waals and EDL forces.81 The features of GFNs, including

706

chemical composition, shape and surface coating; challenge classical DLVO.55,166 For

32

ACS Paragon Plus Environment

Page 32 of 61

Page 33 of 61

Environmental Science & Technology

707

example, due to the two-dimensional nature, the inter-sheet attraction of GFNs could be

708

stronger than that of spherical particles because higher amounts of atoms on GFNs are in

709

close proximity to each other.81 In actual environments, bridging, steric and hydrophobic

710

forces are also present between individual GFN sheets, thus complicating the colloidal

711

system. In this case, expanded theoretical approaches are needed to accurately predict the

712

aggregation/dispersion behavior of GFNs. Hotze and coworkers have summarized

713

expanded DLVO models and the related single forces (e.g, van der Waals attraction,

714

bridging attraction) between colloidal particles,81 which will be helpful in overcoming

715

this challenge. Importantly, multiple forces co-exist in most GFN aggregation cases.

716

Therefore, new approaches should expand/modify existing DLVO theories to combine all

717

relevant forces in a specific model and distinguish their contributions for GFNs and other

718

NPs systems.

719

(3) Initially, GO was synthesized as a byproduct in the oxidative process during

720

graphene production using the oxidation/exfoliation approach.9,70 Currently, a number of

721

distinctive properties of GO that differ from graphene have been recognized, but are

722

poorly understood. It is known that GO can act as surfactant to lower interfacial energy at

723

air-water interface and can emulsify organic solvents at water-oil interfaces.27,70 However,

724

these surfactant sheets are probably unable to form micelles and do not exhibit salient

725

CMC features.74 Further, it is unknown whether GO sheets are able to solubilize/bind

726

organic pollutants similar to conventional chemical surfactants. GO sheets can also

727

spontaneously degrade to humic acid-like materials during long-time exposure (two

728

months) in water.128 Consequently, what surface activity and adsorption behavior will

729

result when the GO sheets begin to transform/degrade? What is the fate of GO in natural

33

ACS Paragon Plus Environment

Environmental Science & Technology

730

aquatic environments? GO behavior is not currently predictable due to lack of relevant

731

data; new approaches and analytical tools are needed to answer the above questions. For

732

example, a passive dosing technique65 is suitable to investigate the solubilization effect of

733

GO towards organic pollutants, especially when the dispersion of GO sheets is decreased

734

during the transformation/degradation period. Long-term (> 2 months) degradation

735

experiments could be conducted to explore the spontaneous degradation of GO by

736

controlling solution conditions such as sunlight irradiation, pHs and temperature.

737

(4) We reviewed the adsorption of environmental adsorbates such as organic molecules

738

and metal ions and their effects on GFN dispersion/aggregation. However, the direct

739

relationship between adsorption and dispersion is poorly understood. In the presence of

740

organic molecules or metal ions, dispersion/aggregation of GFNs could be altered as

741

reviewed above, but the influence of dispersion alteration on GFNs toxicity is poorly

742

understood. Furthermore, despite high adsorption capacity of GFNs, the risk and toxicity

743

of co-existent GFNs and other contaminants are unknown. A complete understanding of

744

the interactions among adsorption, dispersion and toxicity warrant further research in the

745

future.

746 747

ASSOCIATED CONTENT

748

Supporting Information.

749

Seven tables. This material is available free of charge via the Internet at

750

http://pubs.acs.org.

751 752

AUTHOR INFORMATION

753

Corresponding Author

34

ACS Paragon Plus Environment

Page 34 of 61

Page 35 of 61

Environmental Science & Technology

754

*E-mail: [email protected] (Prof. Xing)

755

Notes

756

The authors declare no competing financial interest.

757 758 759

ACKNOWLEDGMENTS This research was in part supported by NSFC (41325013, 41328003, 41120134004),

760

USDA-AFRI (2011-67006-30181), and USDA-AFRI Hatch program (MAS 00978).

761 762

References:

763

(1) Geim, A. K.; Novoselov, K. S. The rise of graphene. Nat. Mater. 2007, 6, 183–191.

764

(2) Novoselov, K. S.; Fal, V. I.; Colombo, L.; Gellert, P. R.; Schwab, M. G.; Kim, K.

765 766 767

A roadmap for graphene. Nature 2012, 490, 192–200. (3) Wu, J.; Pisula, W.; Müllen K. Graphenes as potential material for electronics. Chem. Rev. 2007, 107, 718–747.

768

(4) Yang, W.; Ratinac, K. R.; Ringer, S. P.; Thordarson, P.; Gooding, J. J.; Braet, F.

769

Carbon nanomaterials in biosensors: Should you use nanotubes or graphene? Angew.

770

Chem. Int. Ed. 2010, 49, 2114–2138.

771 772

(5) Kim, J.; Cote, L. J.; Huang, J. Two dimensional soft material: New faces of graphene oxide. Acc. Chem. Res. 2012, 45, 1356–1364.

773

(6) Bianco, A.; Cheng, H. M.; Enoki, T.; Gogotsi, Y.; Hurt, R. H.; Koratkar, N.;

774

Kyotani T.; Monthioux, M.; Park C. R.; Tascon J. M. D.; Zhang, J. All in the graphene

775

family–A recommended nomenclature for two-dimensional carbon materials. Carbon

776

2013, 65, 1–6.

35

ACS Paragon Plus Environment

Environmental Science & Technology

777 778 779 780 781 782

(7) Chen, D.; Feng, H.; Li, J. Graphene oxide: Preparation, functionalization, and electrochemical applications. Chem. Rev. 2012, 112, 6027–6053. (8) Loh, K. P.; Bao, Q.; Ang, P. K.; Yang, J. The chemistry of graphene. J. Mater. Chem. 2010, 20, 2277–2289. (9) Dreyer, D. R.; Park, S.; Bielawski, C. W.; Ruoff, R. S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240.

783

(10) Segal, M. Selling graphene by the ton. Nat. Nanotechnol. 2009, 4, 612–614.

784

(11) Arvidsson, R.; Molander, S.; Sandén, B. A. Review of potential environmental

785

and health risks of the nanomaterial graphene. Hum. Ecol. Risk Assess. 2013, 19, 873–

786

887.

787

(12) Bernard, C.; Nguyen, T.; Pelligrin, B.; Holbrook, R. D.; Zhao, M.; Chin, J. Fate

788

of graphene in polymer nanocomposite exposed to UV radiation. J. Phys.: Conf. Ser.

789

2011, 304, 012063.

790

(13) Zhao, G.; Li, J.; Ren, X., Chen, C., Wang, X. Few-layered graphene oxide

791

nanosheets as superior sorbents for heavy metal ion pollution management. Environ. Sci.

792

Technol. 2011, 45, 10454–10462.

793

(14) Zhang, N.; Qiu, H.; Si, Y.; Wang, W.; Gao, J. Fabrication of highly porous

794

biodegradable monoliths strengthened by graphene oxide and their adsorption of metal

795

ions. Carbon 2011, 49, 827–837.

796

(15) Upadhyay, R. K.; Soin, N.; Roy, S. S. Role of graphene/metal oxide composites

797

as photocatalysts, adsorbents and disinfectants in water treatment: a review. RSC Adv.

798

2014, 4, 3823–3851.

36

ACS Paragon Plus Environment

Page 36 of 61

Page 37 of 61

Environmental Science & Technology

799

(16) Wang, Y.; Gao, S.; Zang, X.; Li, J.; Ma, J. Graphene-based solid-phase extraction

800

combined with flame atomic absorption spectrometry for a sensitive determination of

801

trace amounts of lead in environmental water and vegetable samples. Anal. Chim. Acta

802

2012, 716, 112–118.

803 804

(17) Mishra, A. K.; Ramaprabhu, S. Functionalized graphene sheets for arsenic removal and desalination of sea water. Desalination 2011, 282, 39–45.

805

(18) Sun, H.; Liu, S.; Zhou, G.; Ang, H. M.; Tadé, M. O.; Wang, S. Reduced graphene

806

oxide for catalytic oxidation of aqueous organic pollutants. ACS Appl. Mater. Interfaces

807

2012, 4, 5466−5471.

808

(19) Gao, W.; Majumder, M.; Alemany, L. B.; Narayanan, T. N.; Ibarra, M. A.;

809

Pradhan, B. K.; Ajayan, P. M. Engineered graphite oxide materials for application in

810

water purification. ACS Appl. Mater. Interfaces 2011, 3, 1821–1826.

811

(20) Chowdhury, I.; Duch, M. C.; Mansukhani, N. D.; Hersam, M. C.; Bouchard, D.

812

Deposition and release of graphene oxide nanomaterials using a quartz crystal

813

microbalance. Environ. Sci. Technol. 2014, 48, 961–969.

814

(21) Kotchey, G. P.; Allen, B. L.; Vedala, H.; Yanamala, N.; Kapralov, A. A.; Tyurina,

815

Y. Y.; Klein-Seetharaman L.; Kagan V. E.; Star, A. The enzymatic oxidation of graphene

816

oxide. ACS Nano 2011, 5, 2098–2108.

817

(22) Apul, O. G.; Wang, Q.; Zhou, Y.; Karanfil, T. Adsorption of aromatic organic

818

contaminants by graphene nanosheets: Comparison with carbon nanotubes and activated

819

carbon. Water Res. 2013, 47, 1648–1654.

37

ACS Paragon Plus Environment

Environmental Science & Technology

820

(23) Li, Y.; Du, Q.; Liu, T.; Sun, J.; Jiao, Y.; Xia, Y.; Xia, L.; Wang, Z.; Zhang, W.;

821

Wang, K.; Zhu, H.; Wu, D. Equilibrium, kinetic and thermodynamic studies on the

822

adsorption of phenol onto graphene. Mater. Res. Bull. 2012, 47, 1898–1904.

823

(24) Yang, Z.; Yan, H.; Yang, H.; Li, H.; Li, A.; Cheng, R. Flocculation performance

824

and mechanism of graphene oxide for removal of various contaminants from water.

825

Water Res. 2013, 47, 3037–3046.

826 827 828 829 830 831

(25) Liu, J. Adsorption of DNA onto gold nanoparticles and graphene oxide: surface science and application. Phys. Chem. Chem. Phys. 2012, 14, 10485–10496. (26) Wei, H.; Yang, W.; Xi, Q.; Chen, X. Preparation of Fe3O4@graphene oxide core– shell magnetic particles for use in protein adsorption. Mater. Lett. 2012, 82, 224–226. (27) Kim, J.; Cote, L. J.; Kim, F.; Yuan, W.; Shull, K. R.; Huang, J. Graphene oxide sheets at interfaces. J. Am. Chem. Soc. 2010, 132, 8180–8186.

832

(28) Yang, X.; Li, J.; Wen, T.; Ren, X.; Huang, Y.; Wang, X. Adsorption of

833

naphthalene and its derivatives on magnetic graphene composites and the mechanism

834

investigation. Colloids Sur. A 2013, 422, 118–125.

835

(29) Gao, Y.; Li, Y.; Zhang, L.; Huang, H.; Hu, J.; Shah, S. M.; Su, X. Adsorption and

836

removal of tetracycline antibiotics from aqueous solution by graphene oxide. J. Colloid

837

Interface Sci. 2012, 368, 540–546.

838

(30) Hao, L.; Song, H.; Zhang, L.; Wan, X.; Tang, Y.; Lv, Y. SiO2/graphene

839

composite for highly selective adsorption of Pb (II) ion. J. Colloid Interface Sci. 2012,

840

369, 381–387.

841 842

(31) Shi, G.; Ding, Y.; Fang, H. Unexpectedly strong anion–π interactions on the graphene flakes. J. Comput. Chem. 2012, 33, 1328–1337.

38

ACS Paragon Plus Environment

Page 38 of 61

Page 39 of 61

843 844

Environmental Science & Technology

(32) Pan, B.; Xing, B. Adsorption mechanisms of organic chemicals on carbon nanotubes. Environ. Sci. Technol. 2008, 42, 9005–9013.

845

(33) Chowdhury, I.; Duch, M. C.; Manuskhani, N. D.; Hersam, M. C.; Bouchard, D.

846

Colloidal properties and stability of graphene oxide nanomaterials in the aquatic

847

environment. Environ. Sci. Technol. 2013, 47, 6288–6296.

848 849

(34) Li, D.; Mueller, M. B.; Gilje, S.; Kaner, R. B.; Wallace, G. G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 2008, 3, 101–105.

850

(35) Sharma, P.; Das, M. R. Removal of a cationic dye from aqueous solution using

851

graphene oxide nanosheets: Investigation of adsorption parameters. J. Chem. Eng. Data

852

2013, 58, 151–158.

853

(36) Madadrang, C. J.; Kim, H. Y.; Gao, G.; Wang, N.; Zhu, J.; Feng, H.; Gorring, M.;

854

Kasner, M. L.; Hou, S. Adsorption behavior of EDTA-graphene oxide for Pb (II) removal.

855

ACS Appl. Mater. Interfaces 2012, 4, 1186–1193.

856 857 858 859

(37) Mi, X.; Huang, G.; Xie, W.; Wang, W.; Liu, Y.; Gao, J. Preparation of graphene oxide aerogel and its adsorption for Cu2+ ions. Carbon 2012, 50, 4856–4864. (38) Xu, J.; Wang L.; Zhu, Y. Decontamination of bisphenol A from aqueous solution by graphene adsorption. Langmuir 2012, 28, 8418–8425.

860

(39) Zhang, C.; Wu, L.; Cai, D.; Zhang, C.; Wang, N.; Zhang, J.; Wu, Z. Adsorption

861

of polycyclic aromatic hydrocarbons (fluoranthene and anthracenemethanol) by

862

functional graphene oxide and removal by pH and temperature-sensitive coagulation.

863

ACS Appl. Mater. Interfaces 2013, 5, 4783–4790.

39

ACS Paragon Plus Environment

Environmental Science & Technology

864

(40) Ji, L.; Chen, W.; Xu, Z.; Zheng, S.; Zhu, D. Graphene nanosheets and graphite

865

oxide as promising adsorbents for removal of organic contaminants from aqueous

866

solution. J. Environ. Qual. 2013, 42, 191–198.

867

(41) Deng, X.; Lü, L.; Li, H.; Luo, F. The adsorption properties of Pb (II) and Cd (II)

868

on functionalized graphene prepared by electrolysis method. J. Hazard. Mater. 2010, 183,

869

923–930.

870

(42) Li, Z.; Chen, F.; Yuan, L.; Liu, Y.; Zhao, Y.; Chai, Z.; Shi, W. Uranium(VI)

871

adsorption on graphene oxide nanosheets from aqueous solutions. Chem. Eng. J. 2012,

872

210, 539–546.

873 874 875 876 877 878

(43) Piao, L.; Liu, Q.; Li, Y.; Wang, C. Adsorption of L-phenylalanine on singlewalled carbon nanotubes, J. Phys. Chem. C 2008, 112, 2857–2863. (44) Zheng, J.; Ren, Z.; Guo, P.; Fang, L.; Fan, J. Diffusion of Li+ ion on graphene: A DFT study. Appl. Surf. Sci. 2011, 258, 1651–1655. (45) Krepel, D.; Hod, O. Lithium adsorption on armchair graphene nanoribbons. Surf. Sci. 2011, 605, 1633–1642.

879

(46) AlZahrani, A. Z. First-principles study on the structural and electronic properties

880

of graphene upon benzene and naphthalene adsorption. Appl. Surf. Sci. 2010, 257, 807–

881

810.

882 883 884 885

(47) Yang, K.; Xing, B. Adsorption of organic compounds by carbon nanomaterials in aqueous phase: Polanyi theory and its application. Chem. Rev. 2010, 110, 5989–6008. (48) Xia, X. R.; Monteiro-Riviere, N. A.; Riviere, J. E. An index for characterization of nanomaterials in biological systems. Nat. Nanotechnol. 2010, 5, 671–675.

40

ACS Paragon Plus Environment

Page 40 of 61

Page 41 of 61

Environmental Science & Technology

886

(49) Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.;

887

Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S.; Geim, A. K. Raman spectrum of

888

graphene and graphene layers. Phys. Rev. Lett. 2006, 97, 187401.

889

(50) Wang, C.; Li, H.; Liao, S.; Zheng, H.; Wang, Z.; Pan, B.; Xing, B. Coadsorption,

890

desorption

hysteresis

and

sorption

thermodynamics

of

sulfamethoxazole

891

carbamazepine on graphene oxide and graphite. Carbon 2013, 65, 243–251.

and

892

(51) Xia, X. R.; Monteiro-Riviere, N. A.; Mathur, S.; Song, X.; Xiao, L.; Oldenberg, S.

893

J.; Fadeel, B.; Riviere, J. E. Mapping the surface adsorption forces of nanomaterials in

894

biological systems. ACS Nano 2011, 5, 9074–9081.

895

(52) Zhao, J.; Wang, Z.; Mashayekhi, H.; Mayer, P.; Chefetz, B.; Xing, B. Pulmonary

896

surfactant

897

solubilization and competition as examined by passive dosing technique. Environ. Sci.

898

Technol. 2012, 46, 5369–5377.

899 900 901 902

suppressed

phenanthrene

adsorption

on

carbon

nanotubes

through

(53) Pei, S.; Cheng, H. M. The reduction of graphene oxide. Carbon 2012, 50, 3210– 3228. (54) Camden, A. N.; Barr, S. A.; Berry, R. J. Simulations of peptide-graphene interactions in explicit water. J. Phys. Chem. B 2013, 117, 10691–10697.

903

(55) Lin, S.; Shih, C. J.; Strano, M. S.; Blankschtein, D. Molecular insights into the

904

surface morphology, layering structure, and aggregation kinetics of surfactant-stabilized

905

graphene dispersions. J. Am. Chem. Soc. 2011, 133, 12810–12823.

906 907

(56) Radic, S.; Geitner, N.; Podila, R.; Ke, P. C.; Ding, F. Competitive Binding of Natural Amphiphiles with Graphene derivatives. Sci. Rep. 2013, 3, 2273.

41

ACS Paragon Plus Environment

Environmental Science & Technology

908

(57) Sun, Y.; Yang, S.; Zhao, G.; Wang, Q.; Wang, X. Adsorption of polycyclic

909

aromatic hydrocarbons on graphene oxides and reduced graphene oxides. Chem. Asian J.

910

2013, 8, 2755–2761.

911

(58) Ren, X.; Li, J.; Tan, X.; Wang, X. Comparative study of graphene oxide,

912

activated carbon and carbon nanotubes as adsorbents for copper decontamination. Dalton

913

Trans. 2013, 42, 5266–5274.

914

(59) Li, J.; Zhang, S.; Chen, C.; Zhao, G.; Yang, X.; Li, J.; Wang, X. Removal of Cu

915

(II) and fulvic acid by graphene oxide nanosheets decorated with Fe3O4 nanoparticles.

916

ACS Appl. Mater. Interfaces 2012, 4, 4991–5000.

917

(60) Hartono, T.; Wang, S.; Ma, Q.; Zhu, Z. Layer structured graphite oxide as a novel

918

adsorbent for humic acid removal from aqueous solution. J. Colloid Interface Sci. 2009,

919

333, 114–119.

920 921 922 923

(61) Lin, D.; Tian, X.; Li, T.; Zhang, Z.; He, X.; Xing, B. Surface-bound humic acid increased Pb2+ sorption on carbon nanotubes. Environ. Pollut. 2012, 167, 138–147. (62) Yang, K.; Xing, B. Desorption of polycyclic aromatic hydrocarbons from carbon nanomaterials in water. Environ. Pollut. 2007, 145, 529–537.

924

(63) Ma, X.; Anand, D.; Zhang, X.; Talapatra, S. Adsorption and desorption of

925

chlorinated compounds from pristine and thermally treated multiwalled carbon nanotubes.

