Subscriber access provided by UNIV OF CONNECTICUT

Article

Enhanced Colloidal Stability of CeO2 Nanoparticles by Ferrous Ions: Adsorption, Redox Reaction, and Surface Precipitation Xuyang Liu, Jessica Renee Ray, Chelsea W. Neil, Qingyun Li, and Young-Shin Jun Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es506363x • Publication Date (Web): 07 Apr 2015 Downloaded from http://pubs.acs.org on April 12, 2015

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

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

Page 1 of 31

Environmental Science & Technology

Enhanced Colloidal Stability of CeO2 Nanoparticles by Ferrous Ions: Adsorption, Redox Reaction, and Surface Precipitation  

Xuyang Liu †, Jessica R. Ray †, Chelsea W. Neil †, Qingyun Li, and Young-Shin Jun*   Department of Energy, Environmental and Chemical Engineering, Washington University in St. Louis, St. Louis, MO 63130, United States    

E-mail: [email protected] http://encl.engineering.wustl.edu/ Submitted: December 2014 Revised: March 2015

Environmental Science &Technology

†These

authors contributed equally to the current work.

*To whom correspondence should be addressed.

ACS Paragon Plus Environment

Environmental Science & Technology

  1 

ABSTRACT



Due to the toxicity of cerium oxide (CeO2) nanoparticles (NPs), a better understanding of the redox



reaction-induced surface property changes of CeO2 NPs and their transport in natural and



engineered aqueous systems is needed. This study investigates the impact of redox reactions with



ferrous ions (Fe2+) on the colloidal stability of CeO2 NPs. We demonstrated that under anaerobic



conditions suspended CeO2 NPs in a 3 mM FeCl2 solution at pH = 4.8 were much more stable



against sedimentation than those in the absence of Fe2+. Redox reactions between CeO2 NPs and



Fe2+ lead to the formation of 6-line ferrihydrite on the CeO2 surfaces, which enhanced the colloidal



stability by increasing the zeta potential and hydrophilicity of CeO2 NPs. These redox reactions

10 

can affect the toxicity of CeO2 NPs by increasing cerium dissolution and by creating new Fe(III)

11 

(hydr)oxide reactive surface layers. Thus, these findings have significant implications for

12 

elucidating the phase transformation and transport of redox reactive NPs in the environment.

1   

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

Environmental Science & Technology

  13 

INTRODUCTION

14 

Cerium oxide (CeO2) nanoparticles (NPs) have been enumerated in the priority list of

15 

engineered nanomaterials for risk evaluation by the Organization for Economic Co-operation and

16 

Development (OECD) due to their wide application in industry and daily life.1 CeO2 NPs are ideal

17 

in catalysis applications, such as diesel engine catalytic converters,2 due to their oxygen storage

18 

capabilities3 and their ability to readily participate in Ce4+/Ce3+ redox processes. The use and

19 

production of CeO2 NPs will inevitably result in increased concentrations in natural or manmade

20 

aqueous environments, such as wastewater.4 Therefore, in order to better predict the life cycle of

21 

these engineered nanoparticles, it is vital to improve our understanding of their fate and transport,

22 

particularly in aqueous environments.

23 

Based on a 2001 report on human health risks of cerium from diesel fuels, the average

24 

worldwide estimated level of cerium in soils was 20–60 ppm.5 Recent studies have also shown that

25 

nanometer sized CeO2 particles are found in automobile exhaust.6,

26 

warrants immediate attention to prevent harmful effects to the biosphere. The bioavailability and

27 

toxicity of CeO2 NPs are largely determined by their fate and transport in the environment, which

28 

is in turn affected by their surface charge and aggregation state.8, 9 While recent investigations in

29 

this field have focused on the influence of solution chemistry and organic matter on the fate and

30 

transport of CeO2 NPs,2, 10-16 few studies have considered the effect of redox reactions on CeO2

31 

NP surface properties in the presence of redox-active ions in diverse aqueous environments.17, 18 

7

This high concentration

32 

When released in the environment, CeO2 NPs can coexist and interact with redox sensitive

33 

elements. One important element to consider is ferrous iron (Fe2+), which is widely distributed in

34 

natural aqueous systems (e.g., acid mine drainage),19 as well as in engineered systems for odor and

2   

ACS Paragon Plus Environment

Environmental Science & Technology

  35 

corrosion control, precipitating hydrogen sulfide, and phosphorus removal.20, 21 The concentration

36 

of ferrous ions used in these engineered systems can be as high as 9.85 mM for phosphorus

37 

removal,22 and 1.5–3.0 mM to remove cyanide from industrial wastewater.23 In addition, in a study,

38 

Fe3O4/CeO2-impregnated NPs were synthesized and used as Fenton-like catalysts for hydroxyl

39 

radical generation and 4-chlorophenol degradation with the addition of H2O2.24 Ce4+ has also been

40 

used in the production of silicate glasses to rid the glass of Fe2+ ions.25 Therefore, anthropogenic

41 

CeO2 NPs originated from industrial wastewater can interact with ferrous ions during their

42 

industrial applications, as well as during wastewater treatment processes.26, 27

43 

According to the difference in the standard redox potential of Fe3+/Fe2+ (0.77V) and

44 

Ce4+/Ce3+ (1.44 V), redox reactions between Fe2+ and CeO2 can occur when they coexist in

45 

solution.28 However, it is largely unknown how this reaction with aqueous Fe2+ will affect the

46 

surface properties and colloidal stability of CeO2 NPs in aqueous environments. In this study, we

47 

systematically investigated changes in CeO2 NP surface properties when aqueous Fe2+ is present.

48 

Knowledge obtained in this study can help improve our understanding of the fate and transport of

49 

redox active nanomaterials during their lifetime, which will in turn give insight into the expected

50 

bioavailability of these anthropogenic NPs when released into the environment.

51 

EXPERIMENTAL SECTION

52 

Preparation of Nanoparticle Dispersions for Aggregation and Sedimentation Tests

53 

All solution preparations and the following wet experiments were conducted in an

54 

anaerobic Coy chamber to prevent the influence of dissolved oxygen (DO). By distinguishing Fe2+

55 

oxidation by CeO2 from oxidation by molecular oxygen, we can provide a better understanding of

56 

the redox reactions and mechanisms in aqueous Fe2+–CeO2 NP systems. Depleted oxygen can

3   

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Environmental Science & Technology

  57 

occur under specific conditions, such as in anoxic wastewater treatment processes, underground,

58 

or at the bottom of stratified lakes.29 Furthermore, in acidic environment, the Fe2+ oxidation rate

59 

by dissolved O2(aq) equilibrated at 0.2 atm is quite slow. Thus, our anaerobic experimental

60 

conditions can still be applicable in wastewater systems where Fe2+ is present and the local pH is

61 

acidic. To ensure anoxic conditions, deionized (DI) water was boiled prior to use in these

62 

experiments to remove DO, and then cooled to room temperature in the anaerobic chamber.

63 

We applied commercially available CeO2 NPs (Sigma-Aldrich, St. Louis, MO), so that our

64 

results have closer connections to real engineered NP–environmental systems. CeO2 NP

65 

dispersions of 50 mg/L were created in 50 mL test tubes inside the chamber, and the test tubes

66 

were indirectly ultrasonicated for 1 h before reaction using a Fisher Scientific ultrasonic cleaner

67 

(model no. FS6) with a frequency of 50/60 kHz and power of 30W. For all experiments, the ionic

68 

strength (IS) was maintained at 10 mM. In the reaction system, 3 mM FeCl2 and 1 mM NaCl were

69 

added to the CeO2 dispersions to have an ionic strength (IS) of 10 mM and a pH of 4.8 ± 0.2. The

70 

final CeO2 concentration of the mixed solution was 45 mg/L. A weakly acidic pH can occur in

71 

acid mine drainage environments.30, 31 In addition, when applying iron in wastewater treatment

72 

plants, the hydrolysis of iron could lead to pHs of 5 or less if no pH control is conducted., 21, 32 At

73 

pH 4.8, initial ferrous iron remains largely in the ferrous state, which could have greater

74 

toxicological effects than other iron species on organisms, e.g., the feeding activity and motility of

75 

the mayfly larvae.33, 34 This is also the pH of the system solution after FeCl2 addition, so errors in

76 

the ionic strength through pH adjustment can be minimized. The concentration of 3 mM Fe2+ used

77 

in our experimental system is commonly found in natural environments (e.g. 2–4 mM in an anoxic

78 

lake in Massachusetts) and wastewater treatment processes that use additive ferrous iron.22, 23, 35 In

79 

the control experiment,10 mM NaCl (with no FeCl2) was added to the CeO2 dispersion and the pH 4   

ACS Paragon Plus Environment

Environmental Science & Technology

  80 

was adjusted to pH 4.8 with dilute HCl solution to match that of the Fe2+-containing system and

81 

maintain the final CeO2 concentration of the mixed solution as 45 mg/L. The pH change during

82 

the reaction period was monitored (Figure S6 in the SI). In addition, these conditions give valuable

83 

insight into the interactions that occur between Fe2+ and CeO2, and can be used to bolster further

84 

studies of CeO2 NP stability in more complex aqueous systems.

85 

During sedimentation experiments, aliquots were taken from the supernatant at elapsed

86 

times using a pipette with minimal disturbance to the suspension. The concentrations of CeO2 NPs

87 

were measured by UV-vis spectroscopy (Varian Inc., Cary 50 Bio UV-Vis Spectrophotometer,

88 

Palo Alto, California) at a wavelength of 305 nm, where the highest absorbance was obtained

89 

(Figure S1 in the Supporting Information). Sedimentation experiments were run for 93 hours in

90 

triplicate. To verify the changes in CeO2 surface properties following redox reactions, after

91 

reaction the zeta potentials and particle sizes (Dynamic Light Scattering, DLS) of CeO2

92 

nanoparticle aggregates in the Fe2+ and control systems were measured using a Zetasizer (Nano

93 

ZS, Malvern Instruments Ltd., Westborough, MA). The zeta potential was derived from the

94 

original electrophoretic mobility using the Smoluchowski equation. We have characterized the

95 

particle size using DLS (Table S1) and TEM (Figure S7). The hydrodynamic particle sizes by DLS

96 

were collected for at least three measurements (with each measurement taken over 10 s) and

97 

triplicate or more experiments were conducted for each sample. The hydrodynamic diameter is

98 

defined as the diameter of the aggregate plus that of the hydration layer.

99 

CeO2 NP Dissolution Experiments

100 

Due to the low CeO2 solubility,36 CeO2 suspensions of 250 mg/L and 15 mM FeCl2 were

101 

used in dissolution experiments in order to achieve aqueous Ce concentrations above the 25 μg/L

5   

ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

Environmental Science & Technology

  102 

detection limit of the inductively coupled plasma-optical emission spectrometer (ICP-OES).

103 

Similar or even higher concentrations of CeO2 for dissolution experiments have been also

104 

investigated by other researchers.37, 38 Because the Fe2+ concentration scales up proportionally, we

105 

expect that this elevated concentration system is relevant to lower suspended CeO2

106 

concentrations.37, 39, 40 Thus, the concentration we used for our dissolution tests reasonably allows

107 

us to compare our results with previous CeO2 NP studies.

108 

Reaction and control solutions were prepared in the anaerobic chamber as described for

109 

sedimentation experiments. Triplicate batches of the reaction and control systems were separated

110 

into 5 mL test tubes and put into test tube rotators in the chamber.39, 41, 42 At 1, 2, 3, 4, 5, and 6

111 

hour time points, triplicate samples were taken from the two systems. Samples were placed in

112 

ultracentrifuge tubes (PC Oak Ridge Tubes, Fisher Scientific) in the anaerobic chamber, capped,

113 

and removed from the chamber. To separate the supernatant from the CeO2 NPs, samples were

114 

then ultracentrifuged using a Thermo Scientific Sorvall WX Ultra Series Centrifuge with a T-865

115 

Fixed Angle Rotor at 40,000 rpm (or 115,861 x g) for 30 minutes. In addition to centrifugation,

116 

the samples were placed back inside the anaerobic chamber and filtered by a 0.2 µm filter

117 

(Millipore syringe filter) to ensure the removal of all bulk CeO2 NPs and aggregates. We have

118 

verified the efficacy of combining ultracentrifugation and filtration to separate CeO2 NPs in

119 

preliminary tests, and this method has also been commonly used in CeO2 NP dissolution and

120 

separation studies.39, 41, 42 The filtrate was collected and acidified to 1% v/v nitric acid for ICP-

121 

OES measurements.

6   

ACS Paragon Plus Environment

Environmental Science & Technology

  122 

Phase Identification of Fe(III) (Hydr)oxides on CeO2 NP Surfaces

123 

We conducted replicate experiments for X-ray absorption spectroscopy (XAS) analysis.

124 

After reacting for 6 hours, solutions were transferred to centrifugation tubes and capped in the

125 

anaerobic chamber, then ultracentrifuged at 40,000 rpm for 30 minutes. Once centrifuged, the

126 

supernatant was poured off, leaving the CeO2 NPs fixed at the bottom of the tube. The supernatant

127 

was then replaced with deoxygenated DI water and the tubes were then capped and removed from

128 

the chamber for an additional 30 min of ultracentrifugation to remove excess salt. After the second

129 

ultracentrifugation, the DI water was poured off in the anaerobic chamber, and the solids were

130 

allowed to dry in the chamber overnight to prevent any oxidation. The cerium spectra were

131 

measured in transmission mode, and iron spectra were measured in fluorescence mode.

132 

Experiments were conducted at Beamline 13BM-D at the Advanced Photon Source (APS),

133 

Argonne National Laboratory. This station utilized a Si(111) monochromator, giving it a focused

134 

beam size of 10 m by 30 m and a resolution of 1  10-4 E/E. The energy flux was 1  109 at 10

135 

keV. The energy range for this station was 4.5–70 keV. The iron XANES edge was measured at

136 

7.119 keV and the cerium edge was measured at 40.444 keV.

137 

Secondary Mineral Phase Precipitation on CeO2-Sputtered Substrates by Physical Vapor

138 

Deposition (PVD)

139 

To identify the properties of reaction products on the surface of NPs, CeO2 substrates were

140 

created by sputtering CeO2 NPs on clean Si wafers using a PVD process (Kurt Lesker PVD 75,

141 

Livermore, CA). The DC (diode) mode was applied under 1 mTorr argon pressure, with 100 watts

142 

power input for 2000 seconds. The deposition was monitored in situ via a built-in quartz crystal

143 

microbalance. Because CeO2 nanoparticles were stable, solid phase analyses were conducted under

7   

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Environmental Science & Technology

  144 

atmospheric conditions. The thickness of the CeO2–sputtered wafer was quantified by alpha-SE

145 

ellipsometry (Lincoln, NE) and atomic force microscopy (AFM, Veeco, Nanoscope V) after the

146 

PVD sputtering process. During experiments, the CeO2 substrate was exposed to the same solution

147 

chemistry as the aqueous CeO2 NP experiments, containing either NaCl only or NaCl and Fe2+

148 

ions (Figure S2). A clean Si wafer was also used as a control to compare precipitation in the system

149 

containing NaCl and Fe2+ ions in the absence of CeO2. The precipitates on the CeO2-sputtered

150 

substrate and Si wafer were analyzed by AFM using tapping mode. We collected height, amplitude,

151 

and phase contrast information simultaneously for 5 µm × 5 µm areas, and analyzed the images

152 

using the Nanoscope 7.20 software. Experimental details on the AFM setup have been reported in

153 

our former studies.43, 44

154 

Grazing Incidence Small-Angle X-ray Scattering (GISAXS) and SAXS Measurements

155 

The surface properties and precipitates on the CeO2-sputtered substrates were also

156 

characterized in situ by GISAXS. In this experiment, the solution was prepared under the same

157 

solution chemistry as the CeO2 NP sedimentation tests, i.e., 3 mM FeCl2 and 1 mM NaCl at pH =

158 

4.8. For the control system, a 10 mM NaCl solution was used. The CeO2 substrates were placed

159 

flat at the bottom of a specially designed GISAXS cell. The solutions were injected at the top of

160 

the cell, and then the cell was capped. The substrates were reacted for 1 hour. During GISAXS

161 

measurements, incident X-ray beams at 18 keV were passed through the cell, where they interacted

162 

with particles precipitating on the substrate surface (GISAXS). The scattered X-ray beams were

163 

collected by a 2D detector. X-ray scattering data was processed by cutting along the Yoneda wing.

164 

All data reduction was conducted using the GISAXS shop macro, a software package available at

165 

APS beamline 12-ID. More detailed data fitting procedure descriptions are available in our

166 

previous publications.43-45 Prior to size fitting, the background image of the CeO2 sputtered wafer 8   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

  167 

in DI water was subtracted from all 2D GISAXS scattering images. Thus, observed changes in

168 

particle size were from the precipitation of Fe(III) (hydr)oxides. During the particle size fitting,

169 

the shapes of the particles were assumed to be spherical, thus a form factor for polydisperse

170 

spherical particles. A Schultz size distribution was also assumed for fitting.

171 

RESULTS & DISCUSSION

172 

Enhanced Colloidal Stability of CeO2 NPs in Fe2+ Solutions

173 

First, we determined how the presence of Fe2+ affects the aggregation and subsequent

174 

sedimentation of CeO2 NPs. During the 4 day monitoring period, the concentration of CeO2 NPs

175 

in the supernatant was higher in the system with added Fe2+ than in the system without aqueous

176 

Fe2+ (Figure 1A), indicating that the presence of 3 mM Fe2+ promoted the colloidal stability of

177 

CeO2 NPs. For example, only 46% of the initial CeO2 NP’s concentration remained in the

178 

supernatant in the control system while 94% of the initial CeO2 NPs remained stable after 16 h in

179 

the presence of 3 mM FeCl2. In addition, the significant difference in NP sedimentation due to the

180 

presence of Fe2+ can be observed visually in the solutions after stirring overnight (Figure S3,

181 

Supporting Information). It is noted that the effect of double layer compression was similar in the

182 

solutions because they had same 10 mM ionic strength and same anion identity (i.e., Cl-), which

183 

will dominate double layer compression effects due to the positive charge of CeO2 NPs.

184 

Analyses of particle sizes and zeta potentials in the two experimental systems also help to

185 

illustrate the observed sedimentation trends. The zeta potential of CeO2 was 28.6 ± 1.7 mV in the

186 

presence of Fe2+, while in the absence of Fe2+, the zeta potential was 8.7 ± 0.8 mV (Table S1,

187 

Supporting Information). The isoelectric point, pHiep, of CeO2 NPs in 10 mM NaCl was measured

188 

to be pH 4.95 (Figure S4A), which is within the range of 3.0–7.6 found in the literature. The wide

9   

ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

  189 

variation in this range can be due to the varying surface properties of the nanoparticles produced

190 

through different methods.46-48 In our experimental system, the unreacted CeO2 NPs should be

191 

positively charged under the test conditions.

192 

The hydrodynamic size for stable suspensions (Table S1) and XRD spectra (Figure S4B)

193 

for the unreacted CeO2 NPs are given in the SI. The hydrodynamic size was 137.7 ± 0.7 nm for

194 

stable dispersion of unreacted CeO2 in DI water, while the particle sizes measured by TEM ranged

195 

from 5–30 nm (Figure S7), consistent with the manufacture's nominal size. After 16 hours reaction,

196 

the hydrodynamic diameter (DH) of CeO2 NPs in the 3 mM Fe2+ solution was 154 ±1 nm, while

197 

the DH of CeO2 NPs in the 10 mM NaCl system was 3829 ± 358 nm (Table S1). The smaller size

198 

of CeO2 aggregates when Fe2+ is present is in accordance with the higher suspended concentrations

199 

of CeO2 NPs compared to the system without Fe2+, because the smaller aggregates will be more

200 

stable in solution. In contrast, larger aggregates tend to settle out because gravitational forces are

201 

predominant over Brownian motion.

202 

Identification of the Reaction Products and CeO2 Dissolution

203 

We hypothesize that the Fe2+-promoted stability of CeO2 NPs results from redox

204 

reactions28 between Fe2+ and CeO2 NPs and the formation of Fe(III)products at the CeO2 NP

205 

surfaces (eqn. (1)). Throughout the current manuscript, we expressed the oxidation state of solid

206 

and sorbed species as Roman numerals and the oxidation state of aqueous species as Arabic

207 

numerals. Arabic numerals were also used when both solid and aqueous species are possible. ≡CeIVO2+ Fe2+ (aq)↔ Ce3+ (aq)(or ≡CeIII2O3) + ≡Fe3+

208 

10   

ACS Paragon Plus Environment

(1)

Environmental Science & Technology

Page 12 of 31

  209 

X-ray absorption spectroscopy (XAS) was carried out to determine the Ce K-edge and Fe

210 

K-edge spectra for the solid CeO2 NPs before and after reaction in the FeCl2 and control systems

211 

(Figure 2). X-ray absorption near edge structure (XANES) results for Ce indicated no significant

212 

change in the oxidation state after reaction. In other words, there is no CeIII contribution in both

213 

initial CeO2 NPs and reacted CeO2 NPs (Figure 2A). This could be due to negligible amounts of

214 

CeIII on the surface compared to bulk CeIV. Considering that the presence of surface CeIII2O3 was

215 

not detected, and the aqueous Ce concentration was significantly enhanced in the presence of Fe2+

216 

(as will discussed shortly after), we suggest that aqueous Ce3+ is the more predominant reaction

217 

product than solid CeIII2O3. For iron, XAS results (Figure 3B) indicate that all iron on the reacted

218 

CeO2 NP surface was oxidized to FeIII. Furthermore, the iron spectrum for the reacted CeO2 NPs

219 

is consistent with that of the ferrihydrite standard (Figure 2B).

