Aqueous Aggregation and Surface Deposition Processes of

All DLS and QCM-D measurements were conducted at pH 7.2 ± 0.2 buffered by .... in the presence of 0, 300, 500, and 800 mM NaCl onto a silica sensor s...
1 downloads 0 Views 721KB Size
Article

Aqueous Aggregation and Surface Deposition Processes of Engineered Superparamagnetic Iron Oxide Nanoparticles for Environmental Applications Wenlu Li, Di Liu, Jiewei Wu, Changwoo Kim, and John Dyer Fortner Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 15 Sep 2014 Downloaded from http://pubs.acs.org on September 29, 2014

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

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

Page 1 of 32

Environmental Science & Technology

1

Aqueous Aggregation and Surface Deposition

2

Processes of Engineered Superparamagnetic Iron

3

Oxide Nanoparticles for Environmental Applications

4

Wenlu Li, Di Liu, Jiewei Wu, Changwoo Kim, and John D. Fortner*

5

Department of Energy, Environmental, and Chemical Engineering

6

Washington University in St. Louis

7

St. Louis, Missouri 63130, USA

8 9 10

Submitted to

11

Environmental Science and Technology

12

July 31th, 2014

13 14 15

*To whom correspondence should be addressed:

16

John D. Fortner: Tel: +1-314-935-9293; Fax: +1-314-935-5464; Email: [email protected]

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 32

17

ABSTRACT

18

Engineered, superparamagnetic, iron oxide nanoparticles (IONPs) have significant potential as

19

platform materials for environmental sensing, imaging and remediation due to their unique size,

20

physicochemical and magnetic properties.

21

chemistry of the materials is crucial for such applications in the aqueous phase, and in particular,

22

for porous matrixes with particle-surface interaction considerations.

23

superparamagnetic, highly monodispersed 8 nm IONPs were synthesized and transferred into

24

water as stable suspensions (remaining monodispersed) by way of an interfacial oleic acid

25

bilayer surface.

26

interactions (deposition and release) were quantitatively investigated and described

27

systematically as a function of ionic strength (IS) and type with time-resolved dynamic light

28

scattering (DLS), zeta potential, and real-time quartz crystal microbalance with dissipation

29

monitoring (QCM-D) measurements. The critical coagulation concentration (CCC) for oleic

30

acid bilayer coated iron oxide nanoparticles (OA-IONPs) were determined to be 710 mM for

31

NaCl (matching DLVO predictions) and 10.6 mM for CaCl2, respectively. For all conditions

32

tested, surface deposition kinetics showed stronger, more favorable interactions between OA-

33

IONPs and polystyrene surfaces compared to silica, which is hypothesized to be due to increased

34

particle-surface hydrophobic interactions (when compared to silica surfaces).

To this end, controlling the size and surface

In this study,

Once stabilized and characterized, particle-particle and model surface

ACS Paragon Plus Environment

2

Page 3 of 32

Environmental Science & Technology

35

KEYWORDS

36

Iron oxide nanoparticles (IONPs), nanoparticle aggregation, nanoparticle deposition,

37

nanoparticle stability, environmental sensing, hydrophobic interactions, Quartz Crystal

38

Microbalance with Dissipation (QCM-D)

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 32

39

INTRODUCTION

40

The vision of applying finely divided materials, or even nanoscale materials for simultaneous

41

sensing, mapping and remediating of targeted, complicated environments has recently gained

42

increasing attention, concurrent with advances in related material sciences.1-4 In particular,

43

magnetic based nanoparticles (magnetite and maghemite) provide unique advantages to such

44

applications due to their size (2 to 100 nm) and tunable physicochemical properties. To this

45

point, engineered magnetite (Fe3O4) based nanostructures, differ substantially from it’s bulk

46

phase counterpart due to greatly enhanced specific surface area, reactivity, and magnetic

47

response (e.g., superparamagnetic properties are observed for single domain particles within a

48

size range of ca. 6-20 nm compared to bulk phase ferromagnetism).5 Such properties underpin

49

this class of materials’ potential in therapeutic, biological and environmental imaging,6-10

50

remediation11-15 and sensing technologies.1, 2, 4, 16-19

51

Ultimately though, for many of these applications to be successful, engineered materials such

52

as these must not only process intrinsic, novel functionality but must also allow for controlled

53

physical placement (e.g., within a desired volume or at a specific target interface). As an

54

example, single domain, superparamagnetic iron oxide (magnetite) based nanoparticles have

55

significant potential as T2 type nuclear magnetic resonance (NMR) imaging contrast agents7, 20

56

and are being explored for enhanced environmental (subsurface) imaging.4,

57

technologies are aimed at extending and refining hydrocarbon detection in oil-field rocks and

58

even for NAPL (nonaqueous phase liquids) sensing (and remediation) in contaminated systems.2,

59

3, 23

60

nanoscale materials can be transported through low porosity media,2, 22, 24 potentially reaching

61

non-aqueous liquid phases/interface targets (e.g., oil-water-interfaces), at which point they can

10, 21, 22

Such

Further, based on material size (< 20 nm) and dispersivity, it is possible that monodispersed

ACS Paragon Plus Environment

4

Page 5 of 32

Environmental Science & Technology

62

be responsive and potentially even functional.22-24 However, to date, the bottleneck for such

63

applications is largely due to particle (in)stability challenges.10, 24-26 As the aggregation state

64

changes (based on the surface energy of suspended nanoparticles), transportation of

65

nanoparticles through low porous media is increasingly limited and final deposition ‘misses the

66

mark’. Normally, the stability of the nanoparticles is controlled by the surface chemistries

67

involved27 and in particular, the coatings and charge of the particle surface, the matrix surface

68

chemistries, the ionic strength or salinity of the water, and the interactions with other substances,

69

such as natural organic matter (NOM).28 Therefore, quantitatively understanding the aggregation

70

and deposition/adsorption behavior of (surface)stabilized iron oxide nanoparticles (IONPs) under

71

different water chemistries is crucial for use in any potential in-situ, subsurface technologies.

72

The development of low defect, single domain nanoscale IONPs has drawn considerable

73

attention from a number of research communities over the past two decades.5, 7, 29-31 In contrast

74

to aqueous preparation routes, high temperature organic solution phase methods, in particular,

75

allow for the preparation and control of highly uniformed nanoparticles due to the successful

76

separation of nucleation and the growth phases during synthesis.5 For this work, monodisperse

77

IONPs with controlled sizes were synthesized by decomposition of iron carboxylate salts in

78

high-temperature boiling solvent.29,

79

organic solvent, a favorable stabilizing surface is required for aqueous suspensions, and thus

80

applications. Here, we use a stable, oleic acid bilayer surface coating allowing for the efficient

81

transfer33 of the nanoparticles from nonpolar solvent to water, forming stable, nonaggregating,

82

monodispersed suspensions.

83

available (oleic acid is a major component in a number of natural oils including olive oil and

84

magnetite is a widely found natural mineral), such a “green” core-shell structure may be

32

As the aforementioned materials are synthesized in

Further, as both materials are widely naturally occurring and

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 32

85

desirable for actual environmental applications.33

86

superparamagnetic properties29 of the iron oxide core does not provide attractive magnetic forces

87

between nanoparticles, unlike ferromagnetic particles, unless an external magnetic field is

88

applied, thus simplifying aggregation observations and discussion.

