Multiwalled Carbon Nanotube Dispersion Methods Affect Their

Apr 29, 2015 - National Research Council, National Academy of Science, 2101 Constitution ... 960 College Station Road, Athens, Georgia 30605, United S...
0 downloads 0 Views 1001KB Size
Subscriber access provided by George Washington University Libraries

Article

Multiwalled Carbon Nanotubes Dispersion Methods Affect Their Aggregation, Deposition, and Biomarker Response Xiaojun Chang, W. Matthew Henderson, and Dermont C. Bouchard Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 29 Apr 2015 Downloaded from http://pubs.acs.org on April 29, 2015

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

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

Page 1 of 24

Environmental Science & Technology

1

Multiwalled Carbon Nanotubes Dispersion Methods

2

Affect Their Aggregation, Deposition, and

3

Biomarker Response

4

Xiaojun Chang,a W. Matthew Henderson,b and Dermont C. Bouchardb,* a

5

6

b

National Research Council Research Associate

USEPA Office of Research and Development, National Exposure Research Laboratory, 960

7

College Station Road, Athens, GA 30605, USA

8

*Corresponding author E-mail: [email protected]; 706-355-8333

9

Total Word Length: 5267 + 6 figures*300 = 7067 words (not including references)

10

1

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 24

11

Abstract

12

To systematically evaluate how dispersion methods affect multi-walled carbon nanotubes

13

(MWNTs) environmental behaviors, MWNTs were dispersed in various solutions [e.g.,

14

surfactants, natural organic matter (NOM), and etc.] via ultrasonication (SON) and long-term

15

stirring (LT). The two tested surfactants [anionic sodium dodecyl sulfate (SDS) and nonionic

16

poly(ethylene glycol)-poly(propylene glycol)-poly(ethylene glycol) (PEO-PPO-PEO) triblock

17

copolymers (Pluronic)] could only disperse MWNTs via ultrasonication; while stable aqueous

18

SON/MWNT and LT/MWNTs suspensions were formed in the presence of the two model

19

NOMs [Suwannee river humic acid (SRHA) and fulvic acid (SRFA)]. Due to the inherent

20

stochastic nature for both methods, the formed MWNTs suspensions were highly heterogeneous.

21

Their physicochemical properties, including surface charge, size, and morphology, greatly

22

depended upon the dispersant type and concentration but were not very sensitive to the

23

preparation methods. Aggregation and deposition behaviors of the dispersed MWNTs were

24

controlled by van der Waal and electrostatic forces, as well as other non-DLVO forces (e.g.,

25

steric, hydrophobic forces, etc.). Unlike the preparation method-independent physicochemical

26

properties, LT/NOM-MWNTs and SON/NOM-MWNTs differed in their fathead minnow

27

epithelial cell metabolomics profiles.

28

2

ACS Paragon Plus Environment

Page 3 of 24

Environmental Science & Technology

29

1. Introduction

30

Since their discovery in 1991,1 carbon nanotubes (CNTs) have attracted attention due to

31

exceptional properties,2,3 but their increasing production and extensive applications 4,5 may result

32

in accidental and intentional releases to the environment.6,7 To investigate transport, fate, and

33

impact of CNTs in the environment, a number of studies on their dispersal in various aqueous

34

matrices have been conducted.8-13

35

Unlike conventional chemicals, the environmental behaviors of nanoparticles are closely

36

related to how they are dispersed in environmental matrices.14 Previous studies have shown that

37

properties of fullerene C60 nanoparticle suspensions (nC60) are greatly affected by preparation

38

methods.15-18 The transport and fate of nanomaterials are determined by their physicochemical

39

properties which, in turn, are affected by dispersion methodology and dispersant species. Due to

40

their strong van der Waals interaction energies of ~ 500 eV/µm of tube-tube contact,19 CNTs

41

cannot be readily dispersed in water as de-bundled individual nanotubes. Other than

42

functionalization via covalent bonding,20-23 in which CNTs are chemically changed, the primary

43

approach for dispersing them in the aqueous phase is through ultrasonication in the presence of

44

stabilizing agents such as surfactants13,24-27 and natural organic matter15,26,28,29. The appearance of

45

reactive oxygen species (ROS),30 surfactant degradation,30 and nanoparticle oxidation31 usually

46

occurs during ultrasonication due to the introduction of high local temperature and pressure.

47

Compared to high energy ultrasonication, the milder process of extended mixing CNTs in water

48

is a better mimic of processes nanoparticles undergo after release to the aquatic environment.

49

Ultrasonication is often chosen over extended mixing in studies on nanoparticle toxicity,

50

transport, and fate, however, since it is faster and relatively more controllable. Implicit in this

51

approach is the assumption that physicochemical properties and toxicities of the ultrasonicated

52

nanoparticle suspension are the same as those produced via extended mixing. However,

53

ultrasonication has been reported to be able to introduce the formation of O-containing

54

functional groups and the disappearance of –CHn groups in the basic CNTs structure.32

55

Studies33,34 have also shown that toxicities of ultrasonicated CNTs were different from those

56

dispersed via stirring. In addition, various stabilities and mobilities in porous media have been

57

observed for CNTs in the presence of different dispersants through classic column studies.35,36

58

To our best knowledge, no study has experimentally verified the aforementioned assumption

3

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 24

59

regarding key stability properties and their biological effects governing MWNTs’ behavior in the

60

environment.

61

The purpose of this study is to systematically evaluate effects of the two dispersion methods

62

(ultrasonication and long-term stirring) and various dispersants [natural water, small molecular

63

weight organic acids, surfactants, and natural organic matter (NOM)] on MWNTs’

64

environmental behavior. MWNTs suspension properties including mass concentration, surface

65

charge, morphology, size, and size distribution were characterized with a range of instruments.

66

Dispersion methods effects on MWNT aggregation and deposition, and on exposure biomarkers

67

(fathead minnow cell cultures), were also evaluated.

68

2. Materials and Methods

69

2.1 Materials. MWNTs with purity of 95 wt% were purchased from CheapTubes Inc.

70

(Brattleboro, VT); reported outside diameters and lengths are 20-30 nm and 10-30 µm,

71

respectively. Analytical grade sodium dodecyl sulfate (SDS) and sodium citrate (Na3Cit) were

72

purchased from Thermo-Fisher (Fremont, CA). Analytical grade sodium acetate (NaAce) was

73

purchased from J. T. Baker (Center Valley, PA). Poly(ethylene glycol)-poly(propylene glycol)-

74

poly(ethylene glycol) (PEO-PPO-PEO) triblock copolymers (poloxamers, or Pluronic®, PF) were

75

purchased from Sigma Aldrich (St. Louis, MO). Suwanee river humic acid (SRHA, Standard II)

76

and Suwannee river fulvic acid (SRFA, Standard I), purchased from the International Humic

77

Substances Society (St. Paul, MN), were chosen as model NOMs. The deionized (DI) water used

78

has a resistivity of 18.2 MΩ·cm and was obtained from an Aqua Solutions Type I Water

79

Purification System. The natural surface water used was collected from a tributary of Calls

80

Creek, a small stream near Athens, GA.26 Cell culture reagents were purchased from American

81

Type Culture Collection (ATCC, Manassas, VA) and the assays for cytotoxicity were obtained

82

from Promega Corporation (Madison, WI). All analytical consumables were purchased from

83

Agilent Technologies (Santa Clara, CA).

84

2.2 Suspension Preparation. MWNTs suspensions were produced via extended mixing (i.e.,

85

long-term magnetic stirring) or ultrasonication. Here we refer to MWNTs suspensions produced

86

by extended mixing as XX mg/L LT/YY-MWNTs, where XX indicates concentration of the

87

dispersant and YY indicates type of dispersant. To distinguish suspensions produced by extended

88

mixing, suspensions produced by ultrasonication are termed XX mg/L SON/YY-MWNTs. 4

ACS Paragon Plus Environment

Page 5 of 24

Environmental Science & Technology

89

In the extended mixing method, purchased MWNTs were mixed with aqueous solutions of

90

different dispersants (DI, Calls Creek water, SDS, PF, SRHA, SRFA, etc.) with an initial

91

[MWNTs] of 100 mg/L. Mixtures were then magnetically stirred at a rate of 400 rpm under

92

ambient conditions at room temperature in the dark; stirring stopped after more than 200 days.

93

To prepare SON/MWNTs suspensions, a 40 mL mixture of MWNTs and dispersant solution

94

(initial [MWNT] = 100 mg/L) was ultrasonicated with a probe sonicator (Sonic & Materials,

95

Newton, CT) in an ice-water bath for 10 min at an average energy level of ~33 W. Mixtures

96

obtained from long-term magnetic stirring and ultrasonication were gravitationally settled for 30

97

days and the stable supernatants were removed and used as stock MWNTs suspensions for

98

further studies. The possible microbial growth in the prepared suspensions were tested by

99

culturing suspensions in the non-selective media, and no microbial growth was observed (Details

100

are presented in the Supporting Information.)

101

2.3 Characterization.

102

2.3.1 General Physicochemical Characterization. MWNTs’ concentration in suspension was

103

determined using a pre-determined calibration curve37 and UV/Vis absorbance (measured by an

104

Enspire Multimode Reader 2300, PerkinElmer, MA) at 500 nm, with background correction for

105

the corresponding dispersant solutions. The intensity averaged hydrodynamic diameter (Dh) and

106

polydispersity index (PDI) of the MWNTs suspensions were measured by dynamic light

107

scattering (DLS) using a Nano ZetaSizer (Malvern Instruments, Worcestershire, UK) with a

108

helium/neon laser (λ = 633 nm). Using the same Nano ZetaSizer, the electrophoretic mobility

109

(EPM) of MWNTs suspensions was determined by phase analysis light-scattering. MWNTs

110

suspensions were drop cast on aluminum oxide surfaces (QSX 303, Biolin Scientific, MD), dried

111

overnight, and their morphologies characterized using a Bruker MultiMode 8 Atomic Force

112

Microscope (AFM) with a Nanoscope V controller and a J-Scanner. Images were taken under the

113

ScanAsyst-Air mode using ScanAsyst-Air probes at the speed of 1.0 Hz and resolution of 512 ×

114

512 pixels.

115

2.3.2 Size Distribution Determination by Asymmetric Flow Field Flow Fractionation (AF4).

116

An AF2000 Focus AF4 (Postnova, Salt Lake City, UT) was utilized to fractionate MWNTs

117

based on hydrodynamic size (diffusion coefficient).38,39 The trapezoidal AF4 channel (27.5 cm

118

from tip to tip) has a tapered inlet and outlet with lengths of 4 and 1 cm, respectively. A 350 µm

5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 24

119

spacer, 10 kDa regenerated cellulose membrane, and 500 µL injection loop accomplished the

120

fractionation. A full AF4 fractionation program includes injection and focusing, elution, and

121

rinsing. Injection is a 5 min period with a tip flow rate of 0.3 mL/min. In the elution phase, the

122

cross flow rate was set as 1.0 mL/min, while the focusing flow [which equals the sum of cross

123

flow and detector flow (0.75 mL/min) less tip flow] was set by AF4 software. In the rinsing

124

phase, tip flow equaled detector flow at 1.0 mL/min, while cross flow was 0.0 mL/min.

125

Absorbance at 254 nm and Dh of the fractionated effluents were sequentially measured by the

126

coupled UV-Vis detector (PN 3241, Postnova) and DLS detector (Nano ZetaSizer, Malvern).

127

Four polystyrene nanosphereTM (PS-NP) size standards (NIST-traceable materials, Thermo-

128

Fisher, Fremont, CA), with certified diameters of 30 ± 1, 59 ± 2, 147 ± 3, and 296 ± 6 nm, were

129

standards for testing the relationship between elution time t and particle hydrodynamic size Dh

130

(Figure S1A) using the above-described fractionation program. As shown in Figure S1, its t

131

increases with its Dh for a homogeneous standard. Using this relationship, a UV/Vis absorbance-

132

weighted size distribution curve for a heterogeneous suspension can be derived from the UV-t

133

profile obtained from the AF4/UV-Vis.

134

2.2.3 Electrokinetic, Aggregation, and Deposition Properties. Effects of ionic strength (IS) on

135

MWNTs surface charges were investigated by measuring MWNTs EPMs in NaCl solutions of

136

varying IS; MWNTs aggregation rates were determined by Time-Resolved DLS. Experimental

137

details can be found in our previous study.40 The deposition of MWNTs on silica dioxide

138

surfaces (QSX 303, Biolin Scientific, MD) was studied using a quartz crystal microbalance with

139

dissipation monitoring (QCM-D, Q-Sense E4, Västra Frölunda, Sweden). Experimental details

140

are provided in Supporting Information.

141

2.2.4 Exposure Biomarkers. All fathead minnow cells (ATCC® CCL-42) were incubated and

142

maintained at 28ºC in T25 culture flasks and transferred to 24-well plates for exposure studies.

143

MWNTs suspensions at a concentration of 300 ng/mL, dispersant-only solutions, or deionized

144

water were added to the passaged cells following a 24-h acclimation period. After a 24-h

145

exposure, cells were quenched and suspended in 80% methanol after being washed twice with

146

phosphate buffered saline (PBS). The cell-methanol suspensions were stored at ≤ -20ºC prior to

147

lyophilization. The metabolites were extracted and analyzed by gas chromatography, coupled

148

with time of flight mass spectrometry (GC/ToF-MS). The acquired data were analyzed with

6

ACS Paragon Plus Environment

Page 7 of 24

Environmental Science & Technology

149

partial least square discriminant analysis (PLS-DA), using commercially available software.

150

Detailed description of cell culture, exposure, instrumental analysis, and data analyses can be

151

found in previous studies.41,42

152

3. Results and Discussion

153

3.1 Physicochemical Properties

154

3.1.1 Dispersant Efficacy. MWNTs were dispersed in deionized water (DI), Call’s Creek

155

water, solutions of organic acids salts with short carbon chains (Na3Cit and NaAce), surfactants

156

(SDS and PF), and natural organic matter (SRHA and SRFA) via long-term magnetic stirring or

157

ultrasonication. For stirred samples, the suspended concentrations of MWNTs in DI, Call’s

158

Creek water, Na3Cit and NaAce solutions (1-200 mg/L) were below the detection limit of UV-

159

Vis measurement (0.15 mg/L) after more than 200 days of stirring. Likewise, ultrasonicated

160

MWNTs samples in the same background solutions rapidly settled out from suspension, and no

161

MWNTs were detected by UV-Vis after the 24-h gravitational settling period. The lack of

162

stability of dispersed MWNTs in these solutions indicates that pure water, natural surface water

163

with low dissolved organic carbon (DOC) concentration,26 and solutions of small molecular

164

weight organic acids are unable to disperse and stabilize MWNTs in the aqueous phase.

165

The two commercial surfactants tested were also ineffective in dispersing MWNTs via long-

166

term stirring since no MWNTs were detected in solutions with 1-20 mg/L of SDS, or in solutions

167

with 1-10 mg/L of the non-ionic surfactant PF. A very low MWNT concentration (< 0.5 mg/L)

168

was observed in 20 mg/L PF solution after 200-day stirring. Stable MWNTs suspensions were

169

obtained in both surfactant solutions with 5-20 mg/L dispersant via ultrasonication. Moreover,

170

[MWNTs] for these SON/MWNTs were independent of surfactant concentrations ([MWNTs] ≈

171

60 mg/L, (Figure 1A)).

172

Natural organic matter usually carries a negative charge due to the carboxylic and phenolic

173

moieties in its structure.43 Several previous studies33,44,45 investigated NOM’s capability of

174

facilitating MWNTs dispersion in aqueous media, however, their time scales were relatively

175

short. In the present study, SRHA and SRFA were very effective in MWNTs dispersion: up to

176

~60 mg/L MWNTs were suspended in solutions with [SRHA] and [SRFA] as low as 1 mg/L

177

(total organic concentration, Figure 1A) by ultrasonication. Stable LT/MWNTs suspensions were

178

also obtained through extended mixing: the final [MWNTs] after 200 days of stirring increased 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 24

179

with dispersant concentration, reaching ~30 mg/L in the presence of 20 mg/L NOM (Figure 1D).

180

NOM’s greater efficacy in stabilizing MWNTs is likely due to the existence of aromatic fractions

181

in their molecular structures which is absent in both aliphatic SDS and PF.45 The ability of NOM

182

to disperse MWNTs at relatively high concentrations, even with low-energy mixing, is

183

environmentally significant since NOM is ubiquitous in the aquatic environment.

184 185 186 187

Figure 1 Concentration, surface charge, and size of MWNTs prepared via ultrasonication (top row) and long-term stirring (bottom row) as a function of dispersant concentration. (Note: The concentration of SRHA and SRFA were determined as total organic carbon concentration.)

188

3.1.2 Electrophoretic Mobility. Dispersants also affect the physicochemical properties of

189

MWNTs suspensions. The EPMs of SON/MWNTs for all four tested solutions (Figure 1B) and

190

LT/MWNTs in two NOM solutions (Figure 1E) decreased with dispersant concentration up to 10

191

mg/L, showing that high [dispersant]/[MWNTs] ratios facilitate formation of stable suspensions.

192

The ultrasonication process likely results in some oxidation of both MWNTs and dispersant,18

193

therefore MWNTs dispersed in the non-ionic surfactant Pluronic were slightly charged.

194

SON/MWNTs dispersed in the anionic surfactant SDS are more negatively charged than the

195

SON/PF-MWNTs. EPM values are similar for SON/MWNTs dispersed in the two tested NOMs

196

and SDS solutions (SON/SDS-MWNTs ≈ SON/SRHA-MWNTs > SON/SRFA-MWNTs, Figure 8

ACS Paragon Plus Environment

Page 9 of 24

Environmental Science & Technology

197

1B). The observation that SRFA-MWNTs are more negatively charged than SRHA-MWNTs in

198

both ultrasonicated and long-term stirred samples is consistent with the fact that SRFA is more

199

negatively charged than SRHA.46

200

3.1.3 Morphology, Size, and Size Distribution.

201

Morphology. Representative AFM images of 20 mg/L LT/SRHA-MWNTs and SON/SRHA-

202

MWNTs are shown in Figure S2. As determined by AFM measurements, the diameters for both

203

LT/SRHA-MWNTs and SON/SRHA-MWNTs are ~30 nm which is consistent with the

204

manufacturer’s reported diameters for the pristine materials. We note that MWNTs in both

205

samples had similar lengths (~100 nm to ~1 µm) which were much shorter than the

206

manufacturer’s reported lengths (10 to 30 µm). Kennedy et al.33 reported that ultrasonicated

207

MWNTs were more fragmented than MWNTs that were magnetically stirred for seven days. As

208

observed here, similarity in lengths of MWNTs dispersed via different methodologies indicates

209

that mild shear forces of long-term magnetic stirring are as effective at de-bundling pristine

210

MWNTs into shorter individual nanotubes as the more intense short-term ultrasonication. .

211

Hydrodynamic Diameter. In all four dispersants, average hydrodynamic diameter (Dh) of

212

SON/MWNTs decreased as dispersant concentration increased from 1 mg/L to 20 mg/L. This

213

decrease corresponds to the increasing [dispersant]/[MWNTs] ratio and the resultant more

214

negative surface charge (Figure 1B). The extent of Dh decrease varied (Figure 1C) with

215

dispersant species: slight decreases in Dh were observed in the presence of SDS (from 215 nm to

216

183 nm) and the two NOMs (from ~210 to ~140 nm), while the greatest decrease (from 1460 nm

217

to 167 nm) was observed in PF. The low [MWNTs], less negative EPM, and significantly large

218

Dh for 1 mg/L SON/PF-MWNTs collectively suggest that this low PF concentration is unable to

219

provide sufficient steric stabilization to prevent the SON/MWNTs from aggregating and settling.

220

Ten and 20 mg/L SON/PF-MWNTs, however, had similar Dh values to those dispersed in other

221

dispersants.

222

The LT/MWNTs at the lowest NOM concentration (1 mg/L) yielded very large (> 500 nm)

223

Dh values (Figure 1E), which correspond to less negative EPM measurements (Figure 1F) and

224

result in low concentrations (Figure 1D). Similarly to a previous study,33 as NOM concentration

225

increased (≥ 5 mg/L), the MWNTs dispersed via long-term stirring developed more negatively-

226

charged surfaces (< -3 × 10-8 m2/V-s) and smaller sizes (~150 nm, size decrease of ~10 nm with 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 24

227

[NOM] from 5 to 20 mg/L), and thus more stable LT/MWNTs suspensions with higher mass

228

concentrations (Figure 1D).

