Aggregation Behavior of Inorganic 2D ... - ACS Publications

Mar 15, 2019 - van der Waals attraction has been found to govern that of h-BN. .... affect the colloidal stability of inorganic 2D NMs differently fro...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Environmental Processes

Aggregation Behavior of Inorganic 2D Nanomaterials Beyond Graphene: Insights from Molecular Simulations and Modified DLVO Theory Tashfia M. Mohona, Anusha Gupta, Arvid Masud, Szu-Chia Chien, Li-Chiang Lin, Prathima C. Nalam, and Nirupam Aich Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b05180 • Publication Date (Web): 18 Mar 2019 Downloaded from http://pubs.acs.org on March 21, 2019

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

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

Page 1 of 43

Environmental Science & Technology

1 2 3 4

Aggregation Behavior of Inorganic 2D Nanomaterials Beyond Graphene: Insights from Molecular Modeling and Modified DLVO Theory by

5 6 7

Tashfia M. Mohona,1 Anusha Gupta,2 Arvid Masud,1 Szu-Chia Chien,3 Li-Chiang Lin,4 Prathima Nalam,5 and Nirupam Aich1,*

8 9 10

1Department

of Civil, Structural and Environmental Engineering, University at Buffalo, The State University of New York, Buffalo, NY, USA

11 12

2Department

13 14

3Department

15 16

4William

17 18

5Department

of Civil Engineering, Indian Institute of Technology, Gandhinagar, Gujarat, India of Materials Science and Engineering, The Ohio State University, Columbus, OH, USA G. Lowrie Department of Chemical and Biomolecular Engineering, The Ohio State University, Columbus, OH, USA of Materials Design and Innovation, University at Buffalo, The State University of New York, Buffalo, NY, USA

19 20 21 22 23

Submitted to Environmental Science & Technology (March 15, 2019)

24 25 26 27 28 29 30 31

*Corresponding

Author: Nirupam Aich, Phone: 716-645-0977, Email: [email protected]

ACS Paragon Plus Environment

Environmental Science & Technology

32

Abstract:

33

We report the comparative aggregation behavior of three emerging inorganic 2D nanomaterials

34

(NMs): MoS2, WS2, and h-BN in aquatic media. Their aqueous dispersions were subjected to

35

aggregation under varying concentrations of monovalent (NaCl) and divalent (CaCl2)

36

electrolytes. Moreover, Suwanee River Natural Organic Matter (SRNOM) has been used to

37

analyze the effect of natural macromolecules on 2D NM aggregation. An increase in electrolyte

38

concentration resulted in electrical double-layer compression of the negatively charged 2D NMs,

39

thus displaying classical Derjaguin-Landau-Verwey-Overbeek (DLVO) type interaction. The

40

critical coagulation concentrations (CCC) have been estimated as 37, 60, and 19 mM NaCl and

41

3, 7.2, and 1.3 mM CaCl2 for MoS2, WS2, and h-BN, respectively. Theoretical predictions of

42

CCC by modified DLVO theory have been found comparable to the experimental values when

43

dimensionality of the materials is taken into account and a molecular modeling approach was

44

used for calculating molecular level interaction energies between individual 2D NM nanosheets.

45

Electrostatic repulsion has been found to govern colloidal stability of MoS2 and WS2 while the

46

van der Waals attraction has been found to govern that of h-BN. SRNOM stabilizes the 2D NMs

47

significantly possibly by electrosteric repulsion. The presence of SRNOM completely stabilized

48

MoS2 and WS2 at both low and high ionic strengths. While h-BN still showed appreciable

49

aggregation in the presence of SRNOM, the aggregation rates were decreased by 2.6- and 3.7-

50

fold at low and high ionic strengths, respectively. Overall, h-BN nanosheets will have higher

51

aggregation potential and thus limited mobility in the natural aquatic environment when

52

compared to MoS2 and WS2. These results can also be used to mechanistically explain fate,

53

transport, transformation, organismal uptake, and toxicity of inorganic 2D NMs in the natural

54

ecosystems.

ACS Paragon Plus Environment

Page 2 of 43

Page 3 of 43

Environmental Science & Technology

55

Keywords: Aggregation, 2D Nanomaterials, DLVO, MoS2, WS2, h-BN.

56

1. Introduction

57

The discovery of the unprecedented properties of planar atomically thin carbon allotrope

58

graphene has invoked a keen interest in other two dimensional nanomaterials (2D NMs).1

59

Molybdenum disulfide (MoS2), tungsten disulfide (WS2), and hexagonal boron nitride (h-BN), in

60

particular, have received much attention as inorganic analogues of graphene, owing to their

61

unique properties such as atomic thinness, tunable band-gap, high electron mobility, and thermal

62

conductivity.2, 3 Bulk MoS2, WS2, and h-BN have been historically used as solid lubricants,4, 5

63

while their mono-layered nanoscale counterparts are promising for optoelectronic,6,

64

biomedical,8, 9 energy,10-12 and environmental1, 13, 14 applications. MoS2, WS2, and h-BN are non-

65

carbonaceous and physicochemically distinct from graphene or its derivatives.15 MoS2 and WS2

66

are transition metal dichalcogenides (TMDs) having a triatomic layer structure: one layer of

67

transition metal atoms sandwiched between two layers of sulfur.3 h-BN, on the contrary, is a

68

structural analogue of graphene with honeycomb lattice but having ionic B-N bond as opposed to

69

covalent C-C bond of graphene.13 Unlike superconductive graphene with zero band gap, single

70

layer MoS2, WS2, and h-BN are semiconductors and have direct bandgaps7,

71

quantum confinement.17 This direct band gap imparts features such as silicon like electron

72

mobility,18 excellent current on/off ratio,18 and strong photoluminescence19 enabling use for field

73

effect transistors, photodetectors, photovoltaics, and energy storage and conversion. Moreover,

74

superior optical absorption, photoluminescence, high lubricity, flexibility, and catalytic activity

75

make them suitable for drug delivery, orthodontics, endoscopy, optogenetics, bioimaging, and

76

biosensing.20

77

photocatalysis,21 disinfection,22 and high performance water purification membrane.1 Such rise

16

7

arising due to

Their use has recently been extrapolated for contaminant adsorption,13

ACS Paragon Plus Environment

Environmental Science & Technology

78

in research and potential commercialization point towards inevitable release and exposure of

79

these inorganic 2D NMs to the environment.

80

The increasing manufacture and use of 2D NMs can lead to their release, like any NMs, to

81

different environmental matrices i.e., air, soil, and water.23 Release into ambient or indoor air

82

may occur during their production and handling, and pose occupational risk to workers.24

83

Landfills are expected to receive a considerable amount of NM-laden products, and thus,

84

leaching of the NMs into surrounding soil and water is highly likely.25 Additionally, NMs not

85

captured by wastewater treatment processes are released into the water bodies with effluent.26 It

86

thus becomes imperative to study the environmental fate and toxicity of these emerging

87

inorganic 2D NMs.

88

Colloidal stability of 2D NMs in aquatic environment influences their interactions with

89

microorganisms.27 Bioavailability, and thus toxicity, will be governed by their mobility and

90

aggregation behavior in the corresponding environment.28 Among 2D NMs, graphene has been

91

most widely studied and its toxicity has been most thoroughly scrutinized.29, 30 To date, only few

92

studies have investigated the interactions of inorganic 2D NMs with microorganisms.31 Owing to

93

its higher electron conductivity than bulk MoS2, exfoliated MoS2 is hypothesized to generate

94

more reactive oxygen species (ROS) that lead to enhanced cytotoxicity.31 Membrane stress

95

induced by atomically thin MoS2 nanosheets and their increased surface area have also been

96

accounted for their cytotoxicity.32

97

antibacterial activity with E. coli,22 planktonic cells,33 and biofilms33 compared to their bulk

98

counterparts, suggesting that the degree of exfoliation influences their cytotoxicity. Similarly,

99

WS2 nanosheets demonstrated bactericidal activity toward E. coli and S. aureus mediated by

100

oxidative and membrane stress.34 The enhanced toxicity of h-BN nanosheets compared to its

MoS2 nanosheets have been found to show higher

ACS Paragon Plus Environment

Page 4 of 43

Page 5 of 43

Environmental Science & Technology

101

bulk is attributed to ROS generation by unsaturated boron atoms, acting as radicals, at the

102

nanosheet edges.35 Aggregation of 2D NMs reverses the exfoliation mediated features, including

103

thinness, surface area, and exposed edges, and thereby affects toxicity and its mechanisms.

104

Knowledge of the aggregation behavior of 2D NMs is thus important for their environmental

105

implication studies.

106

The classical Derjaguin-Landau-Verwey-Overbeek (DLVO) theory is able to explain the

107

aggregation behavior of charged colloidal particles.36 The assumption of spherical particles has

108

also been made for sheet-like planar 2D NMs like graphene oxide (GO).37-39 However, the van

109

der Waal (vdW) interaction energy scales inversely to the sixth and fourth powers of distance for

110

spherical particles and 2D NMs respectively.40

111

physicochemical properties of 2D NMs warrant the use of appropriate force laws for properly

112

capturing their aggregation behavior. A recent study on GO has demonstrated agreement of the

113

2D NM’s colloidal stability with DLVO theory;41 however, only when Hamaker constants and

114

vdW forces for GO have been modified to account for the effect of its dimensionality. GO

115

contains carboxyl and hydroxyl functional groups which provide them both hydrophilicity and

116

hig negative surface charge. Hence, electrostatic repulsion is hypothesized to be the governing

117

DLVO force behind GO’s excellent colloidal stability.42

Such difference coupled with the distinct

118

On the contrary, inorganic 2D NMs i.e., MoS2, WS2, and h-BN are considered hydrophobic.