220 

Aqueous Ce levels were monitored for the first 6 hours of reaction in the 250 mg/L CeO2

221 

system (Figure 1B) because the most vigorous interfacial reaction occurs during the early stage.

222 

Based on the concentration of dissolved Ce from CeO2 NPs in the supernatant, we also found that

223 

the dissolved Ce concentration was 15 times higher in the presence of aqueous Fe2+ than in the

224 

control system in the absence of Fe2+ over the course of the reaction period. Because the solubility

225 

of Ce3+ is 25 orders of magnitude higher than CeIVO2, and it was reported that under similar

226 

aqueous conditions to our experimental condition, more than 99% dissolved Ce was Ce3+,36, 49 we

227 

assumed that the aqueous Ce in our system was primarily Ce3+. Thus, the increased concentration

228 

of dissolved Ce could result from the reduction of CeIVO2 by Fe2+. Then, the redox reaction is also

229 

likely to accelerate the precipitation of Fe(III)(hydr)oxides on the surface of CeO2 NPs while the

230 

dissolved Ce3+ ions are released from the CeIVO2 NP surfaces.

11   

ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

  231  232 

Reaction Pathways: Adsorption, Redox Reaction, and Surface Precipitation of Fe(III)(Hydr)oxides on the Surfaces of CeO2 NPs and Dissolution of CeO2 NPs

233 

A series of reactions could contribute to the enhanced colloidal stability of CeO2 in the

234 

presence of Fe2+ ions. MgCl2 extraction revealed that the adsorbed Fe2+ concentration on the

235 

surface of CeO2 NPs was 3.75 mg/L, which is equivalent to 0.05% of the total iron (Figure S5,

236 

Supporting Information). The Fe2+ can be adsorbed on the CeO2 NPs by ion exchange (eq. (2)),

237 

and such a process is expected to shift the zeta potential of CeO2 NPs to be more positive and

238 

release H+ into the solution:50

239 

Fe2+ Sorption: ≡CeOH + Fe2+ ↔ ≡CeOFe+ + H+

240 

On the other hand, the adsorbed FeII can react further with CeIV on the surface of CeO2 NPs

241 

rather than existing as ferrous state, due to the large enough difference in the standard redox

242 

potential for Ce4+/Ce3+ (1.44 V) and Fe3+/Fe2+ (0.77 V).28 As a control experiment, we tested the

243 

reaction of aqueous Ce4+ using Ce(SO4)2 and Fe2+ ions in solution and observed a significant

244 

decrease of Fe2+ concentration by the Ferrozine method after 1 d (Table S2, Supporting

245 

Information). Furthermore, the XANES experiments of the CeO2 NP‒Fe2+ system also revealed

246 

oxidized Fe(III) on the surface of CeO2 NPs under anaerobic conditions, providing direct evidence

247 

of surface redox reactions (Figure 2).

(2)

248 

In the experimental systems, the pH of the reaction solution (CeO2 NPs + FeCl2) was measured

249 

to be lower than either CeO2 NPs or FeCl2 solutions alone, and decreased during the first few hours

250 

of reaction (Figure S6, Supporting Information). The decrease of solution pH can be attributed to

251 

both the adsorption of Fe2+ through ion exchange (eqn. (2)) and the hydrolysis of Fe3+, the redox

252 

reaction product, which releases more H+ into solution than is released by Fe2+ sorption. The Fe2+12   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 31

  253 

only system also has slightly decreased pH due to minor Fe2+ hydrolysis (eqn. 3a), but the rate of

254 

pH decrease is much less than that of the CeO2/Fe2+ system, which undergoes the hydrolysis of

255 

ferric ions (eqn. 3b).

256 

Hydrolysis: Fe2+ + H2O  Fe(OH)+(aq) + H+ or

(3a)

257 

Fe3+ + 3 H2O  Fe(OH)3 (aq) + 3 H+

(3b)

258 

Once hydrolysis of ferric iron forms the monomer (eqn. 3b), then dimers and polymers can

259 

form through continuous olation (hydroxo–bridging) and oxolation (oxo–bridging) reactions.44

260 

Once the polymeric cluster size is larger than the critical nucleus size, stable nuclei form, and

261 

Fe(III) (hydr)oxide precipitation occurs.45

262 

To observe the particle formation of Fe(III) (hydr)oxides and their morphological changes

263 

in Fe2+–CeO2 NP systems, AFM and GISAXS were used on a CeO2-sputtered Si wafer reacted in

264 

the same solution conditions. AFM images show significant precipitation of FeIII solid phase on

265 

the CeO2-sputtered Si wafer substrate after 6 h (Figure 3). The particle size of precipitated Fe(III)

266 

(hydr)oxides on the CeO2 substrates increased from 4.8 ± 0.7 nm (based on 20 particle analysis)

267 

to 10.6 ± 2.8 nm after 6 h, and 25.9 ± 5.4 nm after 1 day (based on 50 particle analysis each). In

268 

contrast, precipitation of Fe(III) (hydr)oxide on a clean Si substrate was negligible within 1 day.

269 

GISAXS results also showed Fe(III) (hydr)oxide nucleation on the surface of CeO2-sputtered

270 

substrates (Figure 4). Figure 4A revealed that particles with an Rg of 1.72 nm were observed on

271 

the surface of CeO2 substrates in the presence of 3 mM Fe2+ in 1 h. In contrast, no particles formed

272 

in the absence of Fe2+due to no scattering increase for the CeO2 only system. SAXS images in

273 

Figure 4B indicates that no particles formed in solution in the absence of CeO2 (e.g., Fe2+ only

274 

system) as well. TEM analysis of CeO2 NPs revealed that nucleated Fe(III) (hydr)oxide covered 13   

ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

  275 

the surfaces of CeO2 NPs and appeared to increase the CeO2 NP particle–particle distance (Figure

276 

S7). Furthermore, electron diffraction analysis of the Fe(III) (hydr)oxide products from the CeO2‒

277 

Fe2+ system revealed the formation of 6-line ferrihydrite, which is consistent with XAS results. It

278 

is important to note that artifacts and phase transformation might be introduced during TEM

279 

sample preparation due to the drying process.43 Therefore, the morphology of the in situ Fe(III)

280 

(hydr)oxide particles might be different than what is observed using TEM. However, the multiple

281 

complementary techniques used in this study provide convincing and consistent evidence that the

282 

precipitation of iron on the CeO2 surfaces greatly affected the surface properties and colloidal

283 

stability of CeO2 NPs. Based on the above discussion, eqn. (1) can be rewritten as follows:

284 

≡CeIVO2+ Fe2+ (aq)↔ ≡CeIVO2 + Ce3+ (aq) + ≡Fe(III)(hydr)oxide (e.g., ferrihydrite).

285 

(4)

Enhanced Stability by Iron (Hydr)oxide Surface Precipitation

286 

First, we analyzed the change in stability of CeO2 NPs using classical Derjaguin–Landau–

287 

Verwey–Overbeek (DLVO) interactions, comprised of electrostatic repulsion forces and van der

288 

Waals attractions. The electrostatic interaction increases with the absolute value of surface

289 

potential (eqns. (S1-3) in Section S4 of the Supporting Information). Nanoparticles with higher

290 

zeta potential have higher electrostatic repulsions each other, and therefore, are more stable in

291 

suspension. Hence, the increased zeta potential due to the adsorption of Fe2+ and precipitation of

292 

Fe(III)(hydr)oxides on the CeO2 NP surfaces can lead to more significant electrostatic repulsions.

293 

In addition, the values of the Hamaker constant for iron oxide phases (magnetite, maghemite, and

294 

hematite) in the literature range from 1.3 to 4.5 × 10-20 J.51 Although the Hamaker constant for

295 

ferrihydrite is not available, the highest reported value for an iron oxide related phase (4.5 × 10-20

296 

J) is still smaller than that of CeO2 (5.6–6.0×10-20 J).2, 52 A smaller Hamaker constant signifies that 14   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

  297 

there are less attractive forces between Fe(III) (hydr)oxide–Fe(III) (hydr)oxide than that of the

298 

CeO2 NP–CeO2 NP. The precipitated Fe(III)(hydr)oxides on the surface of CeO2 will most likely

299 

increase the electrostatic repulsion and decrease the van der Waals attractions.

300 

Non-DLVO interactions also possibly play a significant role in the interactions of CeO2

301 

NPs after being reacted with Fe2+. We hypothesize that the enhanced Ce stability in the presence

302 

of Fe2+ may also result from Fe(III) (hydr)oxide precipitation changing the CeO2 NP surface

303 

hydrophilicity. Due to the distinctive electronic structure of rare earth atoms, the CeO2 surface has

304 

intrinsically less hydrophilic properties.53 In a cerium atom, the outer full octet of electrons in the

305 

5s2p6 shell shields the unfilled 4f orbitals. As a result, CeO2 tends to not exchange electrons and

306 

form hydrogen bonds with surrounding water molecules, making it less hydrophilic than other

307 

metal oxides.53 The precipitated Fe(III) (hydr)oxides could, therefore, make the NP surface more

308 

hydrophilic than the CeO2 surface.53 As a result, the coated NPs would be more stable in the

309 

aqueous phase. To test the extent of hydrophilicity changes of CeO2 NPs by Fe(III) (hydr)oxide

310 

surface coatings, we conducted surface angle measurements. It appeared that the surface of CeO2-

311 

sputtered wafers became more hydrophilic after the precipitation of Fe(III) (hydr)oxides within a

312 

3 day reaction period. (Figure S8, Supporting Information)

313 

ENVIRONMENTAL IMPLICATIONS

314 

In this study, we found that redox reactions between CeO2 NPs and Fe2+ lead to the

315 

formation of 6-line ferrihydrite on the CeO2 surface, which enhanced the colloidal stability by

316 

increasing the zeta potential and hydrophilicity of CeO2 NPs. The findings of this work suggest

317 

longer retention periods and farther transport distances of CeO2 NPs in aquatic environments

318 

containing Fe2+ ions. This study calls for immediate attention to the significant effects of redox

15   

ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

  319 

reactions and surface precipitation on the fate, transport, and bioavailability of engineered NPs in

320 

aquatic environments. These nanoparticles could act as heterogeneous nucleation sites and

321 

adsorption sites when released into the environment, incorporating toxic elements and molecules

322 

into a “hybrid” engineered/natural nanoparticle composite.54 Redox reactive elements, e.g., Fe2+,

323 

Mn2+, and contaminants, such as As, Cr, and U, can be adsorbed on and react with redox reactive

324 

engineered NPs such as CeO2 NPs.55, 56 In addition, the positively charged CeO2 can adsorb and

325 

aggregate with negatively charged natural colloids or polymers. In the presence of Fe2+, the new

326 

hybrid Fe(III) hydroxide coated CeO2 nanoparticles become more positively charged and may take

327 

longer to be destabilized by the natural colloids than in the absence of Fe2+ ions. In addition, the

328 

interactions of adsorbed natural polymers with Fe2+ (e.g., complexation with natural polymers)

329 

can make the redox reaction of CeO2 more complicated.57 The findings of the current study provide

330 

an important starting point for investigating the long term influences of Fe2+ on CeO2 fate, transport,

331 

and toxicity. These reactions have a significant impact on the transport and transformation of both

332 

NPs and contaminants.

333 

Consequential changes in the physicochemical properties of CeO2 NPs can affect the

334 

toxicity by altering factors such as the surface charge, particle size, production of reactive oxygen

335 

species (ROS), dissolution of reactive ions, hydrophilicity, and surface functionality or coatings.

336 

For instance, positively charged CeO2 NPs were found to penetrate C. elegans cell membranes

337 

more easily than the neutral and negatively charged CeO2 NPs, making them more toxic.8 Because,

338 

in general, dissolved Ce3+ is far more toxic than CeO2 NPs,27 the presence of Fe2+can also enhance

339 

their toxicity due to increased dissolution. Conversely, the surface of CeO2 NPs may become

340 

passivated with time by Fe(III) (hydr)oxide coatings. This would retard further dissolution and

341 

could help mitigate CeO2 NP toxicity. CeO2 NP redox processes with Fe2+ and other redox active 16   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

  342 

species could also interact with natural organic matter in wastewater treatment plants, potentially

343 

affecting their toxicity to microorganisms in wastewater treatment plants.14, 39 In addition to CeO2

344 

NPs, many engineered NPs may be subject to surface reactions and precipitation due to ubiquitous

345 

redox reactive elements in the environment. These interactions need to be considered

346 

comprehensively while evaluating the fate of engineered CeO2 NPs and assessing their risk in

347 

aquatic environments.

348 

Supporting Information Available

349 

Supporting information includes experimental descriptions, UV-Vis spectra, DLS data, zeta

350 

potential measurements, TEM images, an ED pattern, and pH monitoring data. This material is

351 

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

352 

Acknowledgments

353 

This work is supported by the National Science Foundation’s Environmental Chemical Science

354 

Program (CHE-1214090) and Washington University’s Faculty Startup. JRR was supported by the

355 

Environmental Protection Agency STAR Fellowship and CWN was supported by the Mr. and Mrs.