229

Size Distribution. Due to the inherent stochastic nature of the top-down dispersion processes,

230

it is expected that highly polydisperse MWNTs suspensions would form through ultrasonication

231

and extended mixing. The high polydispersity indices (PDIs) around 0.25, shown in Table S2,

232

and broad size distribution (from less than< 30 nm to 1 micron (Figure S3)) confirm the high size

233

heterogeneity in suspensions formed by either technique. Counter-intuitively, size (Table S2) and

234

size distribution (Figures S3A, S3C, S3E, S3G, and S3I) measurements by batch DLS showed

235

that LT/MWNTs have slightly smaller average size and size distribution regions than

236

SON/MWNTs prepared by the much more energetic ultrasonication process. Although previous

237

studies have shown that ultrasonication is more effective in reducing nanoparticle size than

238

magnetic stirring,31 stirring time periods were fairly short (e.g., 300 min31). According to the

239

present study’s observed decrease in size with stirring time which is likely the most important

240

factor determining particle size, long-term stirring is as effective as short-term ultrasonication in

241

reducing MWNT size. Also, Raman spectra for the SON/SRHA-MWNTs and LT/SRHA-

242

MWNTs (Figure S4) show no difference (G/D ratios are 0.87 ± 0.04 and 0.85 ± 0.06 for

243

SON/SRHA-MWNTs and LT/SRHA-MWNTs, respectively), indicating that MWNT oxidation

244

by long-term stirring or 10-min of ultrasonication is comparable.

245

For a highly heterogeneous suspension, the intensity-weighted hydrodynamic diameter is a

246

general index of the whole particle population rather than an accurate measurement of individual

247

particles. The DLS technique is not ideal for measuring particle size for non-spherical particles

248

in suspensions with a range of particle sizes since DLS theory assumes the measured particles

249

are monodisperse spheres.14 And since the light-scattering intensity of a particle scales to the

250

sixth power of its diameter,47 DLS measurement tends to overestimate large particles in a

251

heterogeneous suspension.

252

To mitigate occlusion of smaller particles in batch DLS measurements, asymmetric flow

253

field-flow fractionation (AF4) was employed to fractionate the highly polydisperse MWNTs

254

suspensions. The separation is performed in a thin channel with a laminar flow, which has a

255

parabolic flow profile. An applied cross flow, which is perpendicular to the longitudinal laminar

256

flow, force particles to shift towards the permeable accumulation wall. Particles will re-distribute 10

ACS Paragon Plus Environment

Page 11 of 24

Environmental Science & Technology

257

in the channel due to Brownian motion: smaller particles with larger diffusion coefficients

258

migrate farther away from the accumulation wall,

259

particle retention time is therefore a function of diffusion coefficent and the separation is thus

260

achieved.38,39,48 The UV/Vis-t profiles of the 10 mg/L SRHA-MWNTs suspensions (Figure S5)

261

and their scattered light intensity-t profile (Figure 2B) show that the ultrasonicated sample was

262

eluted from AF4 earlier than the long-term stirred sample. Using the relationship between elution

263

time and particle size established from polystyrene nanosphere standards (Figures 2A and S1),

264

the UV/vis-t profile (Figure S5) was transformed to UV/vis absorbance-weighted size

265

distribution profiles. The normalized UV-size profiles derived (Figure 2C) clearly show the 10

266

mg/L SON/SRHA-MWNTs had a narrower and smaller size distribution than the 10 mg/L

267

LT/SRHA-MWNTs which is contrary to batch DLS measurement results (Figure 2D). The AF4-

268

UV/Vis-DLS fractionation-sizing experiments were conducted for all MWNT suspensions listed

269

in Table S2. Differences between these measurements and batch DLS measurements were

270

observed for all five groups (Figure S3). The discrepancy is likely explained by different

271

mechanisms of AF4-UV/Vis and DLS measurements. Mie theory predicts a linear relationship

272

between nanoparticle volume and extinction coefficient.49 Previous studies found that UV/Vis

273

extinction coefficients of various spherical and non-spherical nanoparticles follow a power law

274

with exponents between 2-3.50-55 This relationship indicates that size distribution curves

275

determined by AF4-UV/Vis-size profiles will still overestimate larger particles, however, they

276

are more sensitive to smaller particles than the batch DLS measurement in which scattered light

277

intensity is proportional to the sixth power of particle size. The suspensions having similar batch

278

DLS-determined average hydrodynamic diameters therefore had substantially different size

279

distributions measured by AF4-UV/Vis. These results suggest that determining particle size and

280

size distribution for a heterogeneous suspension is highly dependent upon the measurement

281

method, and that comparing particle size and heterogeneity of different suspensions must be

282

done carefully.

39

where the transport velocity is higher; the

11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 24

283 284 285 286 287

Figure 2 (A) The chromatographs of polystyrene standards obtained by AF4; (B) Scattered light intensity as a function of elution time; (C) normalized UV/vis absorbance weighted size distribution curves; and (D) scattered light intensity weighted size distribution curves of the AF4 fractionated 10 mg/L SRHA-MWNTs solution.

288

3.2 Environmental Behaviors.

289

3.2.1 EPM Changes with Ionic Strength. The EPM of MWNTs as a function of [NaCl] is

290

shown in Figure 3A. Generally, all EPM values of tested MWNTs became less negative with

291

increasing IS, due to the screening effect of sodium ion (Na+) commonly seen in colloidal

292

suspensions.40,56 The SON/PF-MWNTs were the least negatively charged in the absence of NaCl

293

[(-1.558 ± 0.231) × 10-8 m2/V-s] and their EPMs were very close to zero when [NaCl] ≥ 5 mM;

294

the EPMs of SON/SDS-MWNTs and SON/SRHA-MWNTs gradually reached the range of (-1~-

295

2) × 10-8 m2/V-s as [NaCl] increased from 0 to 40 mM; and the EPMs of SON/SRFA-MWNTs

296

changed the least in the tested IS range. For each dispersant type, there was no substantial

297

difference in EPM changes among suspensions with various concentrations (Figure S6) which

298

suggests that dispersant type is the key factor dictating MWNTs surface charge in aqueous

299

solutions.

300

Changes in EPM of two LT/MWNTs suspensions were also measured. Interestingly, there is

301

no difference in EPM between the LT/SRHA-MWNTs and SON/SRHA-MWNTs, however, the

302

LT/SRHA-MWNTs is more sensitive to IS increase than SON/SRFA-MWNTs. This could be

303

caused by the different structures of SRHA and SRFA or by the transformation of SRHA and

304

SRFA from ultrasonication.

12

ACS Paragon Plus Environment

Page 13 of 24

Environmental Science & Technology

305 13

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 24

306 307 308 309

Figure 3 Environmental behaviors comparison between MWNTs suspension in the presence of SDS, PF, SRHA, and SRFA dispersants (20 mg/L dispersant). (A) Ionic strength effects on surface charge; (B) aggregation attachment efficiency (α αA) as a function of [NaCl]; and (C) QCM-D deposition profile at [NaCl] = 20 mM on silica surface.