119

However, MoS2 and WS2,have lone pair of electrons on the chalcogen atoms rendering them

120

active for a wide range of chemical reactions.43 Although they are considered hydrophobic, their

121

exfoliation in aqueous media through ultrasonic (i.e., by ultrasound mediated shear-forces)44 or

122

chemical (i.e., intercalation by lithium ions)45 routes impart high negative charges on MoS2 and

123

WS2 nanosheets.46, 47 In case of h-BN, the interplanar bonding is a composite of vdW forces and

ACS Paragon Plus Environment

Environmental Science & Technology

124

ionic attraction, the latter arising due to charge localization on N atom.48 The electron-deficient

125

boron atoms act as Lewis acids, thereby enriching the h-BN chemistry.49 Thus, despite being

126

considered hydrophobic in bulk form, h-BN is prone to both solvent polarity effect and

127

hydrolysis.50 Sonication-assisted aqueous exfoliation of h-BN causes large sheets to break off

128

along defect sites into smaller flakes. It is hypothesized that the boron edges of these flakes have

129

dangling hydroxyl groups resulting from thermodynamically favorable release of ammonia.50

130

Thus, exfoliated h-BN nanosheets become hydrophilic and colloidally stable in aqueous

131

dispersions. It is anticipated that the interplay of these distinct physicochemical properties may

132

affect the colloidal stability of inorganic 2D NMs differently from each other as well as from

133

GO. The critical knowledge gap exists in understanding the applicability and validity of the

134

modified DLVO theory for inorganic 2D NMs beyond GO.

135

The objective of this paper is to assess and compare the aggregation behavior of these three

136

emerging and important inorganic 2D NMs i.e., MoS2, WS2, and h-BN in aqueous media. An

137

ethanol-water mixture is used to exfoliate and disperse these 2D NMs. The chemical identity,

138

physical morphology, and electrokinetic properties of the 2D NMs are then characterized by UV-

139

vis spectroscopy, high resolution transmission electron microscopy (HRTEM), scanning

140

transmission electron microscopy (STEM), atomic force microscopy (AFM), and electrophoretic

141

measurements.

142

dynamic light scattering (TRDLS) under a broad spectrum of mono- and divalent electrolytic

143

conditions and also in the presence of Suwanee River Natural Organic Matter (SRNOM). The

144

mechanisms underlying the aggregation behavior of the inorganic 2D NMs are determined using

145

modified DLVO theory and interactions at the molecular level are elucidated by computing the

Aggregation kinetics of the 2D NMs are determined using time resolved

ACS Paragon Plus Environment

Page 6 of 43

Page 7 of 43

Environmental Science & Technology

146

van der Waal (vdW) energies between two layers of 2D NMs via summing all pair-wise inter-

147

atomic interactions with the aid from molecular modeling.

148

2. Materials and Methods

149

2.1

150

exfoliated and dispersed in aqueous media in the same method which was already established in

151

the literature and described elsewhere.15 Briefly, 30 mg of MoS2, WS2, or h-BN ultrafine

152

powder (99%, Graphene Supermarket, Calverton, NY) was taken in 25 mL beaker and 10 mL of

153

45:55 mixture of ethanol and deionized (DI) water (by volume) was added into it as dispersant.

154

The beaker was then placed in an ice bath and the mixture was sonicated using a probe tip

155

sonicator Q700 (Qsonica, Newtown, CT) for 60 min. The amplitude of the sonicator was set to

156

50 corresponding to a power input of 12-15 W, and the pulse on and pulse off times were set to 8

157

s and 2 s, respectively. After sonication, the contents of the beaker were poured into a 15 mL

158

centrifuge tube and centrifuged at 3000 rpm for 20 minutes with Centrifuge 5810 R (Eppendorf

159

AG, Hauppauge, NY). Next, the supernatant was collected from the centrifuge tube and stored

160

in 20 mL scintillation vial, wrapped in aluminum foil, at 34 ºF until further use and between

161

experiments. A flow-diagram of this experimental protocol for preparing aqueous dispersion of

162

inorganic 2D NM has been provided in the supporting information (Figure S1).

163

2.2

164

99+%, Acros Organics) salts of a wide range of concentrations were used to replicate natural

165

aquatic environment.

166

dissolving required amount of their powders in DI water and then passing the solutions through

167

0.22 µm cellulose acetate membrane filters. The stock solutions were then diluted with DI water

168

to produce electrolyte solutions of required cation concentrations. In order to determine the

Aqueous 2D NM Suspension Preparation.

All three inorganic 2D NMs were

Solution Chemistry. Mono- (NaCl, 99.8%, Fischer Scientific) and di-valent (CaCl2,

Stock solutions of NaCl (5 M) and CaCl2 (1 M) were prepared by

ACS Paragon Plus Environment

Environmental Science & Technology

169

effect of natural organic matter on 2D NM aggregation, two different electrolyte conditions were

170

used in the presence and absence of SRNOM. These two electrolyte conditions were (1) a

171

mixture 7 mM NaCl and 1 mM CaCl2 which gives an equivalent ionic strength of 10 mM and is

172

referred to as low ionic strength; and (2) a salt solution of 100 mM NaCl with equivalent ionic

173

strength of 100 mM which is referred to as high ionic strength. SRNOM powder was purchased

174

from International Humic Substances Society (St. Paul, Mn) and first dispersed in DI water

175

under magnetic stirring for 6 h and then was filtered through 0.22 µm cellulose acetate

176

membrane filters to obtain a 400 mg/L SRNOM stock. This was further diluted to prepare 2.052

177

mg/L TOC solution when added to the 2D NM suspension for aggregation experiments at both

178

low and high ionic strength conditions.

179

2.3

180

and electrophoresis techniques were used to characterize the physicochemical properties of

181

exfoliated 2D NM suspensions. First, a Cary 60 UV-visible spectroscope (Agilent Technologies,

182

Santa Clara, CA) was used to characterize all three 2D NMs in aqueous dispersions for their

183

respective unique UV-vis spectroscopic signatures. Baseline was established with 45:55 ethanol-

184

water mixture contained in a cleaned quartz cuvette.

185

introduced individually in the cuvette to obtain spectral scans (three scans per 2D NM

186

suspension) for a spectral range of 100 to 900 nm.

Aqueous 2D NM Characterization. UV-visible spectroscopy, HRTEM, STEM, AFM,

Each 2D NM suspension was then

187

For performing HRTEM and STEM, carbon-coated copper TEM grids (Tedpella Inc,

188

Redding, CA) were prepared by adding two to three drops of each 2D NM suspension and air-

189

drying for few minutes. Electron micrographs were obtained at an accelerating voltage of 200

190

kV and a point resolution of 0.19 nm with a JEOL JEM 2010 HRTEM (JEOL USA, Inc.,

191

Peabody, MA), located at the Integrated Nanostructured Systems Instrumentation Facilities

ACS Paragon Plus Environment

Page 8 of 43

Page 9 of 43

Environmental Science & Technology

192

(INSIF) at the University at Buffalo to characterize the physical morphology of exfoliated 2D

193

NMs. Furthermore, STEM imaging was performed at an accelerating voltage of 20 kV using a

194

Hitachi SU70 field emission scanning electron microscope (FESEM). For each 2D NM type, 60

195

individual flakes were analyzed using imageJ software to estimate their lateral dimensions i.e.,

196

planar flake area distributions.

197

AFM imaging was performed to measure the thickness of exfoliated 2D NMs. For this, few

198

microliters of as prepared 2D NM aqueous dispersion was diluted with ethanol and drop casted

199

on silicon wafer which were than imaged in tapping mode using a Bruker Dimension Icon AFM.

200

The electrokinetic behavior of the 2D NMs was determined by measuring the electrophoretic

201

mobility (EPM) using a Malvern ZetaSizer Nano ZS (Malvern Instruments Ltd., Westborough,

202

MA). pH of the suspensions was kept unchanged to avoid introduction of additional ionic

203

species and was measured to be between 4.6 and 4.9. Then, appropriate dilutions of NaCl or

204

CaCl2 salts were added to 2D NM suspensions to get desired salt strength in the final volume of

205

2 mL. Four replicate measurements were taken for each different salt concentration and for each

206

sample.

207

2.4

208

used to measure aggregation kinetics of the 2D NM suspensions. The Malvern ZetaSizer Nano

209

ZS containing a 4 mW He-Ne 633 nm laser was used to perform the TRDLS and to measure the

210

hydrodynamic diameter of the 2D NM dispersions. 2 mL of MoS2, WS2 or h-BN suspensions,

211

each diluted to 100 times of their stock suspension concentrations, with added salt or SRNOM

212

was mixed in Malvern Zetasizer cuvettes, vortexed, and placed inside the ZetaSizer chamber.

213

Continual measurements at a time interval of 15 seconds were taken for 30 min to allow for at

214

least 30% increase in the initial average hydrodynamic radius.

Aggregation Kinetics Studies. Time resolved dynamic light scattering (TRDLS) was

ACS Paragon Plus Environment

Environmental Science & Technology

The initial aggregation rates, k, were determined from the slope of the aggregation history

215 216

Page 10 of 43

i.e., time vs. hydrodynamic diameter (Dh) plots.

217

𝑘∝

1 𝑑𝐷ℎ(𝑡) 𝑁𝑜 𝑑𝑡

[

]

(1) 𝑡→0

218

The initial slope was calculated by linear regression of the hydrodynamic diameter values up

219

to 1.3 times their initial values measured by the instrument at t = 0. For comparatively low ionic

220

strengths, i.e., up to 8.5 - 30 mM NaCl or 0.55 – 1 mM CaCl2, 30% increase in initial diameter

221

was not achieved within 30 min. In such scenario, the slope of the initial linear region of the

222

hydrodynamic diameter versus time plot was estimated to determine the aggregation rate for

223

each ionic strength. Attachment efficiency, α, is then determined by normalizing the initial rate

224

of change of hydrodynamic diameters (or the aggregation rate) at each solution condition by that

225

in the favorable regime, following equation (2).

226

𝛼=

[ [

𝑑𝐷ℎ(𝑡) 𝑑𝑡

] ]

𝑡→0

(2)

𝑑𝐷ℎ(𝑡) 𝑑𝑡

𝑡→0, 𝑓𝑎𝑣

227 228

2.5

Molecular Modeling. To probe the vdW interaction energies between layered 2D NMs,

229

the molecular-level energies of a system composed by two layers of MoS2, WS2 or hBN

230

nanostructures as a function of separation distance were computed using the open-source

231

LAMMPS package.51 For this, we built a number of initial configurations of two parallel AB-

232

stacking MoS2 (or WS2 or h-BN) sheets at separation distances (i.e., the shortest distance

233

between the two layers) ranging from 2.0 Å to 12.0 Å at an interval of 0.1 Å. These MoS2 (or

234

WS2 or h-BN) sheets have a surface area of approximately 70 x 70 Å2 and are assumed to be

235

rigid. Periodic boundary conditions were applied along the directions parallel to the sheets. The

ACS Paragon Plus Environment

Page 11 of 43

Environmental Science & Technology

236

total vdW energy of each of the systems was computed via summing all pair-wise inter-atomic

237

vdW interactions between layers. In these calculations, the 6-12 Lennard-Jones (L-J) potential,

238

with parameters taken from the Universal force field (UFF),52 was used to describe the

239

interactions between MoS2 – MoS2 (or WS2 – WS2 or h-BN – h-BN), and the Lorentz-Berthelot

240

mixing rule was applied to estimate the L-J parameters between dissimilar atoms. The potential

241

was truncated and shifted to zero at a cutoff radius of 12 Å.

242

2.6

243

the 2D NMs and determine if the classical DLVO theory can capture their aggregation behavior,

244

the following procedure adopted from a recent article explaining GO aggregation behavior using

245

a modified DLVO theory was followed.41 The total interaction energy was obtained from the

246

sum of van der Waals interaction energy (WvdW) and electrostatic double layer interaction energy

247

(WEL). The net interaction energy for the 2D TMDs i.e., MoS2 and WS2 will be discussed first as

248

they follow similar calculation procedure.

249 250

251

Net Interaction (DLVO) Energy Calculation. In order to account for the geometry of

WvdW between two MoS2 nanosheets was calculated using the potential law for two thin slabs:

𝑊𝑣𝑑𝑊 = ―

𝐴𝐻 (12𝜋 ){(𝑑1 ) + ((𝑑 + 𝑡1+ 𝑡 ) ) ― ((𝑑 +1𝑡 ) ) ― (𝑑 +1𝑡 ) } 2

1

2

2

1

2

2

2

(3)

252

where, t1 and t2 are the thicknesses of interacting slabs and are taken as 0.65 nm each for

253

MoS2,53 d is the sheet separation distance, A is the surface area of the sheets, and H is the

254

Hamaker constant. The Hamaker constant between two media contained in a third medium can

255

be determined from the following equation:40

256

𝐻𝑡𝑜𝑡𝑎𝑙 ≈

3ℎ𝑣𝑒 8 2

×

(𝑛21 ― 𝑛23)(𝑛22 ― 𝑛23) (𝑛21 + 𝑛23)0.5(𝑛22 + 𝑛23)0.5{((𝑛21 + 𝑛23)0.5(𝑛22 + 𝑛23)0.5)}

ACS Paragon Plus Environment

(4)

Environmental Science & Technology

Page 12 of 43

257

where, n1 = n2 = n is the refractive index of MoS2 in visible regime, n3 is the refractive index

258

of water (1.33), h is the Planck’s constant (6.626 × 10 ―34 Js), ve is the main absorption frequency

259

in the UV region. The refractive index of MoS2 depends on its number of layers.54, 55 For this

260

study, the upper and lower boundaries of the refractive index i.e., for bulk and monolayer MoS2

261

respectively, were found using the following an empirical equation which is strictly valid for

262

covalent oxides, nitrides, and sulfides:56

263

𝑛=

(

1 16 2.8 10

8.8 × 𝜔𝑈𝑉

)

(5)

264

where, 𝜔𝑈𝑉 is absorption frequency corresponding to band gap. For bulk and monolayer

265

MoS2, 𝜔𝑈𝑉 is 1.29 eV and 1.9 eV, respectively.1 The n for monolayer and bulk MoS2 are 3.4 and

266

3.9, respectively. For our calculation, an intermediate value of 3.7 is taken for n since the

267

number of sheets formed from our exfoliation method is between 3 to 4.15

268

The main electronic absorption frequency in the ultraviolet (UV) region, ve, is found from the

269

following equation40

270

𝑣𝑒 = 𝑣𝐼

3

(6)

𝑛2 + 2

271

where, 𝑣𝐼 = absorption frequency of a Bohr atom (3.3 × 1015 s ―1). For n = 3.7, we obtained a

272

𝑣𝑒 value of 1.443 × 1015 s ―1. Using the aforementioned parameters, we obtained a Hamaker

273

constant value of 296 × 10 ―21 J for MoS2. The parameters used for WS2 and h-BN are tabulated

274

in the supporting information, SI (Table S1).

275

constant for WS2 was found to be 320 × 10 ―21 J. For h-BN, the Hamaker constant was estimated

276

using the ratio of favorable interaction energies of h-BN to MoS2, obtained from the energy

277

minimum values computed for molecular systems of two layered 2D NMs of each type as

Applying the same approach, the Hamaker

ACS Paragon Plus Environment

Page 13 of 43

Environmental Science & Technology

278

described above (Section 2.5 Molecular modeling). The vdW interaction energies between

279

layered MoS2 and between layered WS2 were found to be similar, however, the vdW interaction

280

energies between layered h-BN were found substantially higher. The ratio obtained, 2.58,

281

indicates the substantially stronger vdW interaction energy between layered h-BN relative to that

282

between layered MoS2 (and also relative to that between layered WS2). By multiplying the

283

Hamaker constant of MoS2 with this ratio, a Hamaker constant value of 764 × 10 ―21 J was

284

calculated for h-BN, which was then used to calculate the vdW contributions between 2D h-BN

285

nanosheets in the DLVO theory.

286 287 288

The electrostatic double layer (EL) interaction energy between two identical surfaces can be obtained from the following equation:57 𝑊𝐸𝐿 ≈ 2𝐴𝜀𝑟𝜀0𝜅𝜉2𝑒 ―𝜅𝑑

(7)

289

where, 𝜀𝑟= dielectric constant (78.54 at 25 °C), 𝜀0 = permittivity of vacuum (8.854 × 10 ―12 F/m),

290

𝜉 = zeta potential (in Volt), 𝜅 = inverse of debye length (m-1).

291 292 293

The total interaction energy, 𝑊𝐷𝐿𝑉𝑂, is found by adding 𝑊𝑣𝑑𝑊 from equation (3) to twice the value of 𝑊𝐸𝐿 from equation (7) to account for both sides of the sheets being charged.57 𝑊𝐷𝐿𝑉𝑂 = 𝑊𝑣𝑑𝑊 +2𝑊𝐸𝐿

(8)

294 295

3. Results and Discussion

296

3.1 Chemical Identity and Morphological Characteristics. The sonication of inorganic 2D

297

NM in ethanol-water mixture followed by centrifugation resulted in stable aqueous dispersions

298

of exfoliated 2D NMs (supernatants) with different colors i.e., dark green, medium green, and

299

milky white dispersion for exfoliated MoS2, WS2, and h-BN, respectively. This is consistent

300

with previous literature reports.15, 58 The chemical identities of these exfoliated 2D NMs, i.e.,

ACS Paragon Plus Environment

Environmental Science & Technology

301

MoS2, WS2, and h-BN nanosheets in aqueous dispersions, were further confirmed by obtaining

302

their UV-vis spectra as presented in Figure 1 (a), (b), and (c), respectively. The spectrum

303

obtained for MoS2 dispersed in 45:55 ethanol-water mixture shows two absorbance peaks at 627

304

nm and 672 nm for exfoliated MoS2 (Figure 1a) that can be attributed to characteristic A1 and B1

305

direct excitonic transitions.15, 59

306

638 nm, which is the characteristic of excitonic A band, while a shoulder associated with indirect

307

B excitonic transition at 525 nm is not prominent (Figure 1b).60 The peaks indicate that the

308

exfoliated 2D TMD nanosheets are dispersed in the aqueous dispersion as 2H – phase.61 These

309

absorption peaks arise due to energy split of valence-band and spin-orbital coupling.61 Both

310

these results are consistent with values reported in literature.59, 60 The absorption spectrum of

311

aqueous h-BN dispersions does not show any characteristic peak due to its large band gap

312

(Figure 1c).15

313

transparent and transmits more than 99% of the light in the wavelength range of 250 to 900 nm.62

314

The physical morphology of the exfoliated 2D NMs have been determined by HRTEM,

315

STEM, and AFM. Figure S2 presents with the HRTEM images of bulk and exfoliated MoS2

316

where contrast differences between the images help to distinguish between hundreds of 2D

317

layers in the bulk form (high contrast) vs only few layers in the exfoliated form (low contrast).