926

J. Phys. Chem. C 2011, 115, 4552–4557.

927 928

(64) Zhang, Y.; Tang, Y.; Li, S.; Yu, S. Sorption and removal of tetrabromobisphenol A from solution by graphene oxide. Chem. Eng. J. 2013, 222, 94–100.

42

ACS Paragon Plus Environment

Page 42 of 61

Page 43 of 61

Environmental Science & Technology

929

(65) Zhao, J.; Wang, Z.; Zhao, Q.; Xing, B. Adsorption of phenanthrene on multi-layer

930

graphene as affected by surfactant and exfoliation. Environ. Sci. Technol. 2014, 48, 331–

931

339.

932

(66) Creighton, M. A.; Rangel-Mendez, J. R.; Huang, J. X.; Kane, A. B.; Hurt, R. H.;

933

Graphene-induced adsorptive and optical artifacts during in vitro toxicology assays.

934

Small 2013, 9, 1921–1927.

935 936 937 938

(67) Jachak, A. C.; Creighton, M.; Qiu, Y.; Kane, A. B.; Hurt, R. H. Biological interactions and safety of graphene materials. MRS bulletin 2012, 37, 1307–1313. (68) Wang, Z.; Zhao, J.; Li, F.; Gao, D.; Xing, B. Adsorption and inhibition of acetylcholinesterase by different nanoparticles. Chemosphere 2009, 77, 67–73.

939

(69) Petersen, E.; Henry, T.; Zhao, J.; MacCuspie, R.; Kirschling, T.; Dobrovolskaia,

940

M.; Hackley, V.; Xing, B.; White, J. Identification and avoidance of potential artifacts in

941

nanomaterial ecotoxicity measurements. Environ. Sci. Technol. 2014, 48, 4226–4246.

942 943 944 945 946 947

(70) Cote, L. J.; Kim, J.; Tung, V. C.; Luo, J.; Kim, F.; Huang, J. Graphene oxide as surfactant sheets. Pure Appl. Chem. 2011, 83, 95–110. (71) Si, Y.; Samulski, E. T. Synthesis of water soluble graphene. Nano Lett. 2008, 8, 1679–1682. (72) Luo, J.; Cote, L. J.; Tung, V. C.; Tan, A. T.; Goins, P. E.; Wu, J.; Huang, J. Graphene oxide nanocolloids. J. Am. Chem. Soc. 2010, 132, 17667–17669.

948

(73) Yoon, K. Y.; An, S. J.; Chen, Y.; Lee, J. H.; Bryant, S. L.; Ruoff, R. S.; Huh C.;

949

Johnston, K. P. Graphene oxide nanoplatelet dispersions in concentrated NaCl and

950

stabilization of oil/water emulsions. J. Colloid Interface Sci. 2013, 403, 1–6.

43

ACS Paragon Plus Environment

Environmental Science & Technology

Page 44 of 61

951

(74) Shih, C.; Lin, S.; Sharma, R.; Strano, M. S.; Blankschtein, D. Understanding the

952

pH-dependent behavior of graphene oxide aqueous solutions: A comparative

953

experimental and molecular dynamics simulation study. Langmuir 2012, 28, 235–241.

954

(75) Lanphere, J. D.; Luth, C. J.; Walker, S. L. Effects of solution chemistry on the

955

transport of graphene oxide in saturated porous media. Environ. Sci. Technol. 2013, 47,

956

4255–4261.

957

(76) Hong, B. J.; Compton, O. C.; An, Z.; Eryazici, I.; Nguyen, S. T. Successful

958

stabilization

of

graphene

oxide

in

electrolyte

solutions:

959

biofunctionalization and cellular uptake. ACS Nano 2012, 6, 63–73.

enhancement

of

960

(77) Park, S.; Lee, K. S.; Bozoklu, G.; Cai, W.; Nguyen, S. T.; Ruoff, R. S. Graphene

961

oxide papers modified by divalent ions–enhancing mechanical properties via chemical

962

cross-linking. ACS Nano 2008, 2, 572–578.

963 964

(78) Domingos, R. F.; Tufenkji, N.; Wilkinson, K. J. Aggregation of titanium dioxide nanoparticles: role of a fulvic acid. Environ. Sci. Technol. 2009, 43, 1282–1286.

965

(79) Zhao, J.; Wang, Z.; Dai, Y.; Xing, B. Mitigation of CuO nanoparticle-induced

966

bacterial membrane damage by dissolved organic matter. Water Res. 2013, 47, 4169–

967

4178.

968

(80) Zinchenko, A. A.; Maeda, N.; Pu, S.; Murata, S. Entrapping of fullerenes,

969

nanotubes, and inorganic nanoparticles by a DNA−chitosan complex: A method for

970

nanomaterials removal. Environ. Sci. Technol. 2013, 47, 4489–4496.

971

(81) Hotze, E. M.; Phenrat, T.; Lowry, G. V. Nanoparticle aggregation: Challenges to

972

understanding transport and reactivity in the environment. J. Environ. Qual. 2010, 39,

973

1909–1924.

44

ACS Paragon Plus Environment

Page 45 of 61

Environmental Science & Technology

974

(82) Qiu, L.; Yang, X.; Gou, X.; Yang, W.; Ma, Z. F.; Wallace, G. G.; Li, D.

975

Dispersing carbon nanotubes with graphene oxide in water and synergistic effects

976

between graphene derivatives. Chem. Eur. J. 2010, 16, 10653–10658.

977 978 979 980

(83) Liang, L.; Wu, T.; Kang, Y.; Wang, Q. Dispersion of graphene sheets in aqueous solution by oligodeoxynucleotides. ChemPhysChem 2013, 14, 1626–1632. (84) Smith, R. J.; Lotya, M.; Coleman, J. N. The importance of repulsive potential barriers for the dispersion of graphene using surfactants. New J. Phys. 2010, 12, 125008.

981

(85) Guardia, L.; Fernandez-Merino, M. J.; Paredes, J. I.; Solis-Fernandez, P.; Villar-

982

Rodil, S.; Martinez-Alonso, A.; Tascon, J. M. D. High-throughput production of pristine

983

graphene in an aqueous dispersion assisted by non-ionic surfactants. Carbon 2011, 49,

984

1653–1662.

985

(86) Pupysheva, O. V.; Farajian, A. A.; Knick, C. R.; Zhamu, A.; Jang, B. Z.

986

Modeling direct exfoliation of nanoscale graphene platelets. J. Phys. Chem. C 2010, 114,

987

21083–21087.

988 989 990 991

(87) Lotya, M.; King, P. J.; Khan, U.; De, S.; Coleman, J. N. High-concentration, surfactant-stabilized graphene dispersions. ACS Nano 2010, 4, 3155–3162. (88) Green, A. A.; Hersam, M. C. Solution phase production of graphene with controlled thickness via density differentiation. Nano Lett. 2009, 9, 4031–4036.