310

3.2.2 Aggregation. The aggregation attachment efficiency (αA) was obtained from initial

311

aggregation rates determined by time-resolved DLS.40,57 The αA and critical coagulation

312

concentrations (CCCs) obtained from αA profiles were used to quantitatively describe

313

aggregation in different MWNTs suspensions. No size increase was observed for all SON/PF-

314

MWNTs at all [NaCl] levels (as high as 1 M), showing that MWNTs dispersed in pluronic

315

solutions did not aggregate despite their almost neutral surface charges. Lack of aggregation of

316

the weakly charged SON/PF-MWNTs demonstrates that the non-ionic surfactant PF is an

317

effective dispersion agent that stabilizes MWNTs mainly through steric effects. SON/SDS-

318

MWNTs started aggregating at a low IS level ([NaCl] = 10 mM, Figure S7A) which implies that

319

the steric component of the SDS electrosteric stabilization of MWNTs is not as profound as that

320

of the Pluronic co-block polymer and that electrostatic forces are the major stabilizing forces for

321

SON/SDS-MWNTs. SON/SDS-MWNTs’ αA values increased with IS in the range of 10-100

322

mM, reaching 1.0 (favorable, diffusion-controlled aggregation) as IS increased further. Like IS

323

effects on their surface charges, all SON/SDS-MWNTs exhibited very similar aggregation

324

behaviors, and had the same critical coagulation concentrations (CCC) at ~100 mM NaCl which

325

is higher than the CCC of MWNTs previously dispersed in DI.56

326

SON/SRHA-MWNTs were more stable than SON/SDS-MWNTs (Figure S7B). Although

327

their EPM values both became much less negative (≥ -1× 10-8 m2/V-s) when IS ≥ 20 mM (Figure

328

S6C), all SON/SRHA-MWNTs did not start aggregating until IS ≥ 60 mM which indicates the

329

importance of steric effects in SRHA stabilization of MWNTs suspensions. In the range of 60

330

mM ≤ IS ≤ 120 mM, 20 mg/L SON/SRHA-MWNTs were slightly more stable than those

331

dispersed in solutions with lower SRHA concentrations, however, this difference disappeared

332

with increasing [NaCl] and all CCCs were ~ 300 mM NaCl. Aggregation behaviors of

333

SON/SRFA-MWNTs showed clearer dependence upon dispersant concentration (Figure S7C).

334

MWNTs dispersed in solutions with lower SRFA concentrations started aggregating at low IS

335

levels (~20 mM NaCl), while those dispersed in solutions with higher [SRFA] levels started

336

aggregating when IS exceeded 50 mM. The CCCs of 1 and 5 mg/L SON/SRFA-MWNTs were

337

200 mM, while the CCCs of 10 and 20 mg/L SON/SRFA-MWNTs were 300 mM. 14

ACS Paragon Plus Environment

Page 15 of 24

Environmental Science & Technology

338

αA as a function of [NaCl] for MWNTs dispersed in 20 mg/L dispersant solutions are

339

compared in Figure 3B. SON/SRHA-MWNTs are more stable than SON/SDS-MWNTs,

340

although IS effects on their surface charges are very similar (Figure 3A). Besides electrostatic

341

force, the high stability of SON/SRHA-MWNTs and SON/SRFA-MWNTs is attributed to non-

342

DLVO, steric effects of large NOM molecules which associate with MWNTs through π-π

343

interactions between aromatic rings on MWNTs and aromatic moieties in NOM molecules.44,45

344

The αA of two MWNTs suspensions prepared by long-term stirring are also presented in Figure

345

3B. There is no substantial difference between ultrasonicated samples and long-term stirred

346

samples.

347

3.2.3 QCM-D Deposition. In natural waters, aggregation between MWNTs is unlikely to be

348

the major factor determining their environmental fate due to the relatively low MWNTs

349

concentration relative to naturally occurring suspended particulate matter such as sediments and

350

NOM. Rather, interaction of MWNTs with environmental surfaces is expected to determine

351

MWNTs distribution in the aquatic environment. Here, MWNTs deposition on silica surfaces

352

was investigated using QCM-D. Although the mass of deposited MWNTs cannot be directly

353

calculated by the classic Sauerbrey equation,58,59 the frequency change can be used as a relative

354

index to represent the deposited mass: the decreasing third overtone (∆f3) indicates deposition

355

(i.e., larger decrease in ∆f3 suggests more deposited mass), while the increasing ∆f3 indicates

356

release of deposited mass.

357

For comparison, all tested MWNTs concentrations and ionic strengths were adjusted to 2.4

358

mg/L and 20 mM NaCl, respectively. The MWNTs suspensions deposition and release profiles

359

are presented in Figure 3C. Changes in ∆f3 during the deposition of the four MWNTs

360

suspensions decrease in the order of SON/PF-MWNTs (~20 Hz) > SON/SDS-MWNTs (~2 Hz)

361

> SON/SRHA-MWNTs = SON/SRFA-MWNTs (~0 Hz). In the presence of 20 mM NaCl,

362

SON/PF-MWNTs had almost neutral surface charge [EPM = (-0.160 ± 0.132) × 10-8 m2/V-s].

363

Under this condition, the negative DLVO interaction (i.e., double-layer interaction and Van der

364

Waals interaction) energy (Figure S9A) indicates attractive forces between SON/PF-MWNTs

365

and silica surface, resulting in a large amount of deposited SON/PF-MWNTs. In contrast, there

366

is an energy barrier (Figure S9A) between silica and the more negatively charged SON/SDS-

367

MWNTs [EPM = (-1.563 ± 0.144) × 10-8 m2/V-s] at 20 mM NaCl, resulting in less deposition.

368

Although SON/SRHA-MWNTs surface charge is similar to SON/SDS-MWNTs’ (Figure 3A), 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 24

369

the former does not deposit onto the silica surface which again indicates the importance of steric

370

effects induced by large SRHA molecules. The stronger electrostatic repulsive forces between

371

the more negatively charged SON/SRFA-MWNTs [EPM = (-3.048 ± 0.088) × 10-8 m2/V-s] and

372

silica surfaces resulted in no measurable deposition. Results of aggregation and QCM-D

373

deposition collectively indicate that aggregation results alone are inadequate to predict the

374

stability of nanoparticle suspensions: for example, the most aggregation-stable SON/PF-MWNTs

375

have the greatest potential to deposit onto silica surfaces. According to these results, in a system

376

with 20 mM NaCl and available silica surfaces, the major process governing SON/PF-MWNTs

377

removal from the aqueous phase will be surface deposition rather than settlement due to

378

aggregation.

379

MWNTs release, which indicates the ability of deposited MWNTs to re-enter the water

380

column after being deposited on a surface, is an important parameter for modeling MWNTs fate

381

in the aquatic environment. After deposition, no MWNTs were released when background

382

solution conditions were held constant (i.e., background solution without MWNTs flowed across

383

crystal sensor), indicating that MWNT deposition on silica was irreversible in all of the

384

dispersants. However, the two deposited MWNTs differed greatly in release immediately after

385

DI introduction: for SON/SDS-MWNTs, ∆f3 increased from -2.56 Hz to -2.32 Hz, indicating that

386

~15% of the deposited MWNTs were released; for SON/PF-MWNTs, ∆f3 rapidly increased from

387

-20 Hz to 0 Hz, indicating that all SON/PF-MWNTs deposited on silica surfaces were released.

388

The infinite primary minimum described by DLVO interaction energy profiles under release

389

conditions (Figure S9B), however, is not consistent with the observed release. By considering the

390

Born repulsion which is short-range interactions introduced by interpenetration of electron

391

clouds surrounding the atoms on colloid and planar surfaces,60,61 a finite primary minimum can

392

be obtained. The resultant energy barrier between this primary minimum and energy maximum

393

may be responsible for release of the deposited nanoparticles. In a previous study,8 Yi and Chen

394

attributed release of highly oxidized MWNTs to this barrier. In cases of SON/PF-MWNTs and

395

SON/SDS-MWNTs, the presence of surfactants and their association with MWNTs may

396

introduce other non-DLVO interactions such as hydration effect, steric effect, hydrophobic force,

397

etc. that govern the release process. In addition, effects of salt on the hydration of PEO and PPO

398

blocks of PF62 are also likely to affect SON/PF-MWNTs release.

16

ACS Paragon Plus Environment

Page 17 of 24

Environmental Science & Technology

399

3.2.4 Biomarkers Study. Given the observed effects of NOM on MWNTs properties

400

discussed above, and the limited information available on the exposure potential of carbon-based

401

nanoparticles associated with NOM,63 a metabolomics-based investigation of different

402

dispersants and dispersion methodologies was conducted on fathead minnow (FHM) epithelial

403

cells. Twenty mg/L LT/SRHA-MWNTs, SON/SRHA-MWNTs, LT/SRFA-MWNTs, and

404

SON/SRFA-MWNTs were used to dose the cell cultures; following extraction, derivatization,

405

and GC/ToF-MS analysis, a PLS-DA model was constructed to aid in visualizing class

406

separation between exposed cells and control samples. Despite the fact that no substantial

407

difference in physiochemical properties and stabilities was observed between the ultrasonicated

408

and long-term stirred MWNTs samples, a score plot derived from PLS-DA analysis of the

409

chromatograms (Figure S11) demonstrates the greatest class separation between samples

410

prepared by the two approaches in the presence of SRHA or SRFA. Due to the separation along

411

principal component 1 (x-axis), the largest variation appears to occur between ultrasonicated and

412

long-term stirred samples, independent of organic matter class.

413

To understand the biochemical changes induced by exposure to MWNTs suspensions,

414

metabolites between treatment groups were identified by Student’s t-test filtering (p ≤ 0.05). The

415

chromatogram and t-test comparisons were then performed across each nanotube suspension,

416

subtracting its respective vehicle control (Figure 4). The dispersants caused minor changes in the

417

assay (Figure S12), however, SRHA-MWNTs altered metabolites such as glucose, valine,

418

glycine and cholesterol, regardless of dispersion technique. Decreases in cellular levels of

419

glycine, serine and other amino acids potentially result from decreasing energy requirements for

420

FHM cells following exposure to MWNTs. SON/SRHA-MWNTs caused quantifiable increases

421

in the C16-C18 fatty acids, as well as other membrane-associated metabolites. Edgington et al.64

422

found that organic matter coatings resulted in increased toxicity of carbon nanoparticles, and that

423

these suspensions were more bioavailable to aquatic organisms. There is less abundant variation

424

in the statistical difference between metabolites altered in the LT/SRFA-MWNTs samples,

425

compared to the LT/SRHA-MWNTs. Fulvic acid, being of smaller molecular weight,65 has

426

greater potential to penetrate cell membranes. The SON/SRHA-MWNTs and SON/SRFA-

427

MWNTs appear to induce more biochemical changes with respect to glucose and metabolites

428

involved in cellular energetics than their respective long-term counterparts. Interestingly,

429

exposure to SON/SRFA alters more metabolites than SRFA alone or SON/SRFA-MWNTs. 17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 24

430 431 432 433 434

Figure 4: T-test filtered difference chromatograms of the humic or fulvic acid-dispersed MWNT solutions, compared to their respective controls. Peaks above the x-axis indicate metabolites higher in organic matter exposed cells.

4. Environmental Implications.

435

Results of the present study demonstrate that physicochemical properties, aggregation and

436

deposition, which collectively control MWNTs transport and fate in the environment, mainly

437

depend upon the dispersants and not on dispersion methodology. Interestingly, distinct

438

differences were observed between the metabolomic profiles of FHM cells in ultrasonicated and

439

long-term stirred samples. Although mild long-term magnetic stirring is a better mimic of the

440

MWNTs dispersal process in natural waters than intense ultrasonication, many previous studies

441

on MWNTs environmental transport and subsequent toxicity have chosen ultrasonication over

442

stirring. According to our results, transport-related data of SON/MWNTs can be applied to

443

model released MWNTs transport and fate in the environment, however, LT/MWNTs may be a

444

better choice for conducting MWNTs toxicity studies when NOM is present.

445

Supporting Information Available

18

ACS Paragon Plus Environment

Page 19 of 24

Environmental Science & Technology

446

The fractogram and size vs. retention time for polystyrene nanosphere standards; QCM-D

447

deposition experiment protocol; Comparison of the sizes and polydispersity indices between

448

LT/MWNTs and SON/MWNTs; AFM images of representative SON/MWNTs and LT/MWNTs;

449

Size distributions of LT/MWNTs and SON/MWNTs obtained from batch DLS and AF4-UV/Vis

450

measurements; Raman spectra for SON/SRHA-MWNTs and LT/SRHA-MWNTs; AF4-UV/Vis

451

profiles for SRHA-MWNTs; Ionic strength effects on MWNTs surface charge; Aggregation

452

attachment efficiency as a function of [NaCl]; Replicate deposition profiles of 20 mg/L SON/PF-

453

MWNTs in the presence 20 mM NaCl on silica surface; DLVO profiles for SON/SDS-MWNTs

454

and SON/PF-MWNTs under deposition and release condition; QCM-D deposition profiles for

455

NOM dispersed MWNTs; PLS-DA score plot for NOM dispersed MWNTs; T-test filtered

456

difference chromatograms of the SRHA and sonicated SRHA solution exposed cells. This

457

information is available free of charge via the Internet at http://pubs.acs.org.

458

Disclaimer

459

This paper has been reviewed in accordance with the USEPA’s peer and administrative

460

review policies and approved for publication. Mention of trade names or commercial products

461

does not constitute endorsement or recommendation for use.

462

Reference

463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481

(1)Iijima, S., Helical microtubules of graphitic carbon. Nature 1991, 354, 56-58. (2)Hu, Y. H.; Shenderova, O. A.; Brenner, D. W., Carbon nanostructures: Morphologies and properties. J. Comput. Theor. Nanosci. 2007, 4, 199-221. (3)Loiseau, A., Understanding carbon nanotubes: from basics to applications. Springer: 2006; Vol. 677. (4)Hendren, C. O.; Mesnard, X.; Dröge, J.; Wiesner, M. R., Estimating production data for five engineered nanomaterials as a basis for exposure assessment. Environ. Sci. Technol. 2011, 45, 2562-2569. (5)Park, S. I.; Vosguerichian, M.; Bao, Z., A review of fabrication and applications of carbon nanotube film-based flexible electronics. Nanoscale 2013, 5, 1727-1752. (6)Nowack, B.; Bucheli, T. D., Occurrence, behavior and effects of nanoparticles in the environment. Environ. Pollut. 2007, 150, 5-22. (7)Petersen, E. J.; Zhang, L.; Mattison, N. T.; O'Carroll, D. M.; Whelton, A. J.; Uddin, N.; Tinh, N.; Huang, Q.; Henry, T. B.; Holbrook, R. D.; Chen, K. L., Potential release pathways, environmental fate, And ecological risks of carbon nanotubes. Environ. Sci. Technol. 2011, 45, 9837-9856. (8)Yi, P.; Chen, K. L., Release kinetics of multiwalled carbon nanotubes deposited on silica surfaces: Quartz crystal microbalance with dissipation (QCM-D) measurements and modeling. Environ. Sci. Technol. 2014, 48, 4406-4413. 19