318

The HRTEM images of exfoliated 2D NMs as shown in (Figure 1 (d), (e), and (f)) show low

319

contrast of MoS2, WS2, and h-BN nanosheets which implies that each of these exfoliated

320

samples has few layers of 2D NM sheets.63 These images also depict irregular sheet-like

321

structures with lateral sizes which range from nanoscale to several microns as observed

322

elsewhere.64 Exfoliated h-BN have much smaller particle sizes than MoS2 and WS2 and this

323

observation is consistent with literature.15, 58

The spectrum from exfoliated WS2 shows prominent peak at

However, the spectrum decays exponentially as h-BN nanosheet is highly

ACS Paragon Plus Environment

Page 14 of 43

Page 15 of 43

Environmental Science & Technology

324

Size distribution for the 2D nanomaterials were calculated by identifying 60 separate sheets

325

for each type of 2D NMs from their respective STEM images using imageJ. This method has

326

been established in literature for estimating and reporting 2D NM particle size distribution.65-67

327

Figure S3(a-c) show the frequency distribution of flake area for exfoliated MoS2, WS2, and h-

328

BN, respectively. For MoS2 and WS2, the dominant range of flake area was below 1 µm2

329

representing more than 75% and 60% of the distribution respectively. However, the flake areas

330

for h-BN were significantly smaller, all the flakes having areas below 0.1 µm2. Such differences

331

in lateral sizes between exfoliated TMD and h-BN nanosheets have been evidenced in the

332

literature.15, 58

333

AFM imaging of the exfoliated 2D NMs provide with their thickness information. Figure

334

S4(a-c) present AFM images of exfoliated MoS2, WS2, and h-BN, respectively; while Figure

335

S4(d-f) present with their corresponding height profiles. All the three types of 2D NMs had

336

thicknesses between 4-6 nm which indicate towards the presence of 3-4 layers of 2D nanosheets.

337

This is consistent with the literature reported values for the thickness of inorganic 2D NMs

338

exfoliated by the same method.15 All these different physicochemical characterization results

339

confirm the successful exfoliation of the inorganic 2D NMs in aqueous dispersions.

340

Figure 1. UV-vis spectra of aqueous dispersions of exfoliated (a) MoS2, (b) WS2, and (c) h-BN

341

nanosheets. Representative HRTEM micrographs of exfoliated (d) MoS2 (e) WS2, and (f) h-BN

342

nanosheets in aqueous dispersions.

343 344

3.2

Electrokinetic Properties. As mentioned above, the stable aqueous dispersions of 2D

345

NMs were prepared by their exfoliation in a 45:55 ethanol-water mixture – a procedure that was

346

established for the utilization of a mixed solvent system (i.e., the ethanol—water mixture) that

347

was selected based on the solvents’ Hansen solubility parameters.15 The ethanol molecules act

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 43

348

like surfactant molecules as their non-polar –CH3 groups cluster together and are absorbed on the

349

hydrophobic 2D NMs.68 The hydrophilic –OH groups protrude from the nanosheets and are

350

most likely responsible for the surface charge that keeps the dispersion stable.50 A colloidal

351

dispersion is generally considered stable if its zeta potential is more positive than +30 mV or

352

more negative than -30 mV.69 The as synthesized aqueous 2D NMs were all negatively charged

353

with zeta potentials of -37.22.4, -40.30.9, and -44.31.5 mV for MoS2, WS2, and h-BN

354

respectively, indicating good colloidal stability in aqueous dispersion in the absence of ionic

355

species. The zeta potential and electrophoretic mobility (EPM) values for all salt concentrations

356

tested are given in Table S2, while figures 2 (a-b) show the EPM values of MoS2, WS2, and h-

357

BN over a broad range of mono- and di-valent cations, however within a pH range of 4.6-4.9.

358

For each salt concentration, WS2 has higher EPM than MoS2 indicating higher electrostatic (or

359

zeta) potential for WS2.

360

difference in their electron affinities with WS2 having a higher electron affinity of 4.5 eV than

361

MoS2 which has an electron affinity of 3.56 eV.70, 71 For all three 2D NMs at low ionic strength

362

of only 0.001 M NaCl the zeta potential values slightly increased from their as synthesized zeta

363

potential. Such behavior is not uncommon among colloids which show decreasing absolute

364

value of zeta potential only at higher electrolyte conditions.72

365

concentrations, the absolute values of zeta potentials decreased consistently in accordance to the

366

Gouy-Chapmann double layer model.73 Electrokinetic measurements obtained in this study

367

show that for electrolyte concentrations of up to 0.01 M NaCl and 0.001 M CaCl2, h-BN

368

possesses the highest EPM, followed by WS2 and MoS2.

369

aforementioned values, however, EPM values of h-BN decreased faster and are mostly lower

370

than those for the TMDs.

This discrepancy between these 2D TMDs could arise from the

ACS Paragon Plus Environment

For higher electrolyte

For concentrations above the

Page 17 of 43

Environmental Science & Technology

371

The absolute values of EPM decreased more with the increase in CaCl2 concentration than

372

with the increase in NaCl concentration as the divalent counter ions (i.e., Ca2+) screen more of

373

the electric double layer and induce a smaller Debye length than the monovalent counter ions

374

(i.e., Na+).74 Furthermore, specific adsorption of Ca2+ to surfaces and short-ranged attractive

375

non-DLVO forces originating from ion-ion correlations charge fluctuations, surface charge

376

heterogeneities, and depletion forces have also been considered as probable causes behind the

377

marked decrease in diffuse layer potential.75

378 379

Figure 2. Electrophoretic mobility (EPM) values of aqueous MoS2, WS2, and h-BN suspensions

380

as a function of (a) monovalent NaCl and (b) divalent CaCl2 electrolytes. Measurements were

381

carried out at pH 4.6 to 4.9 with no additional buffers and at 25 ºC.

382 383

3.3

Aggregation Kinetics. Aggregation kinetics parameters of MoS2, WS2, and h-BN have

384

been determined from their respective aggregation history profiles as shown in Figure S5. The

385

initial hydrodynamic diameters of MoS2, WS2, and h-BN, as obtained from TRDLS, are 195.3

386

± 4.8, 142.5 ± 2.6, and 211.3 ± 3.9 nm respectively. Without any salt addition, the

387

dispersions remained stable i.e., the average hydrodynamic diameters did not change over the

388

duration of DLS measurement and even up to several months after synthesis.

389

increasing mono- and divalent (NaCl and CaCl2) salt concentrations, the surface charge on the

390

2D NMs gets screened thereby expediting aggregation.

391

hydrodynamic diameters (i.e., the aggregation rates) with electrolytic concentration also

392

increases. Attachment efficiency (α) has been estimated by normalizing initial aggregation rate

393

at any condition with that at the very fast, favorable regime. Figure 3 shows a log-log plot of

394

attachment efficiency values for MoS2, WS2, and h-BN dispersions as functions of

ACS Paragon Plus Environment

With the

Thus, the rate of increase of

Environmental Science & Technology

Page 18 of 43

395

concentrations of mono- and divalent (NaCl and CaCl2) salts. These plots of the 2D NMs,

396

otherwise known as stability plots, display distinct unfavorable or reaction-limited (RLCA) and

397

favorable or diffusion-limited (DLCA) regimes implying agreement with classical DLVO-type

398

interaction. It is observed that the attachment efficiency increases as the ionic strength of the

399

salts is increased. This is due to the suppression of range and magnitude of the electric double

400

layer, and thereby the height of the energy barrier.76

401

The transition between the two regimes, i.e. from RLCA to DLCA, is used to estimate

402

critical coagulation concentration (CCC). The CCC values have been approximated as 37 mM,

403

60 mM, and 19 mM NaCl, and 3 mM, 7.2 mM, and 1.3 mM CaCl2 for MoS2, WS2, and h-BN

404

respectively. This is the first report of CCC values for WS2 and h-BN dispersions. However,

405

there has been one study that determined the CCC of MoS2 from salt-initiated aggregation.77

406

According to that study, the estimated CCC values for MoS2 were 4.4 mM NaCl and 0.036 mM

407

CaCl2, which are significantly different from those obtained in our study. That previous study

408

was performed using sodium cholate assisted aqueous dispersion of MoS2, while in our case, the

409

MoS2 (or all 2D NMs) was exfoliated in ethanol-water mixture and the difference possibly

410

contributed to different surface properties and aggregation behavior.

411

The lower CCC values for the divalent salt compared to the monovalent salt is in

412

accordance to the Schulze-Hardy rule, even though the 2D NMs are not spherical.

413

phenomenon has also been observed for the 2D NM GO.78 Quantitatively, the Schulze-Hardy

414

rule translates to proportionality of CCC to Z-n , where Z is the counter ion valence (Z equals to

415

1 for Na+ and 2 for Ca2+) and n is 2 or 6 for low and high zeta potentials, respectively.79 For our

416

study, the exponents have been found to be 3.62, 3.06, and 3.87 for MoS2, WS2, and h-BN

417

respectively, which lie within the 2 to 6 range dictated by the Schulze-Hardy rule even though

ACS Paragon Plus Environment

This

Page 19 of 43

Environmental Science & Technology

418

the 2D NMs are not spherical. The n values for GO have also been found to vary between 5 to 6,

419

conforming to the Schulze-Hardy rule for spherical particles.37, 80, 81

420 421

Figure 3. Stability plots of aqueous MoS2, WS2, and h-BN suspension as a function of (a)

422

monovalent NaCl and (b) divalent CaCl2 electrolytes. Measurements were carried out at pH 4.6

423

to 4.9 with no additional buffers and at 25 ºC.