992

(89) Backes, C.; Hauke, F.; Hirsch, A. The potential of perylene bisimide derivatives

993

for the solubilization of carbon nanotubes and graphene. Adv. Mater. 2011, 23, 2588–

994

2601.

45

ACS Paragon Plus Environment

Environmental Science & Technology

995

(90) Parviz, D.; Das, S.; Ahmed, H. T.; Irin, F.; Bhattacharia, S.; Green, M. J.

996

Dispersions of non-covalently functionalized graphene with minimal stabilizer. ACS

997

Nano 2012, 6, 8857–8867.

998

(91) Gao, J.; Liu, F.; Liu, Y.; Ma, N.; Wang, Z.; Zhang, X. Environment-friendly

999

method to produce graphene that employs vitamin C and amino acid. Chem. Mater. 2010,

1000

22, 2213–2218.

1001

(92) Song, H.; Hao, L.; Tian, Y.; Wan, X.; Zhang, L.; Lv, Y. Stable and water-

1002

dispersible graphene nanosheets: Sustainable preparation, functionalization, and high-

1003

performance adsorbents for Pb2+. ChemPlusChem 2012, 77, 379–386.

1004 1005

(93) Ciesielski, A.; Samorì, P. Graphene via sonication assisted liquid-phase exfoliation. Chem. Soc. Rev. 2013, 43, 381–398.

1006

(94) Saleh, N. B.; Pfefferle, L. D.; Elimelech, M. Influence of biomacromolecules and

1007

humic acid on the aggregation kinetics of single-walled carbon nanotubes. Environ. Sci.

1008

Technol. 2010, 44, 2412–2418.

1009 1010

(95) Lin, D.; Xing, B. Tannic acid adsorption and its role for stabilizing carbon nanotube suspensions. Environ. Sci. Technol. 2008, 42, 5917–5923.

1011

(96) Krishnamoorthy, K.; Veerapandian, M.; Zhang, L. H.; Yun, K.; Kim, S. J.

1012

Antibacterial efficiency of graphene nanosheets against pathogenic bacteria via lipid

1013

peroxidation. J. Phys. Chem. C 2012, 116, 17280–17287.

1014

(97) Liu, S.; Zeng, T. H.; Hofmann, M.; Burcombe, E.; Wei, J.; Jiang, R.; Kong, J.;

1015

Chen, Y. Antibacterial activity of graphite, graphite oxide, graphene oxide, and reduced

1016

graphene oxide: Membrane and oxidative stress. ACS Nano 2011, 5, 6971–6980.

46

ACS Paragon Plus Environment

Page 46 of 61

Page 47 of 61

Environmental Science & Technology

1017

(98) Gurunathan, S.; Han, J. W.; Dayem, A. A.; Eppakayala, V.; Kim, J. H. Oxidative

1018

stress-mediated antibacterial activity of graphene oxide and reduced graphene oxide in

1019

Pseudomonas aeruginosa. Int. J. Nanomed. 2012, 7, 5901–5914.

1020 1021

(99) Akhavan, O.; Ghaderi, E. Toxicity of graphene and graphene oxide nanowalls against bacteria. ACS Nano 2010, 4, 5731–5736.

1022

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

1023

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

1024

Technol. 2009, 43, 2648–2653.

1025

(101) Zanni, E.; De Bellis, G.; Bracciale, M. P.; Broggi, A.; Santarelli, M. L.; Sarto, M.

1026

S.; Palleschi, C.; Uccelletti, D. Graphite nanoplatelets and Caenorhabditis elegans:

1027

Insights from an in vivo model. Nano Lett. 2012, 12, 2740–2744.

1028

(102) Zhang, W.; Wang, C.; Li, Z.; Lu, Z.; Li, Y.; Yin, J.; Zhou, Y.; Gao, X.; Fang, Y.;

1029

Nie, G.; Zhao, Y. Unraveling stress-induced toxicity properties of graphene oxide and the

1030

underlying mechanism. Adv. Mater. 2012, 24, 5391–5397.

1031

(104) Mesarič, T.; Sepčič, K.; Piazza, V.; Gambardella, C.; Garaventa, F.; Drobne, D.;

1032

Faimali, M. Effects of nano carbon black and single-layer graphene oxide on settlement,

1033

survival and swimming behaviour of Amphibalanus amphitrite larvae. Chem. Ecol. 2013,

1034

29, 643–652.

1035 1036

(104) Gollavelli, G.; Ling, Y. C. Multi-functional graphene as an in vitro and in vivo imaging probe. Biomaterials 2012, 33, 2532–2545.

1037

(105) Chen, L.; Hu, P.; Zhang, L.; Huang, S.; Luo, L.; Huang, C. Toxicity of graphene

1038

oxide and multi-walled carbon nanotubes against human cells and zebrafish. Sci. China:

1039

Chem. 2012, 55, 2209–2216.

47

ACS Paragon Plus Environment

Environmental Science & Technology

1040

(106) Zhao, J.; Wang, Z.; Liu, X.; Xie, X.; Zhang, K.; Xing, B. Distribution of CuO

1041

nanoparticles in juvenile carp (Cyprinus carpio) and their potential toxicity. J. Hazard.

1042

Mater. 2011, 197, 304–310.

1043 1044

(107) Kahru, A.; Dubourguier, H. C. From ecotoxicology to nanoecotoxicology. Toxicology 2011, 269, 105–119.

1045

(108) Pretti, C.; Oliva, M.; Di Pietro, R.; Monni, G.; Cevasco, G.; Chiellini, F.;

1046

Pomelli, C.; Chiappe, C. Ecotoxicity of pristine graphene to marine organisms. Ecotox.

1047

Environ. Safe. 2014, 101, 138–145.

1048 1049

(109) Begum, P.; Ikhtiari, R.; Fugetsu, B. Graphene phytotoxicity in the seedling stage of cabbage, tomato, red spinach, and lettuce. Carbon 2011, 49, 3907–3919.

1050

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

1051

thaliana (Columbia ecotype) T87 cell suspensions. J. Hazard. Mater. 2013, 260, 1032–

1052

1041.

1053

(111) Ma, S.; Lin, D. The biophysicochemical interactions at the interfaces between

1054

nanoparticles and aquatic organisms: adsorption and internalization. Environ. Sci.:

1055

Processes Impacts 2013, 15, 145–160.

1056

(112) Liu, S.; Hu, M.; Zeng, T. H.; Wu, R.; Jiang, R.; Wei, J.; Wang, L.; Kang, J.;

1057

Chen, Y. Lateral dimension-dependent antibacterial activity of graphene oxide sheets.