ACS Paragon Plus Environment

Environmental Science & Technology

482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526

Page 20 of 24

(9)Yi, P.; Chen, K. L., Interaction of multiwalled carbon nanotubes with supported lipid bilayers and vesicles as model biological membranes. Environ. Sci. Technol. 2013, 47, 5711-5719. (10)Yi, P.; Chen, K. L., Influence of solution chemistry on the release of wultiwalled carbon nanotubes from silica surfaces. Environ. Sci. Technol. 2013, 12211-12218. (11)Lu, Y.; Xu, X.; Yang, K.; Lin, D., The effects of surfactants and solution chemistry on the transport of multiwalled carbon nanotubes in quartz sand-packed columns. Environ. Pollut. 2013, 182, 269-277. (12)Hwang, Y. S.; Qu, X. L.; Li, Q., The role of photochemical transformations in the aggregation and deposition of carboxylated multiwall carbon nanotubes suspended in water. Carbon 2013, 55, 81-89. (13)Herrero-Latorre, C.; Alvarez-Mendez, J.; Barciela-Garcia, J.; Garcia-Martin, S.; PenaCrecente, R. M., Characterization of carbon nanotubes and analytical methods for their determination in environmental and biological samples: A review. Anal. Chim. Acta 2015, 853, 77-94. (14)Petersen, E. J.; Henry, T. B., Methodological considerations for testing the ecotoxicity of carbon nanotubes and fullerenes: review. Environ. Toxicol. Chem. 2012, 31, 60-72. (15)Duncan, L. K.; Jinschek, J. R.; Vikesland, P. J., C60 colloid formation in aqueous systems: Effects of preparation method on size, structure, and surface, charge. Environ. Sci. Technol. 2008, 42, 173-178. (16)Chang, X.; Duncan, L. K.; Jinschek, J. R.; Vikesland, P. J., Alteration of nC60 in the presence of environmentally relevant carboxylates. Langmuir 2012, 28, 7622-7630. (17)Brant, J.; Lecoanet, H.; Hotze, M.; Wiesner, M., Comparison of electrokinetic properties of colloidal fullerenes (nC60) formed using two procedures. Environ. Sci. Technol. 2005, 39, 63436351. (18)Brant, J.; Lecoanet, H.; Wiesner, M. R., Aggregation and deposition characteristics of fullerene nanoparticles in aqueous systems. J. Nanopart. Res. 2005, 7, 545-553. (19)Girifalco, L. A.; Hodak, M.; Lee, R. S., Carbon nanotubes, buckyballs, ropes, and a universal graphitic potential. Phys. Rev. B 2000, 62, 13104-13110. (20)Sun, Y. P.; Fu, K.; Lin, Y.; Huang, W., Functionalized carbon nanotubes:  Properties and applications. Acc. Chem. Res. 2002, 35, 1096-1104. (21)Banerjee, S.; Hemraj-Benny, T.; Wong, S. S., Covalent surface chemistry of single-walled carbon nanotubes. Adv. Mater. 2005, 17, 17-29. (22)Tasis, D.; Tagmatarchis, N.; Georgakilas, V.; Prato, M., Soluble Carbon Nanotubes. Chem. Eur. J. 2003, 9, 4000-4008. (23)Yi, P.; Chen, K. L., Influence of surface oxidation on the aggregation and deposition kinetics of multiwalled carbon nanotubes in monovalent and divalent electrolytes. Langmuir 2011, 27, 3588-3599. (24)Rastogi, R.; Kaushal, R.; Tripathi, S. K.; Sharma, A. L.; Kaur, I.; Bharadwaj, L. M., Comparative study of carbon nanotube dispersion using surfactants. J. Colloid Interface Sci. 2008, 328, 421-428. (25)Tian, Y. A.; Gao, B.; Ziegler, K. J., High mobility of SDBS-dispersed single-walled carbon nanotubes in saturated and unsaturated porous media. J. Hazard. Mater. 2011, 186, 1766-1772. (26)Bouchard, D.; Zhang, W.; Powell, T.; Rattanaudompol, U., Aggregation kinetics and transport of single-walled carbon nanotubes at low surfactant concentrations. Environ. Sci. Technol. 2012, 46, 4458-4465.

20

ACS Paragon Plus Environment

Page 21 of 24

527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572

Environmental Science & Technology

(27)Dassios, K. G.; Alafogianni, P.; Antiohos, S. K.; Leptokaridis, C.; Barkoula, N.-M.; Matikas, T. E., Optimization of sonication parameters for homogeneous surfactant-Assisted dispersion of multiwalled carbon nanotubes in aqueous solutions. J. Phys. Chem. C 2015, ASAP. (28)Brant, J. A.; Labille, J.; Bottero, J. Y.; Wiesner, M. R., Characterizing the impact of preparation method on fullerene cluster structure and chemistry. Langmuir 2006, 22, 3878-3885. (29)Zhou, X.; Shu, L.; Zhao, H.; Guo, X.; Wang, X.; Tao, S.; Xing, B. S., Suspending multiwalled carbon nanotubes by humic acids from a peat soil. Environ. Sci. Technol. 2012, 46, 38913897. (30)Sesis, A.; Hodnett, M.; Memoli, G.; Wain, A. J.; Jurewicz, I.; Dalton, A. B.; Carey, J. D.; Hinds, G., Influence of acoustic cavitation on the controlled ultrasonic dispersion of carbon nanotubes. J. Phys. Chem. B 2013, 117, 15141-15150. (31)Mejia, J.; Valembois, V.; Piret, J.-P.; Tichelaar, F.; van Huis, M.; Masereel, B.; Toussaint, O.; Delhalle, J.; Mekhalif, Z.; Lucas, S., Are stirring and sonication pre-dispersion methods equivalent for in vitro toxicology evaluation of SiC and TiC? J. Nanopart. Res. 2012, 14, 1-18. (32)Yang, D.; Rochette, J.; Sacher, E., Functionalization of multiwalled carbon nanotubes by mild aqueous sonication. J. Phys. Chem. B 2005, 109, 7788-7794. (33)Kennedy, A. J.; Gunter, J. C.; Chappell, M. A.; Goss, J. D.; Hull, M. S.; Kirgan, R. A.; Steevens, J. A., Influence of nanotube preparation in aquatic bioassays. Environ. Toxicol. Chem. 2009, 28, 1930-1938. (34)Kennedy, A. J.; Hull, M. S.; Steevens, J. A.; Dontsova, K. M.; Chappell, M. A.; Gunter, J. C.; Weiss, C. A., Factors influencing the partitioning and toxicity of nanotubes in the aquatic environment. Environ. Toxicol. Chem. 2008, 27, 1932-1941. (35)Tian, Y.; Gao, B.; Morales, V. L.; Wang, Y.; Wu, L., Effect of surface modification on single-walled carbon nanotube retention and transport in saturated and unsaturated porous media. J. Hazard. Mater. 2012, 239, 333-339. (36)Lu, Y.; Yang, K.; Lin, D., Transport of surfactant-facilitated multiwalled carbon nanotube suspensions in columns packed with sized soil particles. Environ. Pollut. 2014, 192, 36-43. (37)Chang, X.; Vikesland, P. J., UV-vis spectroscopic properties of nC60 produced via extended mixing. Environ. Sci. Technol. 2011, 45, 9967-9974. (38)Wahlund, K. G.; Giddings, J. C., Properties of an asymmetrical flow field-flow fractionation channel having one permeable wall. Anal. Chem. 1987, 59, 1332-1339. (39)Baalousha, M.; Stolpe, B.; Lead, J. R., Flow field-flow fractionation for the analysis and characterization of natural colloids and manufactured nanoparticles in environmental systems: A critical review. J. Chromatogr. A 2011, 1218, 4078-4103. (40)Chang, X.; Bouchard, D. C., Multiwalled carbon nanotube deposition on model environmental surfaces. Environ. Sci. Technol. 2013, 47, 10372-10380. (41)Teng, Q.; Huang, W.; Collette, T. W.; Ekman, D. R.; Tan, C., A direct cell quenching method for cell-culture based metabolomics. Metabolomics 2009, 5, 199-208. (42)West, F. D.; Henderson, W. M.; Yu, P.; Yang, J.-Y.; Stice, S. L.; Smith, M. A., Metabolomic response of human embryonic stem cell derived germ-like cells after exposure to steroid hormones. Toxicol. Sci. 2012, 9-20. (43)Summers, R. S.; Roberts, P. V., Activated carbon adsorption of humic substances: I. Heterodisperse mixtures and desorption. J. Colloid Interface Sci. 1988, 122, 367-381. (44)Hyung, H.; Kim, J. H., Natural organic matter (NOM) adsorption to multi-walled carbon nanotubes: Effect of NOM characteristics and water quality parameters. Environ. Sci. Technol. 2008, 42, 4416-4421. 21