356 

Spencer T. Olin Fellowship. We would like to thank Dr. Seonke Seifert of the Advanced Photon

357 

Source Sector 12-ID-C at Argonne National Laboratory, supported by the U.S. Department of

358 

Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-

359 

06CH11357 and the Institute of Materials Science and Engineering and Nano Research Facility at

360 

WUStL for experimental support. XAS work was performed at GeoSoilEnviroCARS (Sector 13),

361 

Advanced Photon Source (APS), Argonne National Laboratory. GeoSoilEnviroCARS is supported

362 

by the National Science Foundation-Earth Sciences (EAR-1128799) and Department of Energy-

363 

GeoSciences (DE-FG02-94ER14466). We thank Dr. Matt Newville and Dr. Tony Lanzirotti for 17   

ACS Paragon Plus Environment

Page 19 of 31

Environmental Science & Technology

  364 

their help with XAS experiments. We appreciate the assistance and constructive suggestions from

365 

our colleagues in the Environmental NanoChemistry Lab (ENCL) at WUStL.

18   

ACS Paragon Plus Environment

Environmental Science & Technology

  366 

TOC Art

367 

19   

ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

Environmental Science & Technology

  368 

LIST OF FIGURES

369 

Figure 1.(A) Sedimentation kinetics of CeO2 NPs at 10 mM IS and pH 4.8 in the presence and

370 

absence of aqueous Fe2+ ions. The percentages were obtained from suspended nanoparticle

371 

concentration normalized by the initial suspended concentration measured by UV-Vis. The error

372 

bars represent the standard deviation of CeO2 concentration from triplicate experiments. (B)

373 

Dissolved Ce concentrations from CeO2 NPs in the presence and absence of Fe2+ at pH 4.8.

374 

Figure 2. XAS spectra for CeO2 NPs reacted in the presence of 3 mM FeCl2 (A: Ce K-edge and

375 

B: Fe K-edge). The increase in energy of the K-edge position compared to the FeCl2 standard

376 

indicates oxidation of Fe(II) to Fe(III). XAS results showing (A) cerium K-edge and (B) iron K-

377 

edge spectra for reacted samples and standards. Ce K-edge results show no detectable Ce(III) on

378 

the surface. Fe K-edge results show the most similarity between ferrihydrite and the Fe(III) phase

379 

formed on the CeO2 NPs after the 6 h reaction.

380 

Figure 3. Representative AFM images for the precipitation of iron oxide particles on (A) CeO2-

381 

sputtered Si substrates and (B) pure Si control substrates. Height scale (HS) is 10 nm unless

382 

otherwise noted.

383 

Figure 4.1D reduced grazing incidence small angle X-ray scattering (GISAXS) data of

384 

Fe3+precipitation on the surfaces of CeO2 sputtered wafers ( , Figure 4A) and data from the CeO2

385 

NP control system ( , Figure 4A). SAXS raw data of the Fe2+ only system is depicted in Figure

386 

4B. Figure 4A revealed that particles were observed in CeO2 + Fe2+ system with Rg = 1.72 nm (at

387 

1 hr), and no particles formed in the absence of Fe2+ because there was no scattering increase for

388 

the CeO2 only system. Figure 4B (Fe2+ only system) indicates that no particles formed without

389 

CeO2 as well. In Figure 4B, no background subtraction was performed due to no increase of 20   

ACS Paragon Plus Environment

Environmental Science & Technology

  390 

intensity throughout the reaction and some signal around 0.035 Å-1 results from the beamline setup

391 

rather than actual particle contribution.

21   

ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31

Environmental Science & Technology

 

CeO2 conc. (%)

100

3 mM FeCl2 + 1 mM NaCl 10 mM NaCl

80 60 40 20 0

A 0

20

40

60

Time (h)

393 

80

100

Dissolved Ce conc. (mg/L)

392 

2.0 1.5 1.0 0.5 0.0

B 0

1

2

3

4

5

Time (h)

 

394  395 

Figure 1 

22   

ACS Paragon Plus Environment

6

 

Environmental Science & Technology

Page 24 of 31

 

B

Normalized absorbance (a.u.)

A CeO2 standard

CeO2 control 2+ CeO22++Fe Fe(II) CeO

5700

396  397 

5750 5800 Energy (eV)

Hematite

Normalized absorbance (a.u.)

Ce(III) Nitrate standard

5850

Goethite Lepidocrocite

Ferrihydrite CeO Fe(II) CeO Fe2+ 2+ 2+

7100

 

398 

Figure 2

23   

Fe(II) chloride

ACS Paragon Plus Environment

7150 7200 Energy (eV)

7250

 

Page 25 of 31

Environmental Science & Technology

 

399 

 

400 

Figure 3

401 

24   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 31

 

4

1

A

Intensity (counts)

Intensity (counts)

10

3

10

Rg = 1.72 nm

2

10

1

10

0

CeO2 + Fe(II) CeO Fe2+ CeO22 CeO

10 10

-1 6

2

0.01

B 4 2

0.1 4 2

0.01

3 4 56

2

3 4 5

2

0.1

0.01

-1

q_xy (Å )

402  403 

Figure 4

25   

ACS Paragon Plus Environment

3

4 5 6

0.1 -1 q_xy (Å )

2

3

4

 

Page 27 of 31

Environmental Science & Technology

  404 

REFERENCES

405 

1.

OECD Nanosafety at the OECD: The First Five Years 2006-2010; Organization for Economic Cooperation and Development: Paris, France, 2011.

2.

Liu, X.; Chen, G.; Su, C., Influence of Collector Surface Composition and Water Chemistry on the Deposition of Cerium Dioxide Nanoparticles: QCM-D and Column Experiment Approaches. Environ. Sci. Technol. 2012, 46, (12), 6681-6688.

3.

Migani, A.; Neyman, K. M.; Illas, F.; Bromley, S. T., Exploring Ce3+/Ce4+ cation ordering in reduced ceria nanoparticles using interionic-potential and density-functional calculations. J. Chem. Phys. 2009, 131, (6), 064701-064707.

4.

Liu, X.; Chen, G.; Su, C., Influence of Collector Surface Composition and Water Chemistry on the Deposition of Cerium Dioxide Nanoparticles: QCM-D and Column Experiment Approaches. Environ. Sci. Technol. 2012, 46, (12), 6681-6688.

5.

Evaluation of Human Health Risk from Cerium Added to Diesel Fuel; Health Effects Institute: 2001.

6.

Gantt, B.; Hoque, S.; Willis, R. D.; Fahey, K. M.; Delgado-Saborit, J. M.; Harrison, R. M.; Erdakos, G. B.; Bhave, P. V.; Zhang, K. M.; Kovalcik, K.; Pye, H. O. T., Near-Road Modeling and Measurement of Cerium-Containing Particles Generated by Nanoparticle Diesel Fuel Additive Use. Environ. Sci. Technol. 2014, 48, (18), 10607-10613.

7.

Nanomaterials EPA is Assessing. http://www.epa.gov/nanoscience/quickfinder/nanomaterials.htm

8.

Collin, B.; Oostveen, E.; Tsyusko, O. V.; Unrine, J. M., Influence of Natural Organic Matter and Surface Charge on the Toxicity and Bioaccumulation of Functionalized Ceria Nanoparticles in Caenorhabditis elegans. Environ. Sci. Technol. 2013, 48, (2), 1280-1289.

9.

Thill, A.; Zeyons, O.; Spalla, O.; Chauvat, F.; Rose, J.; Auffan, M.; Flank, A. M., Cytotoxicity of CeO2 nanoparticles for Escherichia coli. Physico-chemical insight of the cytotoxicity mechanism. Environ. Sci. Technol. 2006, 40, (19), 6151-6156.

10.

Zhang, W.; Crittenden, J.; Li, K.; Chen, Y., Attachment Efficiency of Nanoparticle Aggregation in Aqueous Dispersions: Modeling and Experimental Validation. Environ. Sci. Technol. 2012, 46, (13), 7054-7062.

11.

Keller, A. A.; Wang, H.; Zhou, D.; Lenihan, H. S.; Cherr, G.; Cardinale, B. J.; Miller, R.; Ji, Z., Stability and Aggregation of Metal Oxide Nanoparticles in Natural Aqueous Matrices. Environ. Sci. Technol. 2010, 44, (6), 1962-1967.

406  407  408  409  410  411  412  413  414  415  416  417  418  419  420  421  422  423  424  425  426  427  428  429  430  431  432  433  434  435  436  437  438 

12. Petosa, A. R.; Ohl, C.; Rajput, F.; Tufenkji, N., Mobility of nanosized cerium dioxide and polymeric capsules in quartz and loamy sands saturated with model and natural groundwaters. Water Res. 2013, 47, (15), 5889-5900. 26   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 31

  439 

13.

440  441 

Buettner, K. M.; Rinciog, C. I.; Mylon, S. E., Aggregation kinetics of cerium oxide nanoparticles in monovalent and divalent electrolytes. Colloids Surf., A 2010, 366, (1-3), 7479.

445 

14. Quik, J. T. K.; Lynch, I.; Hoecke, K. V.; Miermans, C. J. H.; Schamphelaere, K. A. C. D.; Janssen, C. R.; Dawson, K. A.; Stuart, M. A. C.; Meent, D. V. D., Effect of natural organic matter on cerium dioxide nanoparticles settling in model fresh water. Chemosphere 2010, 81, (6), 711-715.

446 

15.

Van Hoecke, K.; De Schamphelaere, K. A. C.; Van der Meeren, P.; Smagghe, G.; Janssen, C. R., Aggregation and ecotoxicity of CeO2 nanoparticles in synthetic and natural waters with variable pH, organic matter concentration and ionic strength. Environ. Pollut. 2011, 159, (4), 970-976.

16.

Barton, L. E.; Auffan, M.; Bertrand, M.; Barakat, M.; Santaella, C.; Masion, A.; Borschneck, D.; Olivi, L.; Roche, N.; Wiesner, M. R.; Bottero, J.-Y., Transformation of Pristine and Citrate-Functionalized CeO2 Nanoparticles in a Laboratory-Scale Activated Sludge Reactor. Environ. Sci. Technol. 2014, 48, (13), 7289-7296.

17.

Singh, V.; Joung, D.; Zhai, L.; Das, S.; Khondaker, S. I.; Seal, S., Graphene based materials: Past, present and future. Prog. Mater. Sci. 2011, 56, (8), 1178-1271.

18.

Yu, P.; Hayes, S. A.; O'Keefe, T. J.; O'Keefe, M. J.; Stoffer, J. O., The Phase Stability of Cerium Species in Aqueous Systems: II. The Systems. Equilibrium Considerations and Pourbaix Diagram Calculations. J. Electrochem. Soc. 2006, 153, (1), C74-C79.

442  443  444 

447  448  449  450  451  452  453  454  455  456  457  458  460 

19. Stumm, W.; Lee, G. F., Oxygenation of Ferrous Iron. Ind. Eng. Chem. 1961, 53, (2), 143146.

461 

20.

Nielsen, A. H.; Hvitved-Jacobson, T.; Vollertsen, J., Effects of pH and iron concentrations on sulfide precipitation in wastewater collection systems. Water Environ. Res. 2008, 80, (4), 380-384.

21.

Wang, Y.; Tng, K. H.; Wu, H.; Leslie, G.; Waite, T. D., Removal of phosphorus from wastewaters using ferrous salts - A pilot scale membrane bioreactor study. Water Res. 2014, 57, 140-150.

459 

462  463  464  465  466 

469 

22. Ivanov, V.; Kuang, S.; Stabnikov, V.; Guo, C., The removal of phosphorus from reject water in a municipal wastewater treatment plant using iron ore. J. Chem. Technol. Biot. 2009, 84, (1), 78-82.

470 

23.

Park, D.; Kim, Y. M.; Lee, D. S.; Park, J. M., Chemical treatment for treating cyanidescontaining effluent from biological cokes wastewater treatment process. Chem. Eng. J. 2008, 143, (1–3), 141-146.

24.

Xu, L.; Wang, J., Magnetic Nanoscaled Fe3O4/CeO2 Composite as an Efficient Fenton-Like Heterogeneous Catalyst for Degradation of 4-Chlorophenol. Environ. Sci. Technol. 2012, 46, (18), 10145-10153.

467  468 

471  472  473  474  475 

27   

ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

  476 

25.

Schreiber, H. D.; Lauer Jr, H. V.; Thanyasiri, T., Oxidation-reduction equilibria of iron and cerium in silicate glasses: Individual redox potentials and mutual interactions. J. Non-Cryst. Solids 1980, 38-39, (PART 2), 785-790.

26.

Limbach, L. K.; Bereiter, R.; Muller, E.; Krebs, R.; Galli, R.; Stark, W. J., Removal of Oxide Nanoparticles in a Model Wastewater Treatment Plant: Influence of Agglomeration and Surfactants on Clearing Efficiency. Environ. Sci. Technol. 2008, 42, (15), 5828-5833.

27.

Dahle, J.; Arai, Y., Effects of Ce(III) and CeO2 Nanoparticles on Soil-Denitrification Kinetics. Arch. Environ. Contam. Toxicol. 2014, 1-9.

477  478  479  480  481  482  483  485 

28. Bard, A. J.; Faulkner, L. R., Electrochemical Methods: Fundamentals and Applications John Wiley & Sons, INC.: New York, 2001.

486 

29.

Wilson, N.; Webster-Brown, J., The fate of antimony in a major lowland river system, the Waikato River, New Zealand. Appl. Geochem. 2009, 24, (12), 2283-2292.

30.

Bigham, J. M.; Schwertmann, U.; Carlson, L.; Murad, E., A poorly crystallized oxyhydroxysulfate of iron formed by bacterial oxidation of Fe(II) in acid mine waters. Geochim. Cosmochim. Ac. 1990, 54, (10), 2743-2758.

484 

487  488  489  490  492 

31. Stumm, W.; Morgan, J., Aquatic Chemistry: Chemical Equilibria and Rate in Natural Waters. 3rd ed.; Wiley: New York, 1996.

493 

32.

Liu, X.; Gong, W.; Liu, L., Treatment of sulfate-rich and low pH wastewater by sulfate reducing bacteria with iron shavings in a laboratory. Water Sci. Technol. 2014, 69, (3), 595600.

33.

Rousch, J. M.; Simmons, T. W.; Kerans, B. L.; Smith, B. P., Relative acute effects of low pH and high iron on the hatching and survival of the water mite (Arrenurus manubriator) and the aquatic insect (Chironomus riparius). Environ. Toxicol. Chem. 1997, 16, (10), 2144-2150.

34.

Gerhardt, A., Effects of subacute doses of iron (Fe) on Leptophlebia marginata (Insecta: Ephemeroptera). Freshwater Biol. 1992, 27, (1), 79-84.

35.

Díez, S.; Noonan, G. O.; MacFarlane, J. K.; Gschwend, P. M., Ferrous iron oxidation rates in the pycnocline of a permanently stratified lake. Chemosphere 2007, 66, (8), 1561-1570.

491 

494  495  496  497  498  499  500  501  502  504 

36. Dahle, J. T.; Livi, K.; Arai, Y., Effects of pH and phosphate on CeO2 nanoparticle dissolution. Chemosphere 2015, 119, 1365-1371.

505 

37.

Hoecke, K. V.; Quik, J. T. K.; Mankiewicz-Boczek, J.; Schamphelaere, K. A. C. D.; Elsaesser, A.; Meeren, P. V. d.; Barnes, C.; McKerr, G.; Howard, C. V.; Meent, D. V. D.; Rydzynski, K.; Dawson, K. A.; Salvati, A.; Lesniak, A.; Lynch, I.; Silversmit, G.; Samber, B. D.; Vincze, L.; Janssen, C. R., Fate and Effects of CeO2 Nanoparticles in Aquatic Ecotoxicity Tests. Environ. Sci. Technol. 2009, 43, (12), 4537-4546.

38.

Rico, C. M.; Morales, M. I.; McCreary, R.; Castillo-Michel, H.; Barrios, A. C.; Hong, J.; Tafoya, A.; Lee, W.-Y.; Varela-Ramirez, A.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L.,

503 

506  507  508  509  510  511 

28   

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

 

Cerium Oxide Nanoparticles Modify the Antioxidative Stress Enzyme Activities and Macromolecule Composition in Rice Seedlings. Environ. Sci. Technol. 2013, 47, (24), 14110-14118.

512  513  514  515 

39.

Rogers, N. J.; Franklin, N. M.; Apte, S. C.; Batley, G. E.; Angel, B. M.; Lead, J. R.; Baalousha, M., Physico-chemical behaviour and algal toxicity of nanoparticulate CeO2 in freshwater. Environ. Chem. 2010, 7, (1), 50-60.

40.

Rico, C. M.; Hong, J.; Morales, M. I.; Zhao, L.; Barrios, A. C.; Zhang, J.-Y.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L., Effect of Cerium Oxide Nanoparticles on Rice: A Study Involving the Antioxidant Defense System and In Vivo Fluorescence Imaging. Environ. Sci. Technol. 2013, 47, (11), 5635-5642.

41.

Baalousha, M.; Ju-Nam, Y.; Cole, P. A.; Gaiser, B.; Fernandes, T. F.; Hriljac, J. A.; Jepson, M. A.; Stone, V.; Tyler, C. R.; Lead, J. R., Characterization of cerium oxide nanoparticles— Part 1: Size measurements. Environ. Toxicol. Chem. 2012, 31, (5), 983-993.

42.

Baalousha, M.; Le Coustumer, P.; Jones, I.; Lead, J. R., Characterisation of structural and surface speciation of representative commercially available cerium oxide nanoparticles. Environ. Chem. 2010, 7, (4), 377-385.

43.

Ray, J. R.; Lee, B.; Baltrusaitis, J.; Jun, Y.-S., Formation of Iron(III) (Hydr)oxides on Polyaspartate- and Alginate-Coated Substrates: Effects of Coating Hydrophilicity and Functional Group. Environ. Sci. Technol. 2012, 46, (24), 13167-13175.

44.

Hu, Y.; Lee, B.; Bell, C.; Jun, Y.-S., Environmentally Abundant Anions Influence the Nucleation, Growth, Ostwald Ripening, and Aggregation of Hydrous Fe(III) Oxides. Langmuir 2012, 28, (20), 7737-7746.

45.

Jun, Y.-S.; Lee, B.; Waychunas, G. A., In Situ Observations of Nanoparticle Early Development Kinetics at Mineral−Water Interfaces. Environ. Sci. Technol. 2010, 44, (21), 8182-8189.

46.

De Faria, L. A.; Trasatti, S., The Point of Zero Charge of CeO2. J. Colloid Interf. Sci. 1994, 167, (2), 352-357.

47.

Kosmulski, M., The pH-Dependent Surface Charging and the Points of Zero Charge. J. Colloid Interf. Sci. 2002, 253, (1), 77-87.

48.

Kosmulski, M., The pH-dependent surface charging and points of zero charge: V. Update. J. Colloid Interf. Sci. 2011, 353, (1), 1-15.

49.

Cornelis, G.; Ryan, B.; McLaughlin, M. J.; Kirby, J. K.; Beak, D.; Chittleborough, D., Solubility and Batch Retention of CeO2 Nanoparticles in Soils. Environ. Sci. Technol. 2011, 45, (7), 2777-2782.

516  517  518  519  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 

50. Hunter, R. J., Foundations of Colloid Science. second ed.; Clarendon Press: Oxford, U.K., 2001; Vol. 210, p 125-125. 29   

ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

 

549 

51. Faure, B.; Salazar-Alvarez, G.; Bergström, L., Hamaker constants of iron oxide nanoparticles. Langmuir 2011, 27, (14), 8659-8664.

550 

52.

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, 1-9.

53.

Azimi, G.; Dhiman, R.; Kwon, H. M.; Paxson, A. T.; Varanasi, K. K., Hydrophobicity of rare-earth oxide ceramics. Nature Materials 2013, 12, (4), 315-320.

548 

551  552  553  555 

54. Grassian, V. H.; Hamers, R. J. Workshop report on "Nanomaterials and the Environment: The Chemistry and Materials Perspective"; Arlington, 2011.

556 

55.

Donald, L. M.; Katherine, W.-D., Redox Chemistry and Natural Organic Matter (NOM): Geochemists? Dream, Analytical Chemists? Nightmare. In Aquatic Redox Chemistry, American Chemical Society: 2011; Vol. 1071, pp 85-111.

56.

Neil, C. W.; Lee, B.; Jun, Y.-S., Different Arsenate and Phosphate Incorporation Effects on the Nucleation and Growth of Iron(III) (Hydr)oxides on Quartz. Environ. Sci. Technol. 2014, 48, (20), 11883-11891.

554 

557  558  559  560  561  562  563 

57. Rose, A. L., Effect of Dissolved Natural Organic Matter on the Kinetics of Ferrous Iron Oxygenation in Seawater. Environ. Sci. Technol. 2003, 37, (21), 4877-4886.

564  565 

30   

ACS Paragon Plus Environment

Enhanced Colloidal Stability of CeO2 Nanoparticles by Ferrous Ions: Adsorption, Redox Reaction, and Surface Precipitation.

Due to the toxicity of cerium oxide (CeO2) nanoparticles (NPs), a better understanding of the redox reaction-induced surface property changes of CeO2 ...
470KB Sizes 0 Downloads 8 Views