1058

Langmuir 2012, 28, 12364–12372.

1059

(113) Das, S.; Singh, S.; Singh, V.; Joung, D.; Dowding, J. M.; Reid, D.; Anderson, J.;

1060

Zhai, L.; Khondaker, S. I.; Self, W. T., Seal, S. Oxygenated functional group density on

1061

graphene oxide: Its effect on cell toxicity. Particle Syst. Characterization 2013, 30, 148–

1062

157.

48

ACS Paragon Plus Environment

Page 48 of 61

Page 49 of 61

Environmental Science & Technology

1063

(114) Eda, G.; Fanchini, G.; Chhowalla, M. Large-area ultrathin films of reduced

1064

graphene oxide as a transparent and flexible electronic material. Nat. Nanotechnol. 2008,

1065

3, 270–274.

1066

(115) Kempaiah, R.; Salgado, S.; Chung, W. L.; Maheshwari, V. Graphene as

1067

membrane for encapsulation of yeast cells: protective and electrically conducting. Chem.

1068

Commun. 2011, 47, 11480–11482.

1069

(116) Tu, Y.; Lv, M.; Xiu, P.; Huynh, T.; Zhang, M.; Castelli, M.; Liu, Z.; Huang, Q.;

1070

Fan, C.; Fang, H, Zhou, R. Destructive extraction of phospholipids from Escherichia coli

1071

membranes by graphene nanosheets. Nat. Nanotechnol. 2013, 8, 594–601.

1072

(117) Chang, Y.; Yang, S. T.; Liu, J. H.; Dong, E.; Wang, Y.; Cao, A.; Liu, Y.; Wang,

1073

H. In vitro toxicity evaluation of graphene oxide on A549 cells. Toxicol. Lett. 2011, 200,

1074

201–210.

1075

(118) Wang, A.; Pu, K.; Dong, B.; Liu, Y.; Zhang, L.; Zhang, Z.; Duan, D.; Zhu, Y.

1076

Role of surface charge and oxidative stress in cytotoxicity and genotoxicity of graphene

1077

oxide towards human lung fibroblast cells. J. Appl. Toxicol. 2013, 33, 1156–1164.

1078 1079 1080 1081

(119) Akhavan, O.; Ghaderi, E.; Emamy, H.; Akhavan, F. Genotoxicity of graphene nanoribbons in human mesenchymal stem cells. Carbon 2013, 54, 419–431. (120) Bianco, A. Graphene: Safe or Toxic? The two faces of the medal. Angew. Chem. Int. Ed. 2013, 52, 4986–4997.

1082

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

1083

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

1084

Adv. Healthcare Mater. 2013, 2, 1489–1500.

49

ACS Paragon Plus Environment

Environmental Science & Technology

1085

(122) Wang, Z.; Li, N.; Zhao, J.; White, J. C.; Qu, P.; Xing, B. CuO nanoparticle

1086

interaction with human epithelial cells: cellular uptake, location, export, and genotoxicity.

1087

Chem. Res. Toxicol. 2012, 25, 1512–1521.

1088

(123) Hardman, R. A toxicologic review of quantum dots: toxicity depends on

1089

physicochemical and environmental factors. Environ. Health Perspect. 2006, 114, 165–

1090

172.

1091

(124) Zhang, X.; Sun, H.; Zhang, Z.; Niu, Q.; Chen, Y.; Crittenden, J. C. Enhanced

1092

bioaccumulation of cadmium in carp in the presence of titanium dioxide nanoparticles.

1093

Chemosphere 2007, 67, 160–166.

1094

(125) Ferguson, P. L.; Chandler, G. T.; Templeton, R. C.; DeMarco, A.; Scrivens, W.

1095

A.; Englehart, B. A. Influence of sediment−amendment with single-walled carbon

1096

nanotubes and diesel soot on bioaccumulation of hydrophobic organic contaminants by

1097

benthic invertebrates. Environ. Sci. Technol. 2008, 42, 3879–3885.

1098

(126) Guardia, L.; Villar-Rodil, S.; Paredes, J. I.; Rozada, R.; Martínez-Alonso, A.;

1099

Tascón, J. M. D. UV light exposure of aqueous graphene oxide suspensions to promote

1100

their direct reduction, formation of graphene–metal nanoparticle hybrids and dye

1101

degradation. Carbon 2012, 50, 1014–1024.

1102 1103

(127) Chen, W.; Yan, L.; Bangal, P. R. Chemical reduction of graphene oxide to graphene by sulfur-containing compounds. J. Phys. Chem. C 2010, 114, 19885–19890.

1104

(128) Dimiev, A. M.; Alemany, L. B.; Tour, J. M. Graphene oxide. Origin of acidity,

1105

its instability in water, and a new dynamic structural model. ACS Nano 2013, 7, 576–588.

50

ACS Paragon Plus Environment

Page 50 of 61

Page 51 of 61

Environmental Science & Technology

1106

(129) Kim, S.; Zhou, S.; Hu, Y.; Acik, M.; Chabal, Y. J.; Berger, C.; de Heer, W.;

1107

Bongiorno, A.; Riedo, E. Room-temperature metastability of multilayer graphene oxide

1108

films. Nat. Mater. 2012, 11, 544-549.

1109 1110

(130) Williams, G.; Seger, B.; Kamat, P. V. TiO2-graphene nanocomposites. UVassisted photocatalytic reduction of graphene oxide. ACS Nano 2008, 2, 1487–1491.

1111

(131) Kim, S. R.; Parvez, M. K.; Chhowalla, M. UV-reduction of graphene oxide and

1112

its application as an interfacial layer to reduce the back-transport reactions in dye-

1113

sensitized solar cells. Chem. Phys. Lett. 2009, 483, 124–127.

1114

(132) Fan, Z.; Wang, K.; Wei, T.; Yan, J.; Song, L.; Shao, B. An environmentally

1115

friendly and efficient route for the reduction of graphene oxide by aluminum powder.

1116

Carbon 2010, 48, 1686–1689.

1117

(133) Fan, Z. J.; Kai, W.; Yan, J.; Wei, T.; Zhi, L. J.; Feng, J.; Ren, Y.; Song, L.; Wei,

1118

F. Facile synthesis of graphene nanosheets via Fe reduction of exfoliated graphite oxide.

1119

ACS Nano 2010, 5, 191–198.

1120

(134) Barrows, S. E.; Cramer, C. J.; Truhlar, D. G.; Elovitz, M. S.; Weber, E. J.

1121

Factors controlling regioselectivity in the reduction of polynitroaromatics in aqueous

1122

solution. Environ. Sci. Technol. 1996, 30, 3028–3038.

1123

(135) Wang, Y.; Shi, Z.; Yin, J. Facile synthesis of soluble graphene via a green

1124

reduction of graphene oxide in tea solution and its biocomposites. ACS Appl. Mater.

1125

Interfaces 2011, 3, 1127–1133.

1126

(136) Akhavan, O.; Kalaee, M.; Alavi, Z. S.; Ghiasi, S. M. A.; Esfandiar, A.

1127

Increasing the antioxidant activity of green tea polyphenols in the presence of iron for the

1128

reduction of graphene oxide. Carbon 2012, 50, 3015–3025.

51

ACS Paragon Plus Environment

Environmental Science & Technology

1129

(137) Fernandez-Merino, M. J.; Guardia, L.; Paredes, J. I.; Villar-Rodil, S.; Solis-

1130

Fernandez, P.; Martinez-Alonso, A.; Tascon, J. M. D. Vitamin C is an ideal substitute for

1131

hydrazine in the reduction of graphene oxide suspensions. J. Phys. Chem. C 2010, 114,

1132

6426–6432.

1133

(138) Bose, S.; Kuila, T.; Mishra, A. K.; Kim, N. H.; Lee, J. H. Dual role of glycine as

1134

a chemical functionalizer and a reducing agent in the preparation of graphene: an

1135

environmentally friendly method. J. Mater. Chem. 2012, 22, 9696–9703.

1136 1137 1138 1139

(139) Chen, D.; Li, L.; Guo, L. An environment-friendly preparation of reduced graphene oxide nanosheets via amino acid. Nanotechnology 2011, 22, 325601. (140) Zhu, C.; Guo, S.; Fang, Y.; Dong, S. Reducing sugar: new functional molecules for the green synthesis of graphene nanosheets. ACS Nano 2010, 4, 2429–2437.

1140

(141) Peng, H.; Meng, L.; Niu, L.; Lu, Q. Simultaneous reduction and surface

1141

functionalization of graphene oxide by natural cellulose with the assistance of the ionic

1142

liquid. J. Phys. Chem. C 2012, 116, 16294–16299.

1143

(142) Liu, J.; Fu, S.; Yuan, B.; Li, Y.; Deng, Z. Toward a universal “adhesive

1144

nanosheet” for the assembly of multiple nanoparticles based on a protein-induced

1145

reduction/decoration of graphene oxide. J. Am. Chem. Soc. 2010, 132, 7279–7281.

1146

(143) Yang, F.; Liu, Y.; Gao, L.; Sun, J. pH-Sensitive highly dispersed reduced

1147

graphene oxide solution using lysozyme via an in situ reduction method. J. Phys. Chem.

1148

C 2010, 114, 22085–22091.

1149

(144) Kuila, T.; Bose, S.; Khanra, P.; Mishra, A. K.; Kim, N. H.; Lee, J. H. A green

1150

approach for the reduction of graphene oxide by wild carrot root. Carbon 2012, 50, 914–

1151

921.

52

ACS Paragon Plus Environment

Page 52 of 61

Page 53 of 61

1152 1153 1154 1155

Environmental Science & Technology

(145) Salas, E. C.; Sun, Z.; Lüttge, A.; Tour, J. M. Reduction of graphene oxide via bacterial respiration. ACS Nano 2010, 4, 4852–4856. (146) Wang, G.; Qian, F.; Saltikov, C. W.; Jiao, Y.; Li, Y. Microbial reduction of graphene oxide by Shewanella. Nano Res. 2011, 4, 563–570.

1156

(147) Jiao, Y.; Qian, F.; Li, Y.; Wang, G.; Saltikov, C. W.; Gralnick, J. A.

1157

Deciphering the electron transport pathway for graphene oxide reduction by Shewanella

1158

oneidensis MR-1. J. Bacteriol. 2011, 193, 3662–3665.

1159 1160

(148) Akhavan, O.; and Ghaderi, E. Escherichia coli bacteria reduce graphene oxide to bactericidal graphene in a self-limiting manner. Carbon 2012, 50, 1853–1860.

1161

(149) Khanra, P.; Kuila, T.; Kim, N. H.; Bae, S. H.; Yu, D. S.; Lee, J. H. Simultaneous

1162

bio-functionalization and reduction of graphene oxide by baker's yeast. Chem. Eng. J.

1163

2012, 183, 526–533.

1164

(150) Kang, S. M.; Park, S.; Kim, D.; Park, S. Y.; Ruoff, R. S.; Lee, H. Simultaneous

1165

Reduction and Surface Functionalization of Graphene Oxide by Mussel-Inspired

1166

Chemistry. Adv. Funct. Mater. 2011, 21, 108–112.

1167

(151) Chen, J.; Gu, B.; Royer, R. A.; Burgos, W. D. The roles of natural organic

1168

matter in chemical and microbial reduction of ferric iron. Sci. Total Environ. 2003, 307,

1169

167–178.

1170

(152) Blasco, A. J.; Crevillén, A. G.; González, M. C.; Escarpa, A. Direct

1171

electrochemical sensing and detection of natural antioxidants and antioxidant capacity in

1172

vitro systems. Electroanalysis 2007, 19, 2275–2286.

53

ACS Paragon Plus Environment

Environmental Science & Technology

1173

(153)

Akhavan,

O.;

Abdolahad,

M.;

Esfandiar,

A.;

Page 54 of 61

Mohatashamifar,

M.

1174

Photodegradation of graphene oxide sheets by TiO2 nanoparticles after a photocatalytic

1175

reduction. J. Phys. Chem. C 2010, 114, 12955–12959.

1176

(154) Allen, B. L.; Kichambare, P. D.; Gou, P.; Vlasova, I. I.; Kapralov, A. A.;

1177

Konduru, N.; Kagan, V. E.; Star, A. Biodegradation of single-walled carbon nanotubes

1178

through enzymatic catalysis. Nano Lett. 2008, 8, 3899–3903.

1179

(155) Allen, B. L.; Kotchey, G. P.; Chen, Y.; Yanamala, N. V. K.; Klein-Seetharaman,

1180

J.; Kagan, V. E.; Star, A. Mechanistic investigations of horseradish peroxidase-catalyzed

1181

degradation of single-walled carbon nanotubes. J. Am. Chem. Soc. 2009, 131, 17194–

1182

17205.

1183 1184

(156) Zhao, Y.; Allen, B. L.; Star, A. Enzymatic degradation of multiwalled carbon nanotubes. J. Phys. Chem. A 2011, 115, 9536–9544.

1185

(157) Andón, F. T.; Kapralov, A. A.; Yanamala, N.; Feng, W.; Baygan, A.; Chambers,

1186

B. J.; Hultenby, K.; Ye, F.; Toprak, M. S.; Brandner, D.; Fornara, A.; Klein-Seetharaman,

1187

J.; Kotchey, G. P.; Star, A.; Shvedova, A. A.; Fadeel, B.; Kagan, V. E. Biodegradation of

1188

single-walled carbon nanotubes by eosinophil peroxidase. Small 2013, 9, 2721–2729.

1189

(158) Kagan, V. E.; Konduru, N. V.; Feng, W.; Allen, B. L.; Conroy, J.; Volkov, Y.;

1190

Vlasova, I. I.; Belikova, N. A.; Yanamala, Y.; Kapralov, A.; Tyurina, Y. Y.; Shi, J.; Kisin,

1191

E. R.; Murray, A. R.; Franks, J.; Stolz, D.; Gou, P.; Klein-Seetharaman, J.; Fadeel, B.;

1192

Star, A.; Shvedova, A. A. Carbon nanotubes degraded by neutrophil myeloperoxidase

1193

induce less pulmonary inflammation. Nat. Nanotechnol. 2010, 5, 354–359.

1194

(159) Chrissafis, K.; Antoniadis, G.; Paraskevopoulos, K. M.; Vassiliou, A.; Bikiaris,

1195

D. N. Comparative study of the effect of different nanoparticles on the mechanical

54

ACS Paragon Plus Environment

Page 55 of 61

Environmental Science & Technology

1196

properties and thermal degradation mechanism of in situ prepared poly (ε-caprolactone)

1197

nanocomposites. Compos. Sci. Technol. 2007, 67, 2165–2174.

1198

(160) Zhu, Y.; Murali, S.; Stoller, M. D.; Velamakanni, A.; Piner, R. D.; Ruoff, R. S.

1199

Microwave assisted exfoliation and reduction of graphite oxide for ultracapacitors.

1200

Carbon 2010, 48, 2118–2122.

1201

(161) Fan, X.; Peng, W.; Li, Y.; Li, X.; Wang, S.; Zhang, G.; Zhang, F.

1202

Deoxygenation of exfoliated graphite oxide under alkaline conditions: A green route to

1203

graphene preparation. Adv. Mater. 2008, 20, 4490–4493.

1204

(162) Kuzyakov, Y.; Subbotina, I.; Chen, H.; Bogomolova, I.; Xu, X. Black carbon

1205

decomposition and incorporation into soil microbial biomass estimated by C-14 labeling.

1206

Soil Biol. Biochem. 2009, 41, 210–219.

1207 1208 1209 1210

(163) Masiello, C. A.; Druffel, E. R. M. Black carbon in deep-sea sediments. Science 1998, 280, 1911–1913. (164) Chng, E. L. K.; Pumera, M. The toxicity of graphene oxides: Dependence on the oxidative methods used. Chem. Eur. J. 2013, 19, 8227–8235.

1211

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

1212

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

1213

Interfaces 2011, 3, 2607–2615.

1214 1215

(166) Cheng, C.; Li, D. Solvated graphenes: An emerging class of functional soft materials. Adv. Mater. 2013, 25, 13–30.

1216

55

ACS Paragon Plus Environment

Environmental Science & Technology

1217

Figures and Tables

1218

A

1219

B

1220

C

O

OH OH

O

1221

O

O

1222

Figure 1. Structural models of single-layer graphene (A), graphene oxide (B) and

1223

reduced graphene oxide (C). Functional groups such as hydroxyl, carboxyl and epoxy

1224

groups can be introduced on graphene oxide (B) after the oxidative exfoliation process.

1225

For reduced graphene oxide (C), hydroxyl and carboxyl groups could still remain on the

1226

edge of the graphitic sheets due to incomplete reduction.

1227

56

ACS Paragon Plus Environment

Page 56 of 61

Page 57 of 61

Environmental Science & Technology

1228 1229

2 Adsorption capacity (mmol/g)

inorganic ions on GO organic molecules on rGO 1.5

organic molecules on GO R² = 0.9359

1

0.5 R² = 0.8702 0 0

200

400 600 BET surface area (m2/g)

800

1000

1230 1231

Figure 2. Relationship between BET surface area of GFNs and Langmuir-fitted

1232

adsorption capacities for organic and metal adsorbates: Adsorption of inorganic ions and

1233

organic molecules on GO and rGO, respectively; Adsorption of organic molecules on GO.

1234

The detailed adsorption data and other information are presented in Table S4. For

1235

inorganic ions on rGO, no published data (both surface area of rGO and adsorption

1236

capacity) were found in the current literature.

1237

57

ACS Paragon Plus Environment

Environmental Science & Technology

1238 1239

Figure 3. Dispersion/aggregation behaviors of GO (A) and rGO (B) in aquatic

1240

environments. For GO, the suspension can be stable at the whole pH range of aquatic

1241

environments (pH 5-9) (I). Compression of EDL and the screening effect are responsible

1242

for the aggregation of GO at high ionic strength, and the bridging would occur when

1243

divalent or polyvalent cations are present (II). Macromolecules and colloidal particle (e.g.,

1244

clays) can also influence the stability of GO suspension (III and IV). For rGO, the sheets

1245

could only disperse at high pH and low ionic strength (V). Surfactants (both nonionic and

1246

ionic) and macromolecules, however, can disperse rGO via different mechanisms (VI and

1247

VII).

58

ACS Paragon Plus Environment

Page 58 of 61

Page 59 of 61

Environmental Science & Technology

1248 1249

Figure 4. Proposed interactions of GFNs with aquatic organisms at the cellular level.

1250

Direct penetration and endocytosis are two pathways for GFNs internalization by cells.

1251

The direct penetration is able to cause physical membrane damage via extracting and

1252

cutting effects. Covering on the external surface of cells could lead to indirect toxicity

1253

because of blocking ion/gas exchange. The internalized GFNs can cause oxidative stress,

1254

mitochondrial dysfunction and DNA damage.

1255

59

ACS Paragon Plus Environment

Environmental Science & Technology

A

1256

B

1257

1258 1259

Figure 5. Transformation and degradation of GFNs. (A) Transformation from GO to rGO

1260

with the assistance of (1) UV irradiation, (2) inorganic reductants (e.g., zero-valent

1261

metals, and sulfur-containing compounds), (3) organic compounds (e.g., polyphenols, 60

ACS Paragon Plus Environment

Page 60 of 61

Page 61 of 61

Environmental Science & Technology

1262

vitamins, and proteins) and (4) organisms (e.g., bacteria, fungi and plants). (B)

1263

Photodegradation of GO under UV irradiation, and biodegradation by enzyme (e.g.,

1264

horseradish peroxidase) and macrophage. In panel B, it should be noted that the

1265

photodegradation from rGO to degraded rGO (indicated by the dotted arrow) was only

1266

reported in ethanol, no report in aquatic solution was found.

61

ACS Paragon Plus Environment