ACS Paragon Plus Environment

Environmental Science & Technology

573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617

Page 22 of 24

(45)Hyung, H.; Fortner, J. D.; Hughes, J. B.; Kim, J. H., Natural organic matter stabilizes carbon nanotubes in the aqueous phase. Environ. Sci. Technol. 2007, 41, 179-184. (46)Sutton, R.; Sposito, G., Molecular structure in soil humic substances: the new view. Environ. Sci. Technol. 2005, 39, 9009-9015. (47)Hiemenz, P. C.; Rajagopalan, R., Principles of Colloid and Surface Chemistry. Marcel Dekker, Inc.: New York, 1997. (48)Isaacson, C. W.; Bouchard, D., Asymmetric flow field flow fractionation of aqueous C60 nanoparticles with size determination by dynamic light scattering and quantification by liquid chromatography atmospheric pressure photo-ionization mass spectrometry. J. Chromatogr. A 2010, 1217, 1506-1512. (49)Wang, Y.; Herron, N., Nanometer-sized semiconductor clusters - Materials Synthesis, quantum size effects, and photophysical properties. J. Phys. Chem. 1991, 95, 525-532. (50)Yu, W. W.; Qu, L.; Guo, W.; Peng, X., Experimental determination of the extinction coefficient of CdTe, CdSe, and CdS nanocrystals. Chem. Mat. 2003, 15, 2854-2860. (51)Jain, P. K.; Lee, K. S.; El-Sayed, I. H.; El-Sayed, M. A., Calculated absorption and scattering properties of gold Nanoparticles of different size, shape, and composition: Applications in biological imaging and biomedicine. J. Phys. Chem. B 2006, 110, 7238-7248. (52)Striolo, A.; Ward, J.; Prausnitz, J. M.; Parak, W. J.; Zanchet, D.; Gerion, D.; Milliron, D.; Alivisatos, A. P., Molecular weight, osmotic second virial coefficient, and extinction coefficient of colloidal CdSe nanocrystals. J. Phys. Chem. B 2002, 106, 5500-5505. (53)Liu, X.; Atwater, M.; Wang, J.; Huo, Q., Extinction coefficient of gold nanoparticles with different sizes and different capping ligands. Colloid Surf. B-Biointerfaces 2007, 58, 3-7. (54)Dai, Q.; Wang, Y.; Li, X.; Zhang, Y.; Pellegrino, D. J.; Zhao, M.; Zou, B.; Seo, J. T.; Wang, Y.; Yu, W. W., Size-dependent composition and molar extinction coefficient of PbSe semiconductor nanocrystals. Acs Nano 2009, 3, 1518-1524. (55)Chang, X.; Vikesland, P. J., Uncontrolled variability in the extinction spectra of C60 nanoparticle suspensions. Langmuir 2013, 9685-9693. (56)Saleh, N. B.; Pfefferle, L. D.; Elimelech, M., Aggregation kinetics of multiwalled carbon nanotubes in aquatic systems: Measurements and environmental implications. Environ. Sci. Technol. 2008, 42, 7963-7969. (57)Chen, K. L.; Elimelech, M., Influence of humic acid on the aggregation kinetics of fullerene (C60) nanoparticles in monovalent and divalent electrolyte solutions. J Colloid. Interf. Sci. 2007, 309, 126-134. (58)Sauerbrey, G., Verwendung von Schwingquarzen zur Wagung dunner schichten und zur mikrowagung. Zeitschrift Fur Physik 1959, 155, 206-222. (59)Reviakine, I.; Johannsmann, D.; Richter, R. P., Hearing what you cannot see and visualizing what you hear: interpreting quartz crystal microbalance data from solvated interfaces. Anal. Chem. 2011, 83, 8838-8848. (60)Elimelech, M.; Gregory, J.; Jia, X.; Willams, R. A., Particle Deposition and Aggregatuion Measurement, Modelling and Simulation. Butterworth-Heinemann: Oxford, England, 1995. (61)Ruckenstein, E.; Prieve, D. C., Adsorption and desorption of particles and their chromatographic separation. AICHE J. 1976, 22, 276-283. (62)Pandit, N.; Trygstad, T.; Croy, S.; Bohorquez, M.; Koch, C., Effect of salts on the micellization, clouding, and solubilization behavior of Pluronic F127 solutions. J. Colloid Interface Sci. 2000, 222, 213-220.

22

ACS Paragon Plus Environment

Page 23 of 24

618 619 620 621 622 623 624 625 626

Environmental Science & Technology

(63)Kim, K.-T.; Jang, M.-H.; Kim, J.-Y.; Xing, B. S.; Tanguay, R. L.; Lee, B. G.; Kim, S. D., Embryonic toxicity changes of organic nanomaterials in the presence of natural organic matter. Sci. Total Environ. 2012, 426, 423-429. (64)Edgington, A. J.; Roberts, A. P.; Taylor, L. M.; Alloy, M. M.; Reppert, J.; Rao, A. M.; Mao, J.; Klaine, S. J., The influence of natural organic matter on the toxicity of multiwalled carbon nanotubes. Environ. Toxicol. Chem. 2010, 29, 2511-2518. (65)Baalousha, M.; Motelica-Heino, M. l.; Coustumer, P. L., Conformation and size of humic substances: Effects of major cation concentration and type, pH, salinity, and residence time. Colloid Surf. A-Physicochem. Eng. Asp. 2006, 272, 48-55.

627

23

ACS Paragon Plus Environment

Environmental Science & Technology

49x46mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 24 of 24