Lastly, within this size range, the

89

To the best of our knowledge, this is the first study systematically quantifying the aggregation

90

kinetics of engineered superparamagnetic IONPs and their deposition behaviors onto different

91

model environmental surfaces in the presence of sodium and calcium cations. IONP suspensions

92

were developed, synthesized and characterized with size of 8 nm as stable, monodispersed

93

aqueous suspensions by way of an oleic acid surface bilayer. Particle-particle and model surface

94

interactions were quantitatively investigated and described as a function of ionic strength/type

95

with time resolved dynamic light scattering (DLS), zeta (ζ) potential and real-time quartz crystal

96

microbalance with dissipation monitoring (QCM-D) measurements (surface deposition and

97

release), using two model environmental surfaces (hydrophilic vs. hydrophobic).

98

MATERIALS AND METHODS

99

Materials: Iron(III) oxide (hydrated, catalyst grade), 1-octadecene (technical grade, 90%),

100

oleic acid (technical grade, 90%), sodium chloride (ACS reagent, ≥99.0%), calcium chloride

101

dihydrate (ACS reagent, ≥99%), sodium hydroxide (ACS reagent, ≥97.0%) , sodium bicarbonate

102

(ACS reagent, 99.7-100.3%) and nitric acid (trace metal grade) were all purchased from Sigma-

103

Aldrich. All materials were used without any further purification.

104

Preparation of Oleic Acid Bilayer Coated Iron Oxide Nanoparticles (OA-IONPs):

105

Detailed synthesis of IONPs and phase transfer method for OA-IONPs are given in the

106

Supporting Information (SI). The concentration of the water stable OA-IONPs stock suspension

ACS Paragon Plus Environment

6

Page 7 of 32

Environmental Science & Technology

107

was 50 mg/L as determined by inductively coupled plasma mass spectrometry (ICP-MS, Agilent

108

7500ce).

109

Solution Chemistry: Various concentrations of NaCl and CaCl2 electrolyte stock solutions

110

were prepared and filtrated (pore size of 0.2 um, Millipore) before use. All DLS and QCM-D

111

measurements were conducted at pH 7.2 ± 0.2 buffered by 0.15 mM NaHCO3 and at room

112

temperature (22 °C).

113

Dynamic Light Scattering: The DLS measurements were performed on a Zetasizer (Nano

114

ZS, Malvern). The OA-IONPs stock solution was diluted 50 times to 1 mg/L for the DLS

115

experiments and pH was adjusted to 7.2. For each measurement, a predetermined volume of

116

diluted nanoparticle suspension was added into a vial. After that, a certain amount of electrolyte

117

stock solution was injected into the vial to make the total volume of sample to be 1 mL. After a

118

short time of vortex (1.5 s), samples were quickly put into the DLS chamber and measured

119

immediately. The scattered light intensity was detected by a photodetector at a scattering angle

120

of 173°. Data points were measured every 15 s and recorded continuously. DLS samples were

121

left to aggregate between 20 to 60 min, depending on the aggregation rate of each sample. The

122

attachment efficiency (α) of the OA-IONP aggregates in the presence of various electrolyte

123

concentrations was calculated by the following equation34

124

1

α= W = k

k fast

(1)

125

where k is the initial aggregation rate constant at different salt concentrations and kfast is the

126

aggregation rate constant under diffusion-limited (fast) aggregation conditions.

127

Quartz Crystal Microbalance with Dissipation: QCM-D measurements were performed

128

with a Q-Sense E4 (Q-sense AB, Sweden) unit by simultaneously monitoring the changes in

129

frequency (∆f) and energy dissipation (∆D) of a 5 MHz silica (SiO2) coated QCM-D crystal

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 32

130

(QSX-303, Q-sense) and a 5 MHz polystyrene (PS) coated QCM-D crystal (QSX-305, Q-sense).

131

Protocols used to clean the sensors and detailed deposition experiment conditions are described

132

in the Supporting Information (SI). All deposition measurements were allowed to proceed for a

133

period between 10 to 30 min to determine the nanoparticle deposition rates. As deposition of

134

OA-IONPs occurred onto the silica or polystyrene sensor surfaces, the increase in the mass of the

135

crystal induced a continuous shift in the frequency (f), as described by the Sauerbrey

136

relationship.35

∆ = −  ∆

137

(2)

138

where n is the overtone number (1, 3, 5, 7, 9, etc…) and C is the crystal constant (17.7 ng/(Hz

139

cm2)). The deposition process on the crystal surface will also enhance the crystal’s ability to

140

dissipate energy, as shown in the increase in the energy dissipation (D) via the following

141

equation: E

dissipated D= 2πE

142

stored

(3)

143

where Edissipated is the energy loss in one oscillation cycle and Estored is the total energy stored in

144

the oscillator. Since deposited mass is linear to the frequency shift, the nanoparticle deposition

145

rate can be calculated from the initial slope of frequency change.

146

normalized frequency at the third overtone in a given period (t) was used to quantify the

147

deposition rate (rd):34

148 149

d =

d∆ d

The initial change of

(4)

In order to obtain the deposition attachment efficiency (αD) as shown in equation 5,

150

favorable deposition conditions were conducted by coating the silica sensor with a positively

151

charged poly-L-lysine (PLL) layer as described elsewhere:36

ACS Paragon Plus Environment

8

Page 9 of 32

Environmental Science & Technology

 = "

152

∆ ⁄  ∆ ⁄  #fav

(5)

153

RESULTS AND DISCUSSION

154

IONPs Stabilization and Characterization

155

IONPs of 8 nm were synthesized and used specifically for this study as shown in TEM

156

micrograph in Figure 1a. These nanoparticles are highly uniform in size, with size distribution

157

variability well under 10 %; the size distribution histograms are shown in Figure S1a. The HR-

158

TEM image (SI Figure S1b) of 8 nm IONPs shows that they are highly crystallized and the

159

lattice fringe was indexed to (440) plane of the inverse spinel iron oxide structure.29, 37 Once

160

synthesized, all nanoparticles used for this work were capped with an oleic acid layer with

161

hydrophobic tail pointing outside, rendering them stable in the non-polar solvent (in this case,

162

hexanes). A water stabilizing, surface bilayer was achieved by mixing a small amount of

163

additional oleic acid with water and the iron oxide suspension in hexane (creating a liquid-liquid

164

system) and then using a sonication probe to facilitate phase transfer.38-40 Corresponding TEM

165

micrographs (Figure 1b) along with DLS measurements discussed below, indicate aqueous

166

transferred IONPs remain monodisperse with no core size change. The stable, homogenous

167

bilayer, as described by others, is formed when the hydrophobic oleic acid tail of the second

168

layer aligns with the original hydrophobic tail of the first layer (via hydrophobic and dispersive

169

Van der Waals forces),41 and the hydrophilic head group (carboxylic acid functionality) of the

170

second layer aligns at the aqueous interface, rendering the nanoparticles water stable.33 The

171

number-weighted hydrodynamic diameter of the OA-IONPs was measured to be 15.6 ± 1.3 nm

172

for 8 nm IONPs originally dispersed in hexane (SI Figure S2). The DLS value for 8 nm IONPs

173

is consistent with the value reported before (14.2 ± 2.6 nm) which takes into account the

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 32

174

thickness of the bilayer.33 For all aqueous suspensions, no significant change in hydrodynamic

175

diameter was observed after 3 months and the OA-IONPs did not precipitate over 6 months.

176

Figure 2 shows ζ-potential measurements of 8 nm OA-IONPs under various NaCl and CaCl2

177

electrolyte concentrations. Due to the dissociation of carboxylate groups at pH 7.2, OA-IONPs

178

exhibit negatively charged surfaces (OA pKa = 4.95).42 ζ-potentials of OA-IONPs were negative

179

for the entire range of NaCl and CaCl2 concentrations examined in this study.

180

conditions, as expected, the nanoparticle ζ-potentials become less negative with increasing ionic

181

strength due to the basic charge screening effect.43 ζ-potential values of OA-IONPs became less

182

negative from -53.4 mV to -46.5 mV as NaCl concentration increased from 10 to 300 mM. As

183

low as 0.05 mM of CaCl2 significantly increased the ζ-potential values of OA-IONPs to -30.7

184

mV, indicating the enhanced effect of charge neutralization from divalent cations, as observed in

185

other studies.34, 43, 44

Under all

186

Aggregation Kinetics

187

Extensive studies, focused mostly on carbon based material and metal/metal oxides, have been

188

conducted to evaluate the aqueous aggregation behavior(s) of a number of engineered

189

nanomaterials.27 By monitoring effective particle size change via dynamic light scattering (DLS),

190

aggregation kinetics (as a function of one variable) can be determined. For most studies, such

191

aggregation behaviors of engineered materials can be described by Derjaguin-Landau-Verwey-

192

Overbeek (DLVO) theory27 and exhibit reaction-limited (slow) and diffusion-limited (fast)

193

regimes.45 Further, previous reports primarily focus on the effects of Na+ and Ca2+, as sodium

194

and calcium are the most abundant cations in most natural aquifers and even deep saline water

195

resevoirs.46

ACS Paragon Plus Environment

10

Page 11 of 32

Environmental Science & Technology

196

Here, the aggregation of OA-IONPs was also studied in the presence of NaCl and CaCl2 at pH

197

7.2. The aggregation rate of OA-IONPs increased with increasing NaCl concentrations below

198

700 mM (SI Figure S3a). This is consistent with zeta potential measurements as an increase in

199

NaCl concentration, nanoparticle zeta potential became less negative (Figure 2). However, at

200

higher NaCl concentrations (800 and 1000 mM), the increase in electrolyte concentration has a

201

negligible effect on the aggregation rate. Similarly, the aggregation kinetics of OA-IONPs in the

202

presence of CaCl2 follows the same trend (SI Figure S3b), but at lower concentrations as

203

expected with divalent Ca2+.34, 44, 47 Figure 3 shows the aggregation attachment efficiencies of

204

OA-IONPs as a function of both electrolyte concentrations. For these systems, aggregation

205

kinetics can also be classically divided into two regimes: reaction-limited (slow) and diffusion-

206

limited (fast) aggregation.

207

concentration screens the surface charge, thus effectively reducing the energy barriers to

208

aggregation (SI and SI Figure S4).27,

209

However, when the electrolyte concentrations increase to the critical coagulation concentration,

210

the energy barrier is completely eliminated, leading to diffusion-controlled aggregation. Figure

211

S5 shows good agreement between prediction of DLVO theory (calculated) and experimental

212

data. The critical coagulation concentrations (CCC) for OA-IONPs are 710 mM for NaCl and

213

10.6 mM for CaCl2. Tombácz et al. reported a CCC value of 500 mM NaCl for poly(acrylic acid)

214

– PAA coated magnetite nanoparticles with primary size under 10 nm at pH 6.5.48 Considering

215

the different preparation methods for IONPs and surface modifications, our CCC value for OA-

216

IONPs is comparable. The ratio of CCC values calculated (10.6/710) is proportional to z-6.07,

217

where z is the valence of the calcium counterion (i.e. z=2). This is in good agreement with the

In the reaction-limited regime, an increase in the electrolyte

34

As a result, the aggregation sizes increase rapidly.

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 32

218

empirical Schulze-Hardy rule that the CCC ratio for CaCl2 and NaCl should be z-6 for particles

219

with large ζ-potentials.49

220

Deposition Kinetics

221

Quartz crystal microbalance with dissipation (QCM-D) enables the real-time, high sensitivity

222

surface (interaction) analysis (here as particle deposition and release) for fluid suspended

223

particles and even macromolecules.50, 51 Recently, QCM-D has proven to be useful and powerful

224

in evaluating solution based interfacial behaviors of engineered nanomaterials.27,

225

Representative studies of note include: Chen et al. who examined the deposition of fullerene

226

nanoparticles onto silica surfaces and found a decreased deposition rate at higher electrolyte

227

concentrations;34 Fatisson et al.52 employed this technique to compare the deposition behavior of

228

bare and CMC-modified nanoscale zerovalent iron (nZVI) particles onto silica surfaces over a

229

wide range of solution chemistries; Saleh et al. found the lowest deposition rates from the

230

triblock copolymer coated nanoscale zerovalent iron (nZVI) compared to bare and other

231

surfactant-modified nZVI from QCM-D experiment.53 Further, QCM-D based analyses can be

232

tailored as quartz crystal (sensor) surfaces can be modified via a number of methods to mimic

233

particle or environmental surfaces36 for enhancing the technique’s applicability. As done in this

234

work, interfacial surface interactions can be quantitatively investigated with QCM-D using both

235

(relatively) hydrophilic (as silica) and (relatively) hydrophobic (as polystyrene)54-57 sensor

236

surfaces.

36, 52

237

A representative normalized frequency shift at the third overtone ( ∆"%# ) for OA-IONPs

238

deposition onto a polystyrene surface is shown in Figure S6. The deposition kinetics for OA-

239

IONPs were studied as a function of nanoparticle concentration ranging from 2.5 mg/L to 40

240

mg/L (measured as Fe). In order to have a sufficient frequency shift for all deposition studies,

ACS Paragon Plus Environment

12

Page 13 of 32

Environmental Science & Technology

241

we chose 7.5 mg/L as the concentration for all QCM-D measurements, above this concentration

242

(40 mg/L maximum) no change in deposition kinetics was observed. Figure S7 presents a

243

typical deposition study of 8 nm OA-IONPs in the presence of 0, 300, 500, and 800 mM NaCl

244

onto a silica sensor surface. To begin, there was no deposition of OA-IONPs onto silica sensor

245

surface in the absence of salt.

246

electrostatic interactions between OA-IONPs and the silica surface.

247

silica58 are both negatively charged pH 7.2, the strong repulsive force simply inhibits deposition.