424 425

3.4

DLVO Prediction for Aggregation.

DLVO theory can quantitatively predict

426

aggregation behavior by estimating the CCC.41 It also identifies which of the DLVO forces,

427

electrostatic repulsion or vdW attraction, contributes to CCC for a particular material. First, we

428

made an effort to predict the CCC values for MoS2, WS2, and h-BN by solving equations (1) to

429

(6) for various electrolyte concentrations. The input parameters in these equations include zeta

430

potential values and thickness of the nanosheets. Performing calculations using these equations

431

resulted in theoretically predicted CCC values as 55 mM, 85 mM, and 100 mM NaCl, and 5.5

432

mM, 10 mM, and 10 mM CaCl2 for MoS2, WS2, and h-BN, respectively.

433

experimentally obtained values of CCC may vary from theoretical predictions by a factor of 2.73

434

The factor by which our modified DLVO prediction deviates from experimental observation

435

varies between 1.3 to 1.8 for the TMDs i.e., MoS2 and WS2. However, the factor varies by

436

almost an order of magnitude for h-BN. Thus, it is evident that the modified DLVO theory

437

captured the aggregation behavior of MoS2 and WS2 well but didn’t perform well for h-BN. A

438

close analysis into the apparently aberrant prediction of h-BN revealed that the Hamaker

439

constant (107 × 10 ―21 J ) obtained for h-BN using equations (4) to (6) does not reflect the fast

440

aggregation tendency of h-BN observed in experiments. Since the B-N bond in h-BN is ionic,

441

equation (5) cannot be used to calculate its refractive index and thus equation 4 cannot be used to

ACS Paragon Plus Environment

Typically,

Environmental Science & Technology

442

obtain their Hamaker constant of h-BN. Also, no other Hamaker constant was found for h-BN in

443

the literature.

444

Figure 4. vdW energy profiles of bilayer (a) MoS2 and (b) WS2, and (c) h-BN structures as a

445

function of layer separation distance calculated by summing of all pair-wise inter-atomic

446

interactions between layers using molecular modeling.

447

between MoS2 and WS2.

Inset shows difference in energies

448 449

To address this discrepancy and also to further elucidate the aggregation behavior of 2D

450

NMs, we computed the molecular level interaction energies between layered MoS2, WS2, and h-

451

BN nanosheets individually using molecular modeling (as described above in Section 2.5).

452

Figure 4 shows the vdW energy profiles of bilayer MoS2, WS2, and h-BN structures as a function

453

of separation distance. It can be observed that the separation distance dependent energy profiles

454

and the first minima at a separation distance of approximately 3.5 Å for layered MoS2 and WS2

455

almost overlap with each other indicating similar vdW attraction forces (Figure 4 inset). Based

456

on the ratio of vdW interaction energies at the first minima for WS2 to MoS2 (which was slightly

457

more than 1) and the previously calculated Hamaker constant of 296 × 10 ―21 J for MoS2 (from

458

equation 4), we obtained a Hamaker constant value of 303 × 10 ―21 𝐽 for WS2, which is only

459

slightly less than its previously calculated Hamaker constant of 320 × 10 ―21 𝐽 and provide with the

460

same predicted CCC values for WS2. These two similar Hamaker constant values and the same

461

predicted CCC value of WS2 obtained from two independent computational approach (i.e., using

462

equation 4 and using molecular modeling) validate the use of molecular modeling technique for

463

predicting Hamaker constants for like 2D NMs such as h-BN. Figure 4 also presents that the

464

vdW interaction energy profiles and first minima values are significantly higher (for attraction)

465

for layered h-BN compared to layered TMDs indicating towards much stronger interactions

ACS Paragon Plus Environment

Page 20 of 43

Page 21 of 43

Environmental Science & Technology

466

between h-BN nanosheets. Thus, using the ratio of vdW interaction energies of h-BN to MoS2 at

467

the first minima, we were able to determine a new Hamaker constant of 764 × 10 ―21 J. A

468

previous study used a similar approach to reasonably estimate the Hamaker constant values of

469

higher order fullerenes (i.e., for C70, C76, and C84) based on known Hamaker constant of fullerene

470

C60 and their molecular modeling.82 Using this new Hamaker constant and equations (3), (7),

471

and (8) we then computed the net interaction energy profiles of h-BN for a range of electrolytic

472

strengths to obtain its CCC value. Figure 5 shows the total interaction energy plots as a function

473

of separation distance predicted by the DLVO theory for the three 2D NMs for a range of NaCl

474

and CaCl2 concentrations. From Figure 5, this modified DLVO theory predicted the CCC of h-

475

BN to be 30 mM NaCl and 5.5 mM CaCl2, which are in much better agreement with the

476

experimentally obtained values than previously estimated using equation 4.

477

experimentally obtained CCC values and theoretically predicted CCC values in Table 1.

478

Figure 5. DLVO interaction energy between two 2D NM (MoS2, WS2 or h-BN) sheet in

479

presence of various concentrations of NaCl and CaCl2.

We listed the

480 481

Table 1. Experimentally obtained and modified DLVO theory predicted CCC values of MoS2,

482

WS2, and h-BN nanosheets.

483 484

According to our observed and predicted values of CCC, the colloidal stability of the studied

485

inorganic 2D NMs is in the following order: h-BN < MoS2 < WS2. Our calculated values of

486

Hamaker constants are 764 × 10 ―21 𝐽 , 296 × 10 ―21, and 320 × 10 ―21 𝐽 for h-BN, MoS2, and

487

WS2, respectively. These Hamaker constant values suggest that the vdW attraction force is the

488

highest for h-BN among these three 2D NMs and exceeds significantly compared to the TMDs

489

i.e., MoS2 and WS2. However, this was not the case for the relative colloidal stability of MoS2

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 43

490

and WS2 i.e., WS2 being more colloidally stable than MoS2, although WS2’s Hamaker constant

491

was slightly higher than that of MoS2. This suggests that the electrostatic force of repulsion is

492

the governing DLVO force dictating the colloidal behavior of MOS2 and WS2. The force

493

calculations showed that the electrostatic repulsion forces of WS2 significantly exceeded the

494

repulsion forces for MoS2, at all separation distances and for different ionic strength conditions.

495 496

3.5

Role of Natural Organic Matter on Aggregation. The presence of 2.025 mg/L TOC

497

SRNOM significantly enhanced the stability of 2D NM dispersions and decreased their

498

aggregation rates at both low (7 mM NaCl and 1 mM CaCl2; Figure 6a) and high (100 mM NaCl;

499

Figure 6b) electrolyte concentrations. All the inorganic 2D NM dispersions were significantly

500

stabilized by the presence of SRNOM. Without SRNOM, MoS2 and WS2 had aggregation rates

501

of respectively 0.06 and 0.16 nm/s at low ionic strength and respectively 0.13 and 0.30 nm/s at

502

high ionic strength. However, the presence of SRNOM completely diminished their aggregation

503

tendency. The aggregation rates of h-BN were reduced to 2.6- and 3.7-fold of their original

504

values of 0.31 and 0.88 nm/s for low and high ionic strengths, respectively.

505

aggregation rate of h-BN compared to the TMDs even in presence of SRNOM is in accordance

506

with our experimental finding that h-BN has the least CCC value or the highest vdW interaction

507

energy amongst the three 2D NMs studied. NOM consists of humic substances which are known

508

to stabilize nanomaterials by steric or electrosteric repulsion82, 83 in the presence of electrolytes.

509

NOM adsorbs on the surface of WS2 nanosheets probably by strong coordination of the carboxyl

510

groups with tungsten atoms.84 In case of h-BN, the aromatic fraction of the NOM possibly

511

interacts via Π-Π stacking forces.85

ACS Paragon Plus Environment

The higher

Page 23 of 43

Environmental Science & Technology

The ubiquity of NOM in natural aquatic systems plays an important role in the aggregation

512 513

and thus mobility of the 2D NMs in aquatic environment.

Additionally, NOM influences

514

generation of ROS and photochemical dissolution processes of TMDs, thereby governing

515

bioavailability of the NMs. The relative colloidal stability of the different inorganic 2D NMs

516

imparted by NOM thus holds immense environmental significance.

517 518

Figure 6. Aggregation rates of MoS2, WS2, and h-BN without and with 2.025 mg/L TOC

519

SRNOM and in presence of (a) 7 mM NaCl and 1 mM CaCl2, (b) 100 mM NaCl.

520 521

3.6

Environmental Implications. This study provides valuable insight into the relative

522

colloidal stability of exfoliated MoS2, WS2, and h-BN nanosheets in aqueous media. The

523

inorganic 2D NMs have distinct physicochemical properties which dictate their aggregation

524

behavior. While EDL repulsion appears to govern stability of the TMDs, a combination of non-

525

DLVO and vdW attraction forces might dictate the instability of h-BN. The critical coagulation

526

concentration (CCC) values of the 2D NMs suggest that h-BN will be the least stable while WS2

527

will be the most stable in presence of either mono- or divalent salts. The salinity of fresh and

528

brackish waters are approximately 500 ppm and 35000 ppm respectively while that of ground

529

water may vary from 0 to greater than 13,000 ppm depending on its origin, seawater intrusion,

530

leaching of salt sediments and anthropogenic activities.86 This implies that in freshwater 2D

531

NMs will be carried farther downstream from their point of release, thereby having a greater

532

extent of exposure to aquatic organisms. In brackish waters, the 2D NMs are more likely to

533

aggregate and settle since their CCC values are about an order or magnitude lower than typical

534

seawater salinity. However, at concentrations present in natural aquatic systems, SRNOM will

535

stabilize the 2D NMs i.e., suppress their aggregation, even at systems with high electrolyte

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 43

536

concentration such as marine and estuarine bodies. Thus, 2D NM mobility will depend on the

537

relative amount of NOM present in the brackish water. Similarly, the aggregation of 2D NMs in

538

ground water will depend on the salinity and NOM content with high salinity facilitating

539

aggregation and high NOM suppressing it. The complex mobility and transport of 2D NMs in

540

aquatic systems and their effect on native microorganisms mandate comprehensive evaluation of

541

their environmental risk.

542

Supporting Information

543

The Supporting Information is available free of charge on the ACS Publications website. The

544

supporting information includes figures showing experimental protocol for preparing aqueous

545

dispersions of inorganic 2D NMs; HRTEM images of bulk and exfoliated MoS2; particle size

546

distritbution for exfoliated 2D NMs; AFM images of exfoliated 2D NMs; table showing

547

parameters used for DLVO calculation; table for zeta potential and EPM values for 2D NM

548

dispersions with varied ionic strength; figure showing aggregation history profiles for 2D NM

549

dispersions at various electrolyte conditions.

550

Acknowledgements

551

We thank Prof. Yueling Qin from Department of Physics at the University at Buffalo (UB) for

552

his assistance with HRTEM imaging. We also thank Mr. Peter Bush, Director of South Campus

553

Instrument Center at UB, for his assistance with STEM imaging.

554

Department of Civil, Structural and Environmental Engineering (CSEE) at UB for supporting

555

Ms. Anusha Gupta’s travel to UB and research internship in Prof. Nirupam Aich’s laboratory.

556

References

ACS Paragon Plus Environment

Finally, we thank the

Page 25 of 43

Environmental Science & Technology

557

1.

Wang, Z. Y.; Mi, B. X., Environmental Applications of 2D Molybdenum Disulfide

558

(MoS2) Nanosheets. Environmental Science & Technology 2017, 51, (15), 8229-8244.

559

2.

560

9469.

561

3.

562

F.; Hong, S. S.; Huang, J. X.; Ismach, A. F.; Johnston-Halperin, E.; Kuno, M.; Plashnitsa, V. V.;

563

Robinson, R. D.; Ruoff, R. S.; Salahuddin, S.; Shan, J.; Shi, L.; Spencer, M. G.; Terrones, M.;

564

Windl, W.; Goldberger, J. E., Progress, Challenges, and Opportunities in Two-Dimensional

565

Materials Beyond Graphene. Acs Nano 2013, 7, (4), 2898-2926.

566

4.

Prasad, S.; Zabinski, J., Lubricants: super slippery solids. Nature 1997, 387, (6635), 761.

567

5.

Kimura, Y.; Wakabayashi, T.; Okada, K.; Wada, T.; Nishikawa, H., Boron nitride as a

568

lubricant additive. Wear 1999, 232, (2), 199-206.

569

6.

570

optoelectronics of two-dimensional transition metal dichalcogenides. Nature Nanotechnology

571

2012, 7, (11), 699-712.

572

7.

573

ultraviolet lasing of hexagonal boron nitride single crystal. Nat Mater 2004, 3, (6), 404-409.

574

8.

575

biomedical applications. Chemical Society Reviews 2015, 44, (9), 2681-2701.

576

9.

577

in Synthesis and Biomedical Applications of Two-Dimensional Transition Metal Dichalcogenide

578

Nanosheets. Small 2017, 13, (5).

Zhang, H., Ultrathin Two-Dimensional Nanomaterials. Acs Nano 2015, 9, (10), 9451-

Butler, S. Z.; Hollen, S. M.; Cao, L. Y.; Cui, Y.; Gupta, J. A.; Gutierrez, H. R.; Heinz, T.

Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S., Electronics and

Watanabe, K.; Taniguchi, T.; Kanda, H., Direct-bandgap properties and evidence for

Chen, Y.; Tan, C. L.; Zhang, H.; Wang, L. Z., Two-dimensional graphene analogues for

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

ACS Paragon Plus Environment

Environmental Science & Technology

579

10.

Xie, X. Q.; Ao, Z. M.; Su, D. W.; Zhang, J. Q.; Wang, G. X., MoS2/Graphene Composite

580

Anodes with Enhanced Performance for Sodium-Ion Batteries: The Role of the Two-

581

Dimensional Heterointerface. Advanced Functional Materials 2015, 25, (9), 1393-1403.

582

11.

583

Nanosheets and Performance as Li-Ion Battery Anodes. Journal of Physical Chemistry Letters

584

2012, 3, (11), 1523-1530.

585

12.

586

M., Hexagonal Boron Nitride-Based Electrolyte Composite for Li-Ion Battery Operation from

587

Room Temperature to 150 degrees C. Advanced Energy Materials 2016, 6, (12).

588

13.

589

Activated boron nitride as an effective adsorbent for metal ions and organic pollutants. Scientific

590

Reports 2013, 3.

591

14.

592

nanoparticles using WS2/g-C3N4 hybrid as co-catalyst. Transactions of Nonferrous Metals

593

Society of China 2017, 27, (5), 1117-1126.

594

15.

595

Strategy for Efficient Exfoliation of Inorganic Graphene Analogues. Angewandte Chemie-

596

International Edition 2011, 50, (46), 10839-10842.

597

16.

598

optoelectronics of two-dimensional transition metal dichalcogenides. Nature nanotechnology

599

2012, 7, (11), 699.

600

17.

601

structure of the transition metal sulfide TS2. Physical Review B 2011, 83, (24).

Bhandavat, R.; David, L.; Singh, G., Synthesis of Surface-Functionalized WS2

Rodrigues, M. T. F.; Kalaga, K.; Gullapalli, H.; Babu, G.; Reddy, A. L. M.; Ajayan, P.

Li, J.; Xiao, X.; Xu, X. W.; Lin, J.; Huang, Y.; Xue, Y. M.; Jin, P.; Zou, J.; Tang, C. C.,

Zheng, L. L.; Xiao, X. Y.; Li, Y.; Zhang, W. P., Enhanced photocatalytic activity of TiO2

Zhou, K. G.; Mao, N. N.; Wang, H. X.; Peng, Y.; Zhang, H. L., A Mixed-Solvent

Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S., Electronics and

Kuc, A.; Zibouche, N.; Heine, T., Influence of quantum confinement on the electronic

ACS Paragon Plus Environment

Page 26 of 43

Page 27 of 43

Environmental Science & Technology

602

18.

Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A., Single-layer MoS2

603

transistors. Nature Nanotechnology 2011, 6, (3), 147-150.

604

19.

605

heterostructures. Journal of Semiconductors 2017, 38, (3).

606

20.

607

Nanomaterials. Chemnanomat 2017, 3, (1), 5-16.

608

21.

609

Catalyst of α-Fe2O3/MoS2 Hierarchical Nanoheterostructures: Reutilization for Supercapacitors.

610

Scientific reports 2016, 6, 31591.

611

22.

612

X.; Xu, M. S., Antibacterial activity of two-dimensional MoS2 sheets. Nanoscale 2014, 6, (17),

613

10126-10133.

614

23.

615

2D nanomaterials. Environmental Science-Nano 2017, 4, (8), 1617-1633.

616

24.

617

exposure to manufactured nanoparticles. 2009; Vol. 11, p 1637-1650.

618

25.

619

engineered nanomaterials. Journal of Nanoparticle Research 2013, 15, (6).

620

26.

621

Aquatic and Terrestrial Environments. Accounts of Chemical Research 2013, 46, (3), 854-862.

622

27.

623

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

Huo, N. J.; Yang, Y. J.; Li, J. B., Optoelectronics based on 2D TMDs and

Kenry; Lim, C. T., Biocompatibility and Nanotoxicity of Layered Two-Dimensional

Yang, X.; Sun, H.; Zhang, L.; Zhao, L.; Lian, J.; Jiang, Q., High Efficient Photo-Fenton

Yang, X.; Li, J.; Liang, T.; Ma, C. Y.; Zhang, Y. Y.; Chen, H. Z.; Hanagata, N.; Su, H.

Fojtu, M.; Teo, W. Z.; Pumera, M., Environmental impact and potential health risks of

Schneider, T.; Jensen, K., Relevance of aerosol dynamics and dustiness for personal

Keller, A. A.; McFerran, S.; Lazareva, A.; Suh, S., Global life cycle releases of

Batley, G. E.; Kirby, J. K.; McLaughlin, M. J., Fate and Risks of Nanomaterials in

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

ACS Paragon Plus Environment

Environmental Science & Technology

624

the Toxicological Potential of 2D versus Aggregated Molybdenum Disulfide in the Lung. Small

625

2015, 11, (38), 5079-5087.

626

28.

627

nanoparticles. Langmuir 2006, 22, (26), 10994-11001.

628

29.

629

Bacteria. Acs Nano 2010, 4, (10), 5731-5736.

630

30.

631

and Graphene in Human Erythrocytes and Skin Fibroblasts. Acs Applied Materials & Interfaces

632

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

633

31.

634

Koski, K.; Hurt, R., Biological and environmental interactions of emerging two-dimensional

635

nanomaterials. Chemical Society Reviews 2016, 45, (6), 1750-1780.

636

32.

637

exfoliation. Nanoscale 2014, 6, (23), 14412-14418.

638

33.

639

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

640

the presence of electron donor. Environmental Science-Nano 2015, 2, (4), 370-379.

641

34.

642

mediated antibacterial activity of tungsten disulfide (WS2). Rsc Advances 2017, 7, (60), 37873-

643

37880.

644

35.

645

Biocompatibility of boron nitride nanosheets. Nano Research 2018, 11, (1), 334-342.

Chen, K. L.; Elimelech, M., Aggregation and deposition kinetics of fullerene (C-60)

Akhavan, O.; Ghaderi, E., Toxicity of Graphene and Graphene Oxide Nanowalls Against

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

Wang, Z. Y.; Zhu, W. P.; Qiu, Y.; Yi, X.; von dem Bussche, A.; Kane, A.; Gao, H. J.;

Chng, E. L. K.; Sofer, Z.; Pumera, M., MoS2 exhibits stronger toxicity with increased

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

Liu, X.; Duan, G. X.; Li, W. F.; Zhou, Z. F.; Zhou, R. H., Membrane destruction-

Mateti, S.; Wong, C. S.; Liu, Z.; Yang, W. R.; Li, Y. C.; Li, L. H.; Chen, Y.,

ACS Paragon Plus Environment

Page 28 of 43

Page 29 of 43

Environmental Science & Technology

646

36.

Derjaguin, B.; Landau, L., Theory of the stability of strongly charged lyophobic sols and

647

of the adhesion of strongly charged particles in solutions of electrolytes. Acta physicochim.

648

URSS 1941, 14, (6), 633-662.

649

37.

650

Colloidal Properties and Stability of Graphene Oxide Nanomaterials in the Aquatic Environment.

651

Environmental Science & Technology 2013, 47, (12), 6288-6296.

652

38.

653

Interactions of graphene oxide nanomaterials with natural organic matter and metal oxide

654

surfaces. Environmental science & technology 2014, 48, (16), 9382-9390.

655

39.

656

Aggregation and stability of reduced graphene oxide: complex roles of divalent cations, pH, and

657

natural organic matter. Environmental science & technology 2015, 49, (18), 10886-10893.

658

40.

Israelachvili, J. N., Intermolecular and surface forces. Academic press: 2011.

659

41.

Gudarzi, M. M., Colloidal Stability of Graphene Oxide: Aggregation in Two Dimensions.

660

Langmuir 2016, 32, (20), 5058-5068.

661

42.

662

dispersions of graphene nanosheets. Nature nanotechnology 2008, 3, (2), 101.

663

43.

664

two-dimensional layered transition metal dichalcogenide nanosheets. Nature Chemistry 2013, 5,

665

(4), 263-275.

666

44.

667

M. X.; Vajtai, R.; Lou, J.; Ajayan, P. M., Liquid Phase Exfoliation of Two-Dimensional

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

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

Chowdhury, I.; Mansukhani, N. D.; Guiney, L. M.; Hersam, M. C.; Bouchard, D.,

Li, D.; Müller, M. B.; Gilje, S.; Kaner, R. B.; Wallace, G. G., Processable aqueous

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

Shen, J. F.; He, Y. M.; Wu, J. J.; Gao, C. T.; Keyshar, K.; Zhang, X.; Yang, Y. C.; Ye,

ACS Paragon Plus Environment

Environmental Science & Technology

668

Materials by Directly Probing and Matching Surface Tension Components. Nano Letters 2015,

669

15, (8), 5449-5454.

670

45.

671

of molybdenum disulfide. Coordination Chemistry Reviews 2002, 224, (1-2), 87-109.

672

46.

673

nitrate stimulation after oil pollution in mangrove sediment revealed by Illumina sequencing.

674

Mar. Pollut. Bull. 2016, 109, (1), 281-289.

675

47.

676

species? Encapsulation and ordering of hard electropositive cations. Journal of the American

677

Chemical Society 1999, 121, (50), 11720-11732.

678

48.

679

hexagonal boron nitride. Journal of Physics-Condensed Matter 2006, 18, (1), 97-115.

680

49.

681

Nanosheets. Journal of Physical Chemistry Letters 2010, 1, (1), 277-283.

682

50.

683

Dispersions of Few-Layered and Monolayered Hexagonal Boron Nitride Nanosheets from

684

Sonication-Assisted Hydrolysis: Critical Role of Water. Journal of Physical Chemistry C 2011,

685

115, (6), 2679-2685.

686

51.

687

Computational Physics 1995, 117, (1), 1-19.

688

52.

689

periodic table force field for molecular mechanics and molecular dynamics simulations. Journal

690

of the American Chemical Society 1992, 114, (25), 10024-10035.

Benavente, E.; Santa Ana, M. A.; Mendizabal, F.; Gonzalez, G., Intercalation chemistry

Wang, L.; Huang, X.; Zheng, T. L., Responses of bacterial and archaeal communities to

Heising, J.; Kanatzidis, M. G., Exfoliated and restacked MoS2 and WS2: Ionic or neutral

Ooi, N.; Rairkar, A.; Lindsley, L.; Adams, J. B., Electronic structure and bonding in

Lin, Y.; Williams, T. V.; Connell, J. W., Soluble, Exfoliated Hexagonal Boron Nitride

Lin, Y.; Williams, T. V.; Xu, T. B.; Cao, W.; Elsayed-Ali, H. E.; Connell, J. W., Aqueous

Plimpton, S., Fast Parallel Algorithms for Short-Range Molecular Dynamics. Journal of

Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A.; Skiff, W. M., UFF, a full

ACS Paragon Plus Environment

Page 30 of 43

Page 31 of 43

Environmental Science & Technology

691

53.

Li, H.; Ma, L.; Chen, W. X.; Wang, J. M., Synthesis of MoS2/C nanocomposites by

692

hydrothermal route used as Li-ion intercalation electrode materials. Materials Letters 2009, 63,

693

(15), 1363-1365.

694

54.

695

refractive index of highly crystalline monolayer MoS2 with high confidence. Scientific reports

696

2015, 5.

697

55.

698

S., Investigation of the optical properties of MoS2 thin films using spectroscopic ellipsometry.

699

Applied Physics Letters 2014, 104, (10), 103114.

700

56.

701

interface science 1997, 70, 125-169.

702

57.

703

F. M.; De, S.; Wang, Z. M.; McGovern, I. T.; Duesberg, G. S.; Coleman, J. N., Liquid Phase

704

Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. Journal of the

705

American Chemical Society 2009, 131, (10), 3611-3620.

706

58.

707

Shin, E.; Kim, W.-G.; Lee, H.; Ryu, G. H.; Choi, M.; Kim, T. H.; Oh, J.; Park, S.; Kwak, S. K.;

708

Yoon, S. W.; Byun, D.; Lee, Z.; Lee, C., Direct exfoliation and dispersion of two-dimensional

709

materials in pure water via temperature control. Nature Communications 2015, 6, 8294.

710

59.

711

Andersson, M.; Hummelgard, M.; Olin, H., Exfoliated MoS2 in Water without Additives. Plos

712

One 2016, 11, (4).

Zhang, H.; Ma, Y.; Wan, Y.; Rong, X.; Xie, Z.; Wang, W.; Dai, L., Measuring the

Yim, C.; O'Brien, M.; McEvoy, N.; Winters, S.; Mirza, I.; Lunney, J. G.; Duesberg, G.

Bergström, L., Hamaker constants of inorganic materials. Advances in colloid and

Lotya, M.; Hernandez, Y.; King, P. J.; Smith, R. J.; Nicolosi, V.; Karlsson, L. S.; Blighe,

Kim, J.; Kwon, S.; Cho, D.-H.; Kang, B.; Kwon, H.; Kim, Y.; Park, S. O.; Jung, G. Y.;

Forsberg, V.; Zhang, R. Y.; Backstrom, J.; Dahlstrom, C.; Andres, B.; Norgren, M.;

ACS Paragon Plus Environment

Environmental Science & Technology

713

60.

Pagona, G.; Bittencourt, C.; Arenal, R.; Tagmatarchis, N., Exfoliated semiconducting

714

pure 2H-MoS2 and 2H-WS2 assisted by chlorosulfonic acid. Chemical Communications 2015,

715

51, (65), 12950-12953.

716

61.

717

Zhang, L.; Wang, J., Optical Limiting and Theoretical Modelling of Layered Transition Metal

718

Dichalcogenide Nanosheets. Scientific Reports 2015, 5.

719

62.

720

Lou, J.; Yakobson, B. I., Large scale growth and characterization of atomic hexagonal boron

721

nitride layers. Nano letters 2010, 10, (8), 3209-3215.

722

63.

723

molybdenum disulfide in aqueous media and its application in lithium ion batteries. Nanoscale

724

2015, 7, (21), 9919-9926.

725

64.

726

freestanding MoS 2 nanosheets. Nanoscale research letters 2013, 8, (1), 129.

727

65.

728

Quinn, A. J.; Zhou, W.; Blackley, R., Controllable selective exfoliation of high-quality graphene

729

nanosheets and nanodots by ionic liquid assisted grinding. Chemical Communications 2012, 48,

730

(13), 1877-1879.

731

66.

732

nanoplatelets: capacitance, potential of zero charge and diffusion coefficient. Chemical science

733

2015, 6, (5), 2869-2876.

Dong, N. N.; Li, Y. X.; Feng, Y. Y.; Zhang, S. F.; Zhang, X. Y.; Chang, C. X.; Fan, J. T.;

Song, L.; Ci, L.; Lu, H.; Sorokin, P. B.; Jin, C.; Ni, J.; Kvashnin, A. G.; Kvashnin, D. G.;

Liu, W.; Zhao, C.; Zhou, R.; Zhou, D.; Liu, Z.; Lu, X., Lignin-assisted exfoliation of

Gao, D.; Si, M.; Li, J.; Zhang, J.; Zhang, Z.; Yang, Z.; Xue, D., Ferromagnetism in

Shang, N. G.; Papakonstantinou, P.; Sharma, S.; Lubarsky, G.; Li, M.; McNeill, D. W.;

Poon, J.; Batchelor-McAuley, C.; Tschulik, K.; Compton, R. G., Single graphene

ACS Paragon Plus Environment

Page 32 of 43

Page 33 of 43

Environmental Science & Technology

734

67.

Lauritsen, J. V.; Kibsgaard, J.; Helveg, S.; Topsøe, H.; Clausen, B. S.; Lægsgaard, E.;

735

Besenbacher, F., Size-dependent structure of MoS2 nanocrystals. Nature Nanotechnology 2007,

736

2, 53.

737

68.

738

F., A rational design of cosolvent exfoliation of layered materials by directly probing liquid-solid

739

interaction. Nature Communications 2013, 4.

740

69.

741

measurements using light-scattering techniques. Philosophical Transactions of the Royal Society

742

a-Mathematical Physical and Engineering Sciences 2010, 368, (1927), 4439-4451.

743

70.

744

first principles. Physical Review B 2015, 91, (12), 125304.

745

71.

746

performance photocurrent generation from two-dimensional WS2 field-effect transistors. Applied

747

Physics Letters 2014, 104, (19), 193113.

748

72.

749

polystyrene latices, J. Colloid Interface Sei 1988, 122, 211.

750

73.

751

university press: 1989.

752

74.

753

Influence of Ionic Strength, pH, and Cation Valence on Aggregation Kinetics of Titanium

754

Dioxide Nanoparticles. Environmental Science & Technology 2009, 43, (5), 1354-1359.

Halim, U.; Zheng, C. R.; Chen, Y.; Lin, Z. Y.; Jiang, S.; Cheng, R.; Huang, Y.; Duan, X.

Kaszuba, M.; Corbett, J.; Watson, F. M.; Jones, A., High-concentration zeta potential

Komsa, H.-P.; Krasheninnikov, A. V., Native defects in bulk and monolayer MoS 2 from

Hwan Lee, S.; Lee, D.; Sik Hwang, W.; Hwang, E.; Jena, D.; Jong Yoo, W., High-

Midmore, R.; Hunter, R., The effect of electrolyte and co-ion type on the C-potential of

Russel, W. B.; Saville, D. A.; Schowalter, W. R., Colloidal dispersions. Cambridge

French, R. A.; Jacobson, A. R.; Kim, B.; Isley, S. L.; Penn, R. L.; Baveye, P. C.,

ACS Paragon Plus Environment

Environmental Science & Technology

755

75.

Trefalt, G.; Palberg, T.; Borkovec, M., Forces between colloidal particles in aqueous

756

solutions containing monovalent and multivalent ions. Current Opinion in Colloid & Interface

757

Science 2017, 27, 9-17.

758

76.

759

modelling and simulation. Butterworth-Heinemann: 2013.

760

77.

761

Optical Absorption of Layered MoS2 and DNA Oligonucleotides Induced Dispersion Behavior

762

for Label-Free Detection of Single-Nucleotide Polymorphism. Advanced Functional Materials

763

2015, 25, (23), 3541-3550.

764

78.

765

Morphological Transformation of Graphene Oxide in Aqueous Solutions Containing Different

766

Metal Cations. Environmental Science & Technology 2016, 50, (20), 11066-11075.

767

79.

768

Deposition of Engineered Nanomaterials in Aquatic Environments: Role of Physicochemical

769

Interactions. Environmental Science & Technology 2010, 44, (17), 6532-6549.

770

80.

771

oxide and its nanohybrids with magnetite and elemental silver under environmentally relevant

772

conditions. Journal of Nanoparticle Research 2018, 20, (4).

773

81.

774

Morphology and Surface Chemistry with Aggregation Behavior. Environmental Science &

775

Technology 2016, 50, (13), 6964-6973.

Elimelech, M.; Gregory, J.; Jia, X., Particle deposition and aggregation: measurement,

Li, B. L.; Zou, H. L.; Lu, L.; Yang, Y.; Lei, J. L.; Luo, H. Q.; Li, N. B., Size-Dependent

Yang, K. J.; Chen, B. L.; Zhu, X. Y.; Xing, B. S., Aggregation, Adsorption, and

Petosa, A. R.; Jaisi, D. P.; Quevedo, I. R.; Elimelech, M.; Tufenkji, N., Aggregation and

Park, C. M.; Wang, D. J.; Heo, J.; Her, N.; Su, C. M., Aggregation of reduced graphene

Jiang, Y.; Raliya, R.; Fortner, J. D.; Biswas, P., Graphene Oxides in Water: Correlating

ACS Paragon Plus Environment

Page 34 of 43

Page 35 of 43

Environmental Science & Technology

776

82.

Aich, N.; Boateng, L. K.; Sabaraya, I. V.; Das, D.; Flora, J. R. V.; Saleh, N. B.,

777

Aggregation Kinetics of Higher-Order Fullerene Clusters in Aquatic Systems. Environmental

778

Science & Technology 2016, 50, (7), 3562-3571.

779

83.

780

of Multiwalled Carbon Nanotube-Titanium Dioxide Nanohybrids: Probing the Part-Whole

781

Question. Environmental Science & Technology 2018, 52, (15), 8233-8241.

782

84.

783

as a Platform for Biosensing. Analytical Chemistry 2014, 86, (7), 3610-3615.

784

85.

785

Guo, Y. N.; Sarasua, J. R.; Tofail, S. A. M.; Zeugolis, D. I.; Pandit, A.; Biggs, M. J., Effects of

786

Polydopamine Functionalization on Boron Nitride Nanotube Dispersion and Cytocompatibility.

787

Bioconjugate Chemistry 2015, 26, (10), 2025-2037.

788

86.

789

Johnson, R. A. Locarnini, A. V. Mishonov, T.D. O'Brien, C.R. Paver, J.R. Reagan, D. Seidov, I.

790

V. Smolyar, and M. M. Zweng, World Ocean Database 2013. In NOAA Atlas NESDIS 72, 2013.

Das, D.; Sabaraya, I. V.; Zhu, T.; Sabo-Attwood, T.; Saleh, N. B., Aggregation Behavior

Yuan, Y. X.; Li, R. Q.; Liu, Z. H., Establishing Water-Soluble Layered WS2 Nanosheet

Fernandez-Yague, M. A.; Larranaga, A.; Gladkovskaya, O.; Stanley, A.; Tadayyon, G.;

Boyer, T. P., J. I. Antonov, O. K. Baranova, C. Coleman, H. E. Garcia, A. Grodsky, D. R.

791

ACS Paragon Plus Environment

Environmental Science & Technology

792

793 794 795 796

Figure 1. UV-vis spectra of aqueous dispersion of exfoliated (a) MoS2, (b) WS2, and (c) h-BN

797

nanosheets and representative HRTEM micrographs of exfoliated (d) MoS2 (e) WS2, and (f) h-

798

BN

799 800 801 802 803 804 805 806

ACS Paragon Plus Environment

Page 36 of 43

Page 37 of 43

Environmental Science & Technology

807 808

Figure 2. Electrophoretic mobility (EPM) values of aqueous MoS2, WS2, and h-BN suspensions

809

as a function of (a) monovalent NaCl and (b) divalent CaCl2 electrolytes. Measurements were

810

carried out at pH 4.6 to 4.9 with no additional buffers and at 25 ºC.

811 812 813

ACS Paragon Plus Environment

Environmental Science & Technology

814 815

Figure 3. Stability plots of aqueous MoS2, WS2, and h-BN suspension as a function of (a)

816

monovalent NaCl and (b) divalent CaCl2 electrolytes. Measurements were carried out at pH 4.6

817

to 4.9 with no additional buffers and at 25 ºC.

818 819

ACS Paragon Plus Environment

Page 38 of 43

Page 39 of 43

Environmental Science & Technology

820 821

Figure 4. vdW energy profiles of bilayer (a) MoS2 and (b) WS2, and (c) h-BN structures as a

822

function of layer separation distance calculated by summing of all pair-wise inter-atomic

823

interactions between layers using molecular modeling. Inset shows difference in energies

824

between MoS2 and WS2.

825 826 827 828 829 830 831 832 833 834 835

ACS Paragon Plus Environment

Environmental Science & Technology

836

837 838

Figure 5. DLVO interaction energy between two 2D NM (MoS2, WS2 or h-BN) sheet in

839

presence of various concentrations of NaCl and CaCl2.

ACS Paragon Plus Environment

Page 40 of 43

Page 41 of 43

Environmental Science & Technology

840 841

Figure 6. Aggregation rates of MoS2, WS2, and h-BN without and with 2.025 mg/L TOC

842

SRNOM and in presence of (a) 7 mM NaCl and 1 mM CaCl2, (b) 100 mM NaCl.

843 844 845 846 847 848 849 850 851 852 853 854 855

ACS Paragon Plus Environment

Environmental Science & Technology

Page 42 of 43

856 857

Table 1. Experimentally obtained and modified DLVO theory predicted CCC values of MoS2,

858

WS2, and h-BN nanosheets.

2D NM MoS2 WS2 h-BN

Experimentally obtained CCC values

Modified DLVO theory predicted CCC values

mM NaCl

mM CaCl2

mM NaCl

mM CaCl2

37 60 19

3 7.2 1.3

55 85 30

5.5 10 5.5

859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878

ACS Paragon Plus Environment

Page 43 of 43

879

Environmental Science & Technology

For Table of Contents Use Only

880

ACS Paragon Plus Environment