Visualizing Arsenate Reactions and Encapsulation in a Single Zero

Jan 12, 2017 - The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.est.6b04315. ..... Arsenic r...
0 downloads 16 Views 3MB Size
Subscriber access provided by Fudan University

Article

Visualizing Arsenate Reactions and Encapsulation in a Single Zero-Valent Iron Nanoparticle Lan Ling, and Wei-Xian Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b04315 • Publication Date (Web): 12 Jan 2017 Downloaded from http://pubs.acs.org on January 12, 2017

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 25

Environmental Science & Technology

Visualizing Arsenate Reactions and Encapsulation in a Single Zero-Valent Iron Nanoparticle

Lan Ling and Wei-xian Zhang* State Key Laboratory for Pollution Control School of Environmental Science and Engineering Tongji University 1239 Siping Road Shanghai, China 200092

*To whom correspondence should be addressed. Email: [email protected], Phone: +8621-15221378401, Fax: +86-21-6598 0041.

ACS Paragon Plus Environment

1

Environmental Science & Technology

1

ABSTRACT

2

A nanostructure-based mechanism is presented on the enrichment, separation and immobilization

3

of arsenic with nanoscale Zero-Valent Iron (nZVI). The As-Fe reactions are studied with

4

Spherical Aberration Corrected Scanning Transmission Electron Microscopy (Cs-STEM). Near

5

atomic-resolution (99.9%) for 30 minutes prior to

67

the experiment. The iron nanoparticles for the STEM analysis were collected from batch reactors

68

containing 10 ppm As(V) with pre-determined doses of iron nanoparticles (e.g., 0.2–5 g/L ).

69

Batch reactors were sealed with screw caps and mixed on a shaker table at ambient temperature

70

(approximately 22 °C). The collected nZVI particles were rinsed twice with high-purity nitrogen

71

purged ethanol and stored in a nitrogen-purged sample vial for solid reactions before the STEM

72

analysis.

73

Methods. Electron microscopy characterizations are performed on a FEI Titan™ G2 80-300,

74

which is a STEM (scanning transmission electron microscope) integrated with corrected electron

75

optics. The system is equipped with a FEI X-FEG high brightness Schottky field emission source

76

(brightness 1.8 × 109 A/cm² srad at 200 kV, beam current 1,300 pA in 0.2 nm spot), a CEOS

77

hexapole/transfer doublet design aberration corrector, High-Angle Annular Dark-Field

78

(HAADF) and Bright-Field (BF) detectors, an Electron Energy-Loss Spectrometer (EELS)

79

(Gatan, UK), Energy Filtered TEM (EFTEM, Quantum SE/963 200 kV) and the FEI

80

ChemiSTEM system.19, 20

81

3D-ET and XEDS analysis are achieved with the FEI SuperX operated at a beam current of ~0.8

82

nA and a probe size of ~0.2 nm at 200 KV. Each spectrum image of 512×512 pixels is acquired

ACS Paragon Plus Environment

6

Page 6 of 25

Page 7 of 25

Environmental Science & Technology

83

over a 10 min duration. The tilt series of XEDS images are collected from −64° to +64° with a 2°

84

tilt increment. Data processing is achieved using the Inspect3D software (FEI) with a built-in

85

algorithm of simultaneous iterative reconstruction technique (SIRT). 3D dataset of segmentation

86

is then realized with the Amira software.23

87

RESULTS AND DISCUSSION

88

nZVI, an engineered nanomaterial has been increasingly used in environmental remediation and

89

hazardous waste treatment.13−17, 24−27 For example, numerous reports have documented its

90

effectiveness in detoxifying halogenated hydrocarbons in groundwater.28−32 Figure 1 presents

91

images of nZVI (more are available in the Supporting Information, Figure S1). A fresh nZVI

92

particle is spherical with diameter typically in the range of 20–100 nm. The BF and HAADF