248

When OA-IONPs do interact with the sensor surface, the decrease in the frequency shift

249

indicates the continuous deposition of nanoparticles and as nonlinear deposition behavior34 is

250

observed at high ionic strength (e.g., 800 mM NaCl on silica), where the initial slope is used to

251

calculate the unsaturated (surface) deposition rate. The deposition behaviors of 8 nm OA-IONPs

252

onto silica and polystyrene sensor surfaces were examined over a wide range of NaCl and CaCl2

253

concentrations, as shown in Figure 4. As electrolyte concentrations increased, deposition rates

254

correspondingly increased as the electric double layers were compressed by the charge screening

255

from the additional salt. Further increase in electrolyte concentrations close to or higher than the

256

CCC values, resulted in diffusion limited transport, and the deposition rate then decreased as

257

expected and observed by others.34 From Figure 4a, nanoparticles reach a maximum deposition

258

rate on silica at around 500 mM NaCl, while on the more hydrophobic polystyrene surface, the

259

fastest deposition rate was observed for NaCl concentrations between 200-400 mM.

260

separate peaks in the deposition rates of silica and polystyrene surfaces indicate the higher

261

affinity of (this particular) OA-IONPs deposition onto polystyrene, a relatively more

262

hydrophobic surface. In addition, the maximum deposition rate on polystyrene is around 6.5

263

Hz/min, which is significantly higher than the deposition rate on silica surface (2.5 Hz/min).

This observation can be explained based on the repulsive Since OA-IONPs and

The

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 32

264

From Figure 4b, nanoparticles reach a maximum deposition rate on silica around 5 to 7.5 mM

265

CaCl2, while on the polystyrene surface, the fastest deposition concentration for CaCl2 is

266

between 2-3 mM. The concentrations at which the nanoparticles reach highest deposition rate in

267

the presence of NaCl are significantly higher than that in the presence of CaCl2. In addition to

268

the similar charge screening effect as Na+, Ca2+ can also form complexes with the carboxyl

269

groups and thus further neutralize their surface charge.34 Furthermore, carboxyl groups do not

270

form inner sphere complexes with Na+ (i.e. Na+ can only screen the charges of nanoparticles).59

271

For these reasons, Ca2+ ions are significantly more effective in reducing the energy barrier to

272

OA-IONPs deposition than Na+, which is in line with observations made by others.34, 43, 53, 60

273

Favorable deposition onto PLL coated silica sensor surfaces was also conducted under the

274

same conditions as for the unfavorable conditions.27, 34 Insets from Figure 4 show the deposition

275

rates (frequency shift rates) of OA-IONPs in the presence of NaCl and CaCl2 onto PLL coated

276

sensor surfaces at pH 7.2. Generally, on PLL coated sensor surfaces, deposition rates decreased

277

with increasing ionic strength in both NaCl and CaCl2 solutions.

278

explained due to the significant formation of large aggregates36,

279

concentrations as shown for previous DLS measurements (SI Figure S3). By normalizing the

280

particle deposition rates at unfavorable conditions to the corresponding particle deposition rates

281

at favorable conditions, deposition attachment efficiencies (αD) of OA-IONPs are shown in

282

Figure 5. For deposition onto silica sensor surfaces, αD increased with the increasing IS until the

283

energy barrier is completely eliminated, resulting in αD of 1 (SI Figure S8). The critical

284

deposition concentration36 (CDC) was determined to be 485 mM for NaCl and 5.1 mM for CaCl2.

285

While the deposition kinetics on silica surfaces clearly shows typical unfavorable and favorable

286

regimes, deposition kinetics on polystyrene surfaces do not follow classical deposition behaviors

The observation can be

43, 58

at elevated electrolyte

ACS Paragon Plus Environment

14

Page 15 of 32

Environmental Science & Technology

287

of engineered particles,27 suggesting that hydrophobic interactions mainly influence the

288

attachment of engineered OA-IONPs onto polystyrene surfaces.61 Xiao and Wiesner found that

289

aqu-nC60, a moderately hydrophobic nanoparticle, has the highest retention on the most

290

hydrophobic BSA-coated glass beads, indicating the hydrophobic interaction contributes to the

291

particle attachment.62 A recent study by Song et al.61 suggested hydrophobic interactions are

292

responsible for the increased deposition onto hydrophobic surfaces with increasing

293

hydrophobicity of Ag nanoparticles. As oleic acid bilayer coating is actually amphiphilic,33 we

294

propose that the increased deposition rate onto polystyrene is due to the additional hydrophobic

295

interactions55 between the hydrophobic portion of the coating with the polystyrene sensor surface

296

(which also has less surface charge than silica).56 However, the exact mechanisms behind the

297

hydrophobic interactions (nanoparticle-surface) are still not fully understood.63-65 The relative

298

higher deposition attachment efficiencies of engineered OA-IONPs onto the polystyrene surface

299

indicate that the transport and fate of these nanoparticles will be greatly influenced by the

300

hydrophobic surfaces in the environment (e.g. oil, NAPL or natural organic matter). In another

301

word, if transportation of these engineered IONPs in the subsurface could reach to a target

302

volume, favorable in-situ partitioning at oil / NAPL is likely.

303

Deposited Layer Stiffness and Stability

304

Along with measuring frequency changes, thus mass deposition changes, QCM-D can also

305

measure the change in dissipated energy during and after particle deposition events.60 The slope

306

of D to f (&∆'"%# ⁄∆"%# &) can be used as an estimation of the induced energy dissipation per

307

coupled mass change – which is a measure of particle size and particle-surface contact stiffness

308

for films of discrete nanoparticles.55, 66-68 Here, the slope of D to f exhibited a linear relationship

309

(R2 ≥ 0.99, SI Figure S9) for most cases except at very low ionic strength due to negligible

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 32

Therefore, &∆'"%# ⁄∆"%# & was calculated for all the

310

deposition mass on the sensor surface.

311

deposition experiments except for deposition at very low electrolyte concentrations. The trends

312

of &∆'"%# ⁄∆"%# & as a function of NaCl and CaCl2 are presented in Figure 6. For nanoparticle

313

deposition onto the silica surface, &∆'"%# ⁄∆"%# & increased with increasing electrolyte

314

concentration for both NaCl and CaCl2. The interactions of aggregated nanoparticles with the

315

silica surface become increasingly dissipative and therefore, &∆'"%# ⁄∆"%# & values increase – in

316

other words, the layer becomes more loosely attached to silica with increasing ionic strength due

317

to formation of large aggregates.69

318

coated surfaces are lower than the values on silica surfaces. Interestingly, the &∆'"%# ⁄∆"%# & of

319

nanoparticle deposition on the polystyrene surface is much lower than for a silica surface in the

320

presence of NaCl and CaCl2, indicating the nanoparticles were more rigidly attached onto the

321

polystyrene surface as compared to the silica surface. This phenomenon is hypothesized to be in

322

part, due to stronger hydrophobic interactions,55,

323

attachment of aggregated nanoparticles onto the polystyrene surface.

324

electrolyte concentrations and types, the &∆'"%# ⁄∆"%# & value on the polystyrene surface remains

325

constant, which suggests the stiffness of the nanoparticles’ attachment is independent to the

326

addition of salt and surface chemistry. These observations are in good agreement with Chang et

327

al.’s study which highlights the role of sensor surface chemistry with regard to nanoparticles

328

deposition.55

Under non-repulsive conditions, &∆'"%# ⁄∆"%# & on PLL

61

as discussed above, leading to the rigid Moreover, under all

329

To further evaluate the properties of the sorbed layer, experiments were designed to measure

330

the effective particle release after favorable deposition60 had occurred. OA-IONPs were first

331

deposited onto the silica surface at favorable deposition rate conditions (500 mM NaCl and 7.5

332

mM CaCl2, respectively) and then the (flowing) aqueous solution chemistry over the deposited

ACS Paragon Plus Environment

16

Page 17 of 32

Environmental Science & Technology

333

layer was changed (termed the rinse solution) to investigate the stability of the sorbed layer and

334

release of nanoparticles back into solution. For silica surfaces (Figure 7a), the rinse solutions

335

used here include the corresponding buffer, 1 mM NaCl, water buffered at pH 7.2 and water

336

buffered at pH 12. A partial increase of frequency is observed in Figure 7a where the rinse

337

solution is changed from 500 mM NaCl (step D) to 1 mM NaCl (step E), indicating release of

338

sorbed particles when the solution chemistry changed. However, with the initial 7.5 mM CaCl2

339

based suspension, no obvious frequency change is observed (less than 2 % of mass change

340

calculated with Sauerbrey model, Figure 7b and SI Table S1), thus indicating a stable deposition

341

layer regardless of the changes in water chemistry. As discussed, previous studies have shown

342

that Ca2+ can bind to two adjacent carboxyl groups, acting as a bridge, which leads to a more

343

stable deposited layer.43, 70, 71 The negative frequency shift in step G was due to the increase in

344

viscosity/density in water solution buffered at high pH.34 These results indicate that the sorbed

345

layer properties can also be functions of both ionic type71 and surface, and that the initial

346

conditions are critical, as we observed the majority of sorbed layers, for all cases, remain (mass

347

retained) even after the rinse solution changed back to MilliQ water. Such observations are

348

important when considering particle transport over long distances or times as heterogeneities

349

with regard to local ionic strength/type and surfaces may act as effective surface (particle) sinks.

350

On the other hand, if a stable, rigid (deposition) layer is desirable, divalent or even trivalent

351

cations may be useful in application.

352

In summary, this work fundamentally describes the role of two typical cations, Na+ and Ca2+,

353

on the aggregation, deposition and corresponding release of engineered iron oxide nanoparticles

354

(IONPs) with two model surface types (hydrophilic vs. hydrophobic).

355

quantitatively describe and discern the effects of different cations and ionic strength, which

Measurements

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 32

356

strongly affect the stability of nanoparticles and thus their function with regard to particle

357

deposition and release onto/from different, model environmental surfaces. Data sets indicate

358

these OA-IONPs will be relatively stable in groundwater conditions as the Na+ is typically 1-10

359

mM and Ca2+ is typically 0.1-2 mM.53 However, based on this and other studies, it is still

360

difficult to control the stability and deposition of engineered IONPs under extreme conditions,

361

particularly at high salinity (IS > 1 M) as found, for example in brine aquifers or some petroleum

362

reservoirs.46 Additionally, sensor surface (coatings) also proved to be critical factors governing

363

the deposition rates of these nanoparticles onto environmental surfaces.

364

observed a greater affinity, in all cases, for these nanoparticles to associate with the more

365

hydrophobic surface and that oleic acid coatings are quite sensitive to the presence of calcium.

366

In particular, we

To conclude, aqueous application (e.g. sensing, remediation, etc.) of highly controlled,

367

magnetic engineered nanoparticles remains relatively nascent.

This work provides new

368

fundamental knowledge and understanding for such applications and beyond, while underscoring

369

complexities associated with nanoparticle behavior in multiphase systems (here as water-solid

370

interfaces). The authors certainly recognize that natural environments will be more complicated

371

than systems presented, based on the presence of NOM, redox reactions (resulting in particle

372

aging), (bio) molecules, among others, which all add to the uncertainty of extrapolation with

373

regard to real world understanding and accurate prediction of aggregation and deposition

374

processes.27

375

specifically discerning the role of environmental aging and the presence of NOM for similar

376

IONPs aggregation and deposition behaviors.

Accordingly, ongoing research is being conducted to investigate such factors,

ACS Paragon Plus Environment

18

Page 19 of 32

Environmental Science & Technology

377 378

Figure 1. TEM micrographs of the as-prepared 8 nm IONPs in (a) hexane and (b) water. All

379

scale bars are 50 nm.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 32

380 381

Figure 2. Zeta potentials of OA-IONPs over a range of NaCl and CaCl2 concentrations under pH

382

7.2. Each data point shows the mean of 30 measurements of triplicate samples, where the error

383

bars represent standard deviations.

ACS Paragon Plus Environment

20

Page 21 of 32

Environmental Science & Technology

384 385

Figure 3. Attachment efficiency (α) of OA-IONPs as functions of (a) NaCl and (b) CaCl2

386

concentrations at pH 7.2.

387

measurements is 710 mM NaCl and 10.6 mM CaCl2 for OA-IONPs. The lines are extrapolated

388

from the reaction-limited and diffusion-limited regimes, and the intersections show the

389

respective CCC values.

The critical coagulation concentration (CCC) based on the

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 32

390 391

Figure 4. Deposition rates (frequency shift rates) of OA-IONPs in the presence of (a) NaCl and

392

(b) CaCl2 onto silica and polystyrene sensor surfaces. Insets: deposition rates under favorable

393

conditions. Each data point shows the average of duplicate or triplicate measurements, where the

394

error bars represent standard deviations.

ACS Paragon Plus Environment

22

Page 23 of 32

Environmental Science & Technology

395 396

Figure 5. Deposition attachment efficiency (αD) of OA-IONPs in the presence of (a) NaCl and (b)

397

CaCl2 onto silica and polystyrene sensor surfaces. The lines are extrapolated from the

398

unfavorable and favorable deposition regimes, and the intersections show the respective critical

399

deposition concentration (CDC) values.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 32

400 401

Figure 6. &∆'"%# ⁄∆"%# & of OA-IONPs deposition in the presence of NaCl and CaCl2 onto PLL,

402

silica and polystyrene surfaces.

ACS Paragon Plus Environment

24

Page 25 of 32

Environmental Science & Technology

403 404

Figure 7. Stability of deposited OA-IONPs layer over different rinse solutions when the layer

405

was formed in the presence of (a) 500 mM NaCl and (b) 7.5 mM CaCl2. Baseline was first

406

established in MilliQ water (A) and corresponding electrolytes (B), before the nanoparticles are

407

deposited onto the silica surface (C). The system is then rinsed in the respective electrolytes (D),

408

1 mM NaCl at pH 7.2 (E), MilliQ water at pH 7.2 (F), and finally, MilliQ water with pH adjusted

409

to 12 (G).

ACS Paragon Plus Environment

25

Environmental Science & Technology

410

Page 26 of 32

ASSOCIATED CONTENT

411

Supporting Information. Detailed synthesis and phase transfer of iron oxide nanoparticles,

412

QCM-D sensor cleaning protocols, QCM-D deposition experiment, DLVO calculations, nine

413

figures (nanoparticle size distribution, DLS size of nanoparticles, aggregation profiles, DLVO

414

interaction energy profiles, DLVO predictions, QCM-D deposition experiments, and

415

representative D/f ratio) and one table are available free of charge via the Internet at

416

http://pubs.acs.org.

417

AUTHOR INFORMATION

418

Corresponding Author

419

*To whom correspondence should be addressed:

420

John D. Fortner: Tel: +1-314-935-9293; Fax: +1-314-935-5464; Email: [email protected]

421

Funding Sources

422

The authors would like to thank the American Chemical Society’s Petroleum Research Fund

423

(#52640-DNI10) and the National Science Foundation (CBET, #1236653) for supporting this

424

work.

425

ACKNOWLEDGMENT

426

This work was supported by American Chemical Society’s Petroleum Research Fund and the

427

National Science Foundation. TEM, DLS facilities were provided by the Nano Research Facility

428

(NRF) at Washington University in St. Louis, a member of the National Nanotechnology

429

Infrastructure Network (NNIN), which is supported by the National Science Foundation under

430

Grant No. ECS-0335765.). We also thank Mr. Carl H. Hinton for his contributions to this work.

ACS Paragon Plus Environment

26

Page 27 of 32

Environmental Science & Technology

431

REFERENCES

432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475

1. Hsu, R.-S.; Chang, W.-H.; Lin, J.-J., Nanohybrids of Magnetic Iron-Oxide Particles in Hydrophobic Organoclays for Oil Recovery. ACS Applied Materials & Interfaces 2010, 2, (5), 1349-1354. 2. Saleh, N.; Sirk, K.; Liu, Y. Q.; Phenrat, T.; Dufour, B.; Matyjaszewski, K.; Tilton, R. D.; Lowry, G. V., Surface modifications enhance nanoiron transport and NAPL targeting in saturated porous media. Environ. Eng. Sci. 2007, 24, (1), 45-57. 3. Saleh, N.; Phenrat, T.; Sirk, K.; Dufour, B.; Ok, J.; Sarbu, T.; Matyjaszewski, K.; Tilton, R. D.; Lowry, G. V., Adsorbed Triblock Copolymers Deliver Reactive Iron Nanoparticles to the Oil/Water Interface. Nano Letters 2005, 5, (12), 2489-2494. 4. Ingram, D. R.; Kotsmar, C.; Yoon, K. Y.; Shao, S.; Huh, C.; Bryant, S. L.; Milner, T. E.; Johnston, K. P., Superparamagnetic nanoclusters coated with oleic acid bilayers for stabilization of emulsions of water and oil at low concentration. Journal of Colloid and Interface Science 2010, 351, (1), 225-232. 5. Park, J.; An, K.; Hwang, Y.; Park, J.-G.; Noh, H.-J.; Kim, J.-Y.; Park, J.-H.; Hwang, N.M.; Hyeon, T., Ultra-large-scale syntheses of monodisperse nanocrystals. Nat Mater 2004, 3, (12), 891-895. 6. Gupta, A. K.; Gupta, M., Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials 2005, 26, (18), 3995-4021. 7. Na, H. B.; Song, I. C.; Hyeon, T., Inorganic Nanoparticles for MRI Contrast Agents. Advanced Materials 2009, 21, (21), 2133-2148. 8. Cattoz, B.; Cosgrove, T.; Crossman, M.; Prescott, S. W., Surfactant-Mediated Desorption of Polymer from the Nanoparticle Interface. Langmuir 2011, 28, (5), 2485-2492. 9. Hu, F.; Jia, Q.; Li, Y.; Gao, M., Facile synthesis of ultrasmall PEGylated iron oxide nanoparticles for dual-contrast T 1 - and T 2 -weighted magnetic resonance imaging. Nanotechnology 2011, 22, (24), 245604. 10. Bagaria, H. G.; Xue, Z.; Neilson, B. M.; Worthen, A. J.; Yoon, K. Y.; Nayak, S.; Cheng, V.; Lee, J. H.; Bielawski, C. W.; Johnston, K. P., Iron Oxide Nanoparticles Grafted with Sulfonated Copolymers are Stable in Concentrated Brine at Elevated Temperatures and Weakly Adsorb on Silica. ACS Applied Materials & Interfaces 2013, 5, (8), 3329-3339. 11. Fan, Q.-h.; Li, P.; Chen, Y.-f.; Wu, W.-s., Preparation and application of attapulgite/iron oxide magnetic composites for the removal of U(VI) from aqueous solution. Journal of Hazardous Materials 2011, 192, (3), 1851-1859. 12. Yavuz, C.; Mayo, J. T.; Suchecki, C.; Wang, J.; Ellsworth, A.; D’Couto, H.; Quevedo, E.; Prakash, A.; Gonzalez, L.; Nguyen, C.; Kelty, C.; Colvin, V., Pollution magnet: nano-magnetite for arsenic removal from drinking water. Environ Geochem Health 2010, 32, (4), 327-334. 13. Yean, S.; Cong, L.; Yavuz, C. T.; Mayo, J. T.; Yu, W. W.; Kan, A. T.; Colvin, V. L.; Tomson, M. B., Effect of magnetite particle size on adsorption and desorption of arsenite and arsenate. Journal of Materials Research 2005, 20, (12), 3255-3264. 14. Yavuz, C. T.; Mayo, J. T.; Yu, W. W.; Prakash, A.; Falkner, J. C.; Yean, S.; Cong, L. L.; Shipley, H. J.; Kan, A.; Tomson, M.; Natelson, D.; Colvin, V. L., Low-field magnetic separation of monodisperse Fe3O4 nanocrystals. Science 2006, 314, (5801), 964-967. 15. Jiang, Y.; Wang, W.-N.; Biswas, P.; Fortner, J. D., Facile Aerosol Synthesis and Characterization of Ternary Crumpled Graphene–TiO2–Magnetite Nanocomposites for Advanced Water Treatment. ACS Applied Materials & Interfaces 2014, 6, (14), 11766-11774.

ACS Paragon Plus Environment

27

Environmental Science & Technology

476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519

Page 28 of 32

16. Hua, L.; Zhou, J.; Han, H., Direct electrochemiluminescence of CdTe quantum dots based on room temperature ionic liquid film and high sensitivity sensing of gossypol. Electrochimica Acta 2010, 55, (3), 1265-1271. 17. Su, S.; Wu, W.; Lu, J.; Gao, J.; Fan, C., Nanomaterials-based sensors for applications in environmental monitoring. Journal of Materials Chemistry 2012. 18. Khan, S. D.; Jacobson, S., Remote sensing and geochemistry for detecting hydrocarbon microseepages. Geological Society of America Bulletin 2008, 120, (1-2), 96-105. 19. Strand, S.; Austad, T.; Puntervold, T.; Høgnesen, E. J.; Olsen, M.; Barstad, S. M. F., “Smart Water” for Oil Recovery from Fractured Limestone: A Preliminary Study. Energy & Fuels 2008, 22, (5), 3126-3133. 20. Park, Y.; Whitaker, R. D.; Nap, R. J.; Paulsen, J. L.; Mathiyazhagan, V.; Doerrer, L. H.; Song, Y.-Q.; Hürlimann, M. D.; Szleifer, I.; Wong, J. Y., Stability of Superparamagnetic Iron Oxide Nanoparticles at Different pH Values: Experimental and Theoretical Analysis. Langmuir 2012, 28, (15), 6246-6255. 21. Simpson, A. J.; Simpson, M. J.; Soong, R., Nuclear Magnetic Resonance Spectroscopy and Its Key Role in Environmental Research. Environmental Science & Technology 2012, 46, (21), 11488-11496. 22. Park, Y. C.; Paulsen, J.; Nap, R. J.; Whitaker, R. D.; Mathiyazhagan, V.; Song, Y.-Q.; Hürlimann, M.; Szleifer, I.; Wong, J. Y., Adsorption of Superparamagnetic Iron Oxide Nanoparticles on Silica and Calcium Carbonate Sand. Langmuir 2014, 30, (3), 784-792. 23. Berlin, J. M.; Yu, J.; Lu, W.; Walsh, E. E.; Zhang, L.; Zhang, P.; Chen, W.; Kan, A. T.; Wong, M. S.; Tomson, M. B.; Tour, J. M., Engineered nanoparticles for hydrocarbon detection in oil-field rocks. Energy & Environmental Science 2011, 4, (2), 505-509. 24. Yoon, K. Y.; Kotsmar, C.; Ingram, D. R.; Huh, C.; Bryant, S. L.; Milner, T. E.; Johnston, K. P., Stabilization of Superparamagnetic Iron Oxide Nanoclusters in Concentrated Brine with Cross-Linked Polymer Shells. Langmuir 2011, 27, (17), 10962-10969. 25. Ottofuelling, S.; Von Der Kammer, F.; Hofmann, T., Commercial Titanium Dioxide Nanoparticles in Both Natural and Synthetic Water: Comprehensive Multidimensional Testing and Prediction of Aggregation Behavior. Environmental Science & Technology 2011, 45, (23), 10045-10052. 26. Le, N. Y. T.; Pham, D. K.; Le, K. H.; Nguyen, P. T., Design and screening of synergistic blends of SiO 2 nanoparticles and surfactants for enhanced oil recovery in high-temperature reservoirs. Advances in Natural Sciences: Nanoscience and Nanotechnology 2011, 2, (3), 035013. 27. Petosa, A. R.; Jaisi, D. P.; Quevedo, I. R.; Elimelech, M.; Tufenkji, N., Aggregation and Deposition of Engineered Nanomaterials in Aquatic Environments: Role of Physicochemical Interactions. Environmental Science & Technology 2010, 44, (17), 6532-6549. 28. Li, X.; Lenhart, J. J., Aggregation and Dissolution of Silver Nanoparticles in Natural Surface Water. Environmental Science & Technology 2012, 46, (10), 5378-5386. 29. Yu, W. W.; Falkner, J. C.; Yavuz, C. T.; Colvin, V. L., Synthesis of monodisperse iron oxide nanocrystals by thermal decomposition of iron carboxylate salts. Chemical Communications 2004, (20), 2306-2307. 30. Sun, S.; Zeng, H., Size-Controlled Synthesis of Magnetite Nanoparticles. Journal of the American Chemical Society 2002, 124, (28), 8204-8205.

ACS Paragon Plus Environment

28

Page 29 of 32

520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564

Environmental Science & Technology

31. Hyeon, T.; Lee, S. S.; Park, J.; Chung, Y.; Na, H. B., Synthesis of Highly Crystalline and Monodisperse Maghemite Nanocrystallites without a Size-Selection Process. Journal of the American Chemical Society 2001, 123, (51), 12798-12801. 32. Lee, S. S.; Zhang, C.; Lewicka, Z. A.; Cho, M.; Mayo, J. T.; Yu, W. W.; Hauge, R. H.; Colvin, V. L., Control over the Diameter, Length, and Structure of Carbon Nanotube Carpets Using Aluminum Ferrite and Iron Oxide Nanocrystals as Catalyst Precursors. The Journal of Physical Chemistry C 2012, 116, (18), 10287-10295. 33. Prakash, A.; Zhu, H.; Jones, C. J.; Benoit, D. N.; Ellsworth, A. Z.; Bryant, E. L.; Colvin, V. L., Bilayers as Phase Transfer Agents for Nanocrystals Prepared in Nonpolar Solvents. ACS Nano 2009, 3, (8), 2139-2146. 34. Chen, K. L.; Elimelech, M., Aggregation and Deposition Kinetics of Fullerene (C60) Nanoparticles. Langmuir 2006, 22, (26), 10994-11001. 35. Sauerbrey, G., Use of quartz vibrator for weighing thin layers and as a micro-balance. Z. Phys. 1959, 155, (2), 206-222. 36. Chen, K. L.; Elimelech, M., Interaction of Fullerene (C60) Nanoparticles with Humic Acid and Alginate Coated Silica Surfaces: Measurements, Mechanisms, and Environmental Implications. Environmental Science & Technology 2008, 42, (20), 7607-7614. 37. Teng, X.; Yang, H., Effects of surfactants and synthetic conditions on the sizes and selfassembly of monodisperse iron oxide nanoparticles. Journal of Materials Chemistry 2004, 14, (4), 774-779. 38. Xu, X. Q.; Shen, H.; Xu, J. R.; Li, X. J., Aqueous-based magnetite magnetic fluids stabilized by surface small micelles of oleolysarcosine. Applied Surface Science 2004, 221, (1– 4), 430-436. 39. Meerod, S.; Tumcharern, G.; Wichai, U.; Rutnakornpituk, M., Magnetite nanoparticles stabilized with polymeric bilayer of poly(ethylene glycol) methyl ether–poly(ɛ-caprolactone) copolymers. Polymer 2008, 49, (18), 3950-3956. 40. Maity, D.; Agrawal, D. C., Synthesis of iron oxide nanoparticles under oxidizing environment and their stabilization in aqueous and non-aqueous media. Journal of Magnetism and Magnetic Materials 2007, 308, (1), 46-55. 41. Shen, L. F.; Laibinis, P. E.; Hatton, T. A., Bilayer surfactant stabilized magnetic fluids: Synthesis and interactions at interfaces. Langmuir 1999, 15, (2), 447-453. 42. Morgan, L. J.; Ananthapadmanabhan, K. P.; Somasundaran, P., Oleate adsorption on hematite: Problems and methods. International Journal of Mineral Processing 1986, 18, (1–2), 139-152. 43. 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, (7), 3588-3599. 44. Chowdhury, I.; Duch, M. C.; Mansukhani, N. D.; Hersam, M. C.; Bouchard, D., Colloidal Properties and Stability of Graphene Oxide Nanomaterials in the Aquatic Environment. Environmental Science & Technology 2013, 47, (12), 6288-6296. 45. Lin, M. Y.; Lindsay, H. M.; Weitz, D. A.; Ball, R. C.; Klein, R.; Meakin, P., UNIVERSALITY IN COLLOID AGGREGATION. Nature 1989, 339, (6223), 360-362. 46. Hu, Y.; Ray, J. R.; Jun, Y.-S., Na+, Ca2+, and Mg2+ in Brines Affect Supercritical CO2– Brine–Biotite Interactions: Ion Exchange, Biotite Dissolution, and Illite Precipitation. Environmental Science & Technology 2012, 47, (1), 191-197.

ACS Paragon Plus Environment

29

Environmental Science & Technology

565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608

Page 30 of 32

47. Li, K.; Zhang, W.; Huang, Y.; Chen, Y., Aggregation kinetics of CeO2 nanoparticles in KCl and CaCl2 solutions: measurements and modeling. J Nanopart Res 2011, 13, (12), 64836491. 48. Tombácz, E.; Tóth, I. Y.; Nesztor, D.; Illés, E.; Hajdú, A.; Szekeres, M.; L.Vékás, Adsorption of organic acids on magnetite nanoparticles, pH-dependent colloidal stability and salt tolerance. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2013, 435, (0), 91-96. 49. Gregory, J., Particles in Water: Properties and Processes. 2006. 50. Zhang, X.; Yang, S., Nonspecific Adsorption of Charged Quantum Dots on Supported Zwitterionic Lipid Bilayers: Real-Time Monitoring by Quartz Crystal Microbalance with Dissipation. Langmuir 2011, 27, (6), 2528-2535. 51. Donahoe, C. D.; Cohen, T. L.; Li, W.; Nguyen, P. K.; Fortner, J. D.; Mitra, R. D.; Elbert, D. L., Ultralow Protein Adsorbing Coatings from Clickable PEG Nanogel Solutions: Benefits of Attachment under Salt-Induced Phase Separation Conditions and Comparison with PEG/Albumin Nanogel Coatings. Langmuir 2013, 29, (12), 4128-4139. 52. Fatisson, J.; Ghoshal, S.; Tufenkji, N., Deposition of Carboxymethylcellulose-Coated Zero-Valent Iron Nanoparticles onto Silica: Roles of Solution Chemistry and Organic Molecules. Langmuir 2010, 26, (15), 12832-12840. 53. Saleh, N.; Kim, H.-J.; Phenrat, T.; Matyjaszewski, K.; Tilton, R. D.; Lowry, G. V., Ionic Strength and Composition Affect the Mobility of Surface-Modified Fe0 Nanoparticles in WaterSaturated Sand Columns. Environmental Science & Technology 2008, 42, (9), 3349-3355. 54. Reimhult, K.; Petersson, K.; Krozer, A., QCM-D Analysis of the Performance of Blocking Agents on Gold and Polystyrene Surfaces. Langmuir 2008, 24, (16), 8695-8700. 55. Chang, X.; Bouchard, D. C., Multiwalled Carbon Nanotube Deposition on Model Environmental Surfaces. Environmental Science & Technology 2013. 56. Tammelin, T.; Saarinen, T.; Österberg, M.; Laine, J., Preparation of Langmuir/Blodgettcellulose Surfaces by Using Horizontal Dipping Procedure. Application for Polyelectrolyte Adsorption Studies Performed with QCM-D. Cellulose 2006, 13, (5), 519-535. 57. Wu, J.; Alemany, L. B.; Li, W.; Petrie, L.; Welker, C.; Fortner, J. D., Reduction of Hydroxylated Fullerene (Fullerol) in Water by Zinc: Reaction and Hemiketal Product Characterization. Environmental Science & Technology 2014, 48, (13), 7384-7392. 58. Jiang, X.; Tong, M.; Li, H.; Yang, K., Deposition kinetics of zinc oxide nanoparticles on natural organic matter coated silica surfaces. Journal of Colloid and Interface Science 2010, 350, (2), 427-434. 59. Nguyen, T. H.; Chen, K. L., Role of Divalent Cations in Plasmid DNA Adsorption to Natural Organic Matter-Coated Silica Surface. Environmental Science & Technology 2007, 41, (15), 5370-5375. 60. Fatisson, J.; Domingos, R. F.; Wilkinson, K. J.; Tufenkji, N., Deposition of TiO2 Nanoparticles onto Silica Measured Using a Quartz Crystal Microbalance with Dissipation Monitoring. Langmuir 2009, 25, (11), 6062-6069. 61. Song, J. E.; Phenrat, T.; Marinakos, S.; Xiao, Y.; Liu, J.; Wiesner, M. R.; Tilton, R. D.; Lowry, G. V., Hydrophobic Interactions Increase Attachment of Gum Arabic- and PVP-Coated Ag Nanoparticles to Hydrophobic Surfaces. Environmental Science & Technology 2011, 45, (14), 5988-5995.

ACS Paragon Plus Environment

30

Page 31 of 32

609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639

Environmental Science & Technology

62. Xiao, Y.; Wiesner, M. R., Transport and Retention of Selected Engineered Nanoparticles by Porous Media in the Presence of a Biofilm. Environmental Science & Technology 2013, 47, (5), 2246-2253. 63. Faghihnejad, A.; Zeng, H., Hydrophobic interactions between polymer surfaces: using polystyrene as a model system. Soft Matter 2012, 8, (9), 2746-2759. 64. Meyer, E. E.; Rosenberg, K. J.; Israelachvili, J., Recent progress in understanding hydrophobic interactions. Proceedings of the National Academy of Sciences 2006, 103, (43), 15739-15746. 65. Zhuang, J.; Qi, J.; Jin, Y., Retention and transport of amphiphilic colloids under unsaturated flow conditions: Effect of particle size and surface property. Environmental Science & Technology 2005, 39, (20), 7853-7859. 66. 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. Analytical Chemistry 2011, 83, (23), 8838-8848. 67. Tellechea, E.; Johannsmann, D.; Steinmetz, N. F.; Richter, R. P.; Reviakine, I., ModelIndependent Analysis of QCM Data on Colloidal Particle Adsorption. Langmuir 2009, 25, (9), 5177-5184. 68. Olsson, A. L. J.; Quevedo, I. R.; He, D.; Basnet, M.; Tufenkji, N., Using the Quartz Crystal Microbalance with Dissipation Monitoring to Evaluate the Size of Nanoparticles Deposited on Surfaces. ACS Nano 2013, 7, (9), 7833-7843. 69. Quevedo, I. R.; Olsson, A. L. J.; Tufenkji, N., Deposition Kinetics of Quantum Dots and Polystyrene Latex Nanoparticles onto Alumina: Role of Water Chemistry and Particle Coating. Environmental Science & Technology 2013, 47, (5), 2212-2220. 70. Nguyen, T. H.; Elimelech, M., Plasmid DNA Adsorption on Silica:  Kinetics and Conformational Changes in Monovalent and Divalent Salts. Biomacromolecules 2006, 8, (1), 2432. 71. Chowdhury, I.; Duch, M. C.; Mansukhani, N. D.; Hersam, M. C.; Bouchard, D., Deposition and Release of Graphene Oxide Nanomaterials Using a Quartz Crystal Microbalance. Environmental Science & Technology 2013, 48, (2), 961-969.

ACS Paragon Plus Environment

31

Environmental Science & Technology

640

Page 32 of 32

TOC

641

ACS Paragon Plus Environment

32