93

images (Figure 1a and 1b) provide complementary views of the core-shell structure. As shown in

94

the high-resolution TEM images (Figure. 1c and 1d), nZVI has a classical core-shell structure,

95

that is, an individual nanoparticle has local or nanoscale crystalline structures, comprises of an

96

Fe(0) core and separates from other nanoparticles by a thin (2−5 nm) interfacial layer of iron

97

oxides. Short-range lattice arrangements and some defective structures on the shell surface are

98

also apparent from the high-resolution TEM micrographs (Figure 1d). The extremely thin

99

dimension, disordered and defective nature of the oxide layer offer efficient electron passage and

100

high reactivity via tunneling effects and defect channeling. 33 The core-shell structure bestows

101

the nanoparticles with the reductive characteristics of Fe(0) and coordinative properties of iron

102

oxides, offers a unique and powerful combination for the separation and sequestration of arsenic

103

and heavy metal ions.13-17, 21, 22

ACS Paragon Plus Environment

7

Environmental Science & Technology

104

Figure 1. STEM images of clean nanoscale zero-valent iron (nZVI). (a) Bright Field (BF); (b)

105

High-Angle Annular Dark-Field (HAADF). (c) and (d) are high-resolution TEM images.

106

Previous work has showed that nZVI can quickly remove both As(III) and As(V) from water and

107

also possess large capacity (>200 mg As/g Fe) for arsenic encapsulation. 16 The reactions of

108

arsenate with nZVI can be manifested with the distributions of oxygen, iron and arsenic in spent

109

nZVI particles. Figure 2 presents the XEDS elemental mappings of As(Kα), Fe(Kα), O(Kα),

110

XEDS line profiles, color overlays and spectra of selected areas acquired from a single nZVI

111

nanoparticle reacted with arsenate. The nanoparticles were from a reactor after 48-hour reactions

112

with 10 mg/L arsenate. After the reactions, the nZVI nanoparticles is slightly distorted due to

ACS Paragon Plus Environment

8

Page 8 of 25

Page 9 of 25

Environmental Science & Technology

113

iron dissolution and precipitation of iron hydroxides (Figure 2a). The Fe mapping (Figure 2b)

114

shows footprint of iron with a dense metallic core. From Figure 2d, the ring of oxygen matches

115

the lower intensity area of iron, reconfirming that the surface layer is rich in oxygen. The arsenic

116

mapping (Figure 2c) on the other hand illustrates a distinctive ring of arsenic just on top of the

117

Fe(0) core surface. This ring of arsenic is intense and locates exactly between the Fe(0) core and

118

iron oxide shell, resulting from the reduction of As(V) and accumulation of As(0). Previous work

119

with XPS showed that arsenic accumulated inside the particle is mostly the reduced As(0) and

120

As(III),15, 16 we now prove that arsenic can fully penetrate the surface oxide layer.

121

More detailed information on the intraparticle distributions of oxygen, iron and arsenic in a

122

reacted nZVI particle is obtained from the XEDS line profiles (Figure 2a, Figure S2). The

123

arsenic profiles, peak positions and heights offer valuable information on the rates of

124

intraparticle mass transfer and reactions, also on the speciation [i.e., As(V), As(III) and As(0)]

125

and total amount of arsenic in the nanoparticle. The iron abundance ascends rapidly within first

126

4–5 nm and reaches a plateau at the depth of 8−9 nm under the particle surface. The Fe scan

127

shows a sharp valley matching the location of the arsenic ring. The oxygen profile climaxes at

128

about 3 nm from the water-nanoparticle interface just outside the Fe(0) core. This is consistent

129

with the core-shell model that the outer surface iron is more oxidized with oxygen-rich oxides

130

while the inner part is less oxidized with more Fe(II). From XEDS quantifications (Figure 2g-2j

131

and Figure S2), the arsenic abundance in this thin layer is estimated at 19.44 wt% and accounts

132

for more than half of the arsenic within the whole particle.

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 25

133

Figure 2. High-resolution STEM-XEDS elemental mappings and XEDS quantifications of

134

the Fe-As reactions. (a) HAADF image and XEDS line profiles, (b) Fe, (c) As, (d) O, (e)

135

Fe+As, (f) Fe+As+O, (g-i) XEDS quantifications, and (j) XEDS spectra. Signals were collected

136

from an iron nanoparticle after 48-hour reactions with 10 mg/L arsenate.

ACS Paragon Plus Environment

10

Page 11 of 25

Environmental Science & Technology

137

Figure 3. 3D chemical mapping of spent nanoscale zero-valent iron after reactions with

138

arsenate. (a) 3D elemental distributions of Fe (blue), O (green) and As (red). (b-j) are a series of

139

mappings acquired from spent nanoscale zero-valent iron reacted with arsenate in the tilt range

140

of +64° to -64°. The sample was from a batch reactor containing 10 mg As(V)/L and 0.5 g/L

141

nZVI after 48 hours. A corresponding video of the 3D tomography can be found in the

142

Supporting Information (Movie S1).

143

The nano-structure based encapsulation can be visualized via electron tomography, which can

144

generate high-resolution 3D view of pollutant distribution in a single nanoparticle. To achieve

145

3D visualization, a series of projection (2D) images of a single nanoparticle are acquired from a

146

wide range of angles by tilting the sample relative to the probe. A full 3D visualization is then

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 25

147

established using a cross-correlation algorithm for image shift, tilt alignments and simultaneous

148

iterative reconstruction (FEI Inspect3DTM, details are provided in Supporting Information). 34-39

149

A video of 3D electron tomography showing arsenic in a single nZVI nanoparticle is provided in

150

the Supporting Information (Movie S1). Figure 3 presents a series of projection images from

151

selected angles. The XEDS tilt series and the resulting 3D tomography (Movie S1) discover a

152

distinct feature of the As-nZVI reactions, that is, a thin and continuous layer of arsenic (in red

153

color, ~23 Å in thickness, Figure 3) is deposited beneath the iron oxide surface layer and just on

154

top of the Fe(0) core after 48-hour reactions with 10 mg/L As(V) (Na2HAsO4). The distribution

155

sphere of arsenic is larger than that of the iron core but smaller than the distribution ring of

156

oxygen. This technique generates extraordinary details and insights unattainable from

157

conventional 2D XEDS scans as the state-of-the-art Cs-STEM can obtain atomic resolution

158

structural imaging as well as high-resolution (~ 1 nm3) 3D elemental mapping. Although

159

previous work has suggested that arsenic can be reduced and immobilized by zero-valent iron,

160

this work, for the first time visualizes the formation of a continuous layer of arsenic within nZVI

161

nanoparticles. This clearly points to a large capacity and stability of nZVI for arsenic

162

encapsulation.

163

In direct contrast, configurations of the XEDS mapping and electron tomography on the

164

arsenate-ferric hydroxide (FeOOH) reactions (Figure 4 and 5) show that the arsenate is attracted

165

rather loosely to the surface, that is, the arsenic atoms are to a large degree detached from the

166

surface. The cyan color in Figure 4 and 5 reflects the nature of blue (iron) (Figure 4a) and green

167

(oxygen) (Figure 4b) mixtures, suggesting that the particle body consists of Fe and O with As

168

floating on top and/or near the surface. There is no intraparticle diffusion or accumulation of

169

arsenic inside the oxide nanoparticle. The atomic ratios of oxygen, iron and arsenic from XEDS

ACS Paragon Plus Environment

12

Page 13 of 25

Environmental Science & Technology

170 171

Figure 4. High-resolution STEM-XEDS elemental mappings and XEDS line profiles of

172

ferric hydroxide (FeOOH)-arsenate reactions. (a) Fe, (b) O, (c) As, (d) Fe+As+O, (e)

173

HAADF image, (f) XEDS line profiles (quantifications and spectra are provided in Supporting

174

Information, Figure S3). The sample was from a batch reactor with 10 mg/L As(V) and 0.5 g/L

175

Fe. Reaction time was 48 hrs.

176

quantification implies no arsenate reduction (Figure S3).11, 40 XEDS quantitative analysis shows

177

that weight percentage of arsenic is in the range of 2.48% to 2.97% on the FeOOH surface and

178

average 1.75% for the whole particle. Furthermore, arsenic distributes unevenly on the FeOOH

179

surface as shown in the 3D tomography (Movie S2 and Figure S3). The 3D tomography thus

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 25

180 181

Figure 5. 3D chemical mapping on arsenate-goethite (FeOOH) reactions. (a) 3D elemental

182

distribution of Fe (blue), O (green) and As (red). (b-j) are a series of mappings acquired from

183

spent FeOOH reacted with arsenate in the tilt range of +64° to -64°. The sample was from a

184

batch reactor with 10 mg As(V)/L and 0.5 g/L Fe after 48 hours. A corresponding video of the 3D

185

tomography is provided in the Supporting Information (Movie S2).

186

provides direct confirmation of a widely held postulation that arsenic is attracted to the goethite

187

surface by forming surface complexes.11, 40 It further confirms that arsenic covering on the oxide

188

surface is far from continuous and the amount of arsenic on iron hydroxide (FeOOH) is much

189

less than the continuous layer of arsenic in nZVI.

ACS Paragon Plus Environment

14

Page 15 of 25

Environmental Science & Technology

190

The observed pattern of arsenic in the nZVI nanoparticle reflects the net result of the arsenic

191

diffusion and reactions within the unique nanostructure. As presented earlier, nZVI consists of

192

two nanomaterials. The core is zero-valent or metallic iron, which serves as a reductant for

193

arsenite and arsenate reduction. The surface layer of iron oxides is formed from spontaneous

194

oxidation of Fe(0) and can be represented with an average stoichiometry of iron hydroxide

195

(FeOOH).15,16 Furthermore, the nZVI shell consists of a mixed Fe(II)/Fe(III) phase close to the

196

Fe(0) interface and a predominantly Fe(III) at the outer surface of nZVI, resulting in a clear

197

phase junction. Thus nZVI possesses the combinatorial properties of zero-valent iron and iron

198

oxides. The iron oxide layer is polar and/or charged in water, offers electrostatic attractions and

199

sorption sites to both cations and anions in water while the core supplies electrons or the

200

reducing capability for further diffusion, reduction and deposition of arsenic and heavy metals

201

such as Ni and Cu. The intraparticle diffusion of arsenic is highly favored since the core region

202

has the highest concentration of Fe(0) for the arsenic reduction and immobilization. The grain

203

boundary provides additional support to the observed arsenic diffusion in nZVI particle, that is,

204

arsenic atoms diffuse preferably along the defective, high energy, and non-equilibrium

205

polycrystalline grain boundary of iron oxides 41,

206

patterns presented in Figure 2 and 3 are also consistent fully with results from previous XPS

207

characterizations.15, 16

208

Atomic-resolution STEM images of nZVI and a conceptual model on the sorption, precipitation

209

and reduction of arsenic by nZVI are illustrated in Figure 6. As shown in Figure 6a and 6b, the

210

reacted nZVI particle retains the classical core-shell structure. Furthermore, short-range lattice

211

arrangements in the shell area are also apparent from the high-resolution STEM micrographs

212

with individual iron atoms clearly identified under the HAADF model. nZVI formed from

42

The iron, arsenic and oxygen distribution

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 25

213

sodium borohydride reduction of ferric chloride has a body-centered cubic structure (bcc) in the

214

space group Oh9

215

orientation in unprocessed HAADF image. Higher magnification image (inset in Figure 6a)

216

provides further evidence that a dense core displayed along (111) zone axis. The bright columns

217

are the iron atoms and show subtle details about the shape and faceting of the nanocrystals. The

218

intensity line profile (in red color, Figure 6a) measured across the periodical arrangements

219

clearly signals individual atoms (highlighted with the blue circles). A HAADF intensity peak

220

matches precisely to an atomic column. This is further evidenced by the corresponding electron

221

diffraction patterns.

222

The high-resolution STEM images of nZVI provide remarkable details on the atomic

223

arrangements within the core-shell structure and complement previous work of XPS and EXAFS. 15,

224

16

225

of iron structure model in the (111) lattice plane with the lattice constant a of 2.405 Å. The

226

interatomic spacing along this direction is measured at 1.70 Å, in good agreement with the

227

calculated interatomic spacing along direction, that is,

228

XEDS, this surface layer is enriched with arsenic as a result of the reductive deposition. The

229

observed difference of peak intensity is likely caused by the substitution or replacement of iron

230

for the arsenic atoms. 43, 44

231

As shown in the conceptual model (Figure 6c), formation of the thin layer of arsenic between the

232

Fe(0) core and iron oxide shell offers the most direct evidence on the nano-encapsulation and

233

proves that the outer layer, especially the Fe(0)/oxide interface is the frontline for the As-Fe

234

reactions. Evidence obtained in this work has important implications on applications of iron

235

nanoparticles for arsenic separation and immobilization. The total capacity of nZVI for arsenic

15, 33

Figure 6a shows the structure of spent nZVI predominantly along (111)

In the Annular Dark Field (ADF) image (Figure 6b), the seven red points illustrate a unit-cell

ACS Paragon Plus Environment

16

 √

=1.70 Å. As observed from

Page 17 of 25

Environmental Science & Technology

236

removal is the sum of surface sorption and immobilization by the Fe(0) reduction. Since the

237

driving force for arsenic diffusion in nZVI is chemical reduction and successive encapsulation,

238

the chemical reduction of arsenic is much greater than that of the surface sorption with capacity

239

for arsenic immobilization governed primarily by and proportional to the amount of Fe(0).

240

Laboratory batch experiments have confirmed that arsenic removal capacity can reach as high as

241

200 mg As/g Fe.15,16

242 243

Figure 6. Atomic-resolution STEM images of spent nanoscale zero-valent iron (nZVI) and

244

a conceptual model on arsenate-nZVI reactions. (a) Unprocessed High-angle Annular Dark-

245

field (HAADF) image of a portion of a nanoparticle, inset is a higher-magnification view of the

246

core area. The HAADF intensity line profile (in red color) measured across the periodical

247

arrangements signals individual atoms (highlighted with the solid blue circles). (b) Annular

248

Dark-Field (ADF) image shows that Fe atoms in the core area are predominantly in the body-

249

centered cubic (111) lattice plane. (c) A conceptual model on the sorption, precipitation and

250

reduction of toxic metals with nZVI.

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 25

251

Additionally, the reduced arsenic, As(0) is very stable in the presence of Fe(0). This is critical for

252

the immobilization and stabilization of toxic heavy metals such as mercury, arsenic, and

253

radioactive materials.22 Our laboratory experiments showed the nZVI-bound arsenic is much less

254

sensitive to solution pH changes. For example, optimal pH for arsenic sorption to iron oxides is

255

less than 4 with little sorption observed above pH 8. With nZVI, high efficiency (>90%) can be

256

maintained over a wide range of solution pH (~3−10) (Figure S4). The presence of Fe(0) thus

257

serves as an enduring shield or long-term stabilizer, which inhibits oxidation, dissolution and

258

leaching of pollutants.

259

Reactions with solid surfaces are the principal mechanisms controlling speciation, distribution,

260

transformation, bioavailability, ultimate fate and toxicity of environmental contaminants, thus

261

are fundamental in environmental, geological and colloidal chemistry.

262

knowledge on the surface reactions of most environmental contaminants is still limited, largely

263

impeded by the lack of effective tools for direct and in situ observations. We demonstrate that for

264

the first time, reactions of arsenic with iron nanoparticles can be “seen” via high-resolution 3D

265

electron tomography at near atomic resolution. This technique offers new opportunities in the

266

investigation of pollutant transformation and detoxification by generating simultaneous high-

267

resolution structural characterization and high-accuracy identification and quantification of

268

contaminants in the solid phase. 34, 35 With further improvements and refinements, this technique

269

will undoubtedly expand our understanding on the intraparticle mass transfer, catalysis,

270

transformation and detoxification of environmental toxins in the heterogeneous environmental

271

media.

272

ASSOCIATED CONTENT

ACS Paragon Plus Environment

18

45, 46

Nevertheless, our

Page 19 of 25

Environmental Science & Technology

273

Supporting Information

274

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

275

Movies S1-S2, Figures S1-S4, and more experimental details are available in the Supporting

276

Information.

277

ACKNOWLEDGEMENTS

278

Financial support from the National Natural Science Foundation of China (NSFC Grants

279

21307094, 21277102 and 21677107), the Collaborative Innovation Center for Regional

280

Environmental Quality and the State Key Laboratory for Pollution Control and Resource Reuse

281

are acknowledged.

282

AUTHOR INFORMATION

283

Corresponding author

284

Email: [email protected], Phone: +86-21-15221378401, Fax: +86-21-6598 0041.

285

Notes

286

The authors declare no competing financial interests.

287

REFERENCES

288

1. Geen van, A.; Bostick, B. C.; Trang, P. T. K.; Lan, V. M.; Mai, N. N.; Manh, P. D.; Viet, P. H.;

289

Radloff, K.; Aziz, Z.; Mey, J. L.; Stahl, M. O.; Harvey, C. F.; Oates, P.; Weinman, B.; Stengel,

290

C.; Frei, F.; Kipfer, R.; Berg, M. Retardation of arsenic transport through a Pleistocene aquifer.

291

Nature 2013, 501, 204–208.

292

2. Fendorf, S.; Michael, H. A.; Geen van, A. Spatial and temporal variations of groundwater

293

arsenic in South and Southeast Asia. Science 2010, 328, 1123–1127.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 25

294

3. Nickson1, R.; McArthur, J.; Burgess, W.; Ahmed, K. M.; Ravenscroft, P.; Rahmanñ, M.

295

Arsenic poisoning of Bangladesh ground water. Nature 1998, 395, 338.

296

4. Rodríguez-Lado, L.; Sun, G.; Berg, M.; Zhang, Q.; Xue, H.; Zheng, Q.; Johnson, C. A.

297

Groundwater arsenic contamination throughout China. Science 2013, 341, 866–868.

298

5. Radloff, K. A.; Zheng, Y.; Michael, H. A.; Stute, M.; Bostick, B. C.; Mihajlov, I.; Bounds, M.;

299

Huq, M. R.; Choudhury, I.; Rahman, M. W.; Schlosser, P.; Ahmed, K. M.; Geen van A. Arsenic

300

migration to deep groundwater in Bangladesh influenced by adsorption and water demand.

301

Nature Geosci. 2011, 4, 793–798.

302

6. Stuckey, J. W.; Schaefer, M. V.; Kocar, B. D.; Benner, S. G.; Fendorf, S. Arsenic release

303

metabolically limited to permanently water-saturated soil in Mekong Delta. Nature Geosci. 2016,

304

9, 70–79.

305

7. Senn, D. B.; Hemond, H. F. Nitrate controls on iron and arsenic in an urban lake. Science 2002,

306

296, 2373−2376.

307

8. Polizzotto, M. L.; Kocar, B. D.; Benner, S. G.; Sampson, M.; Fendorf, S. Near-surface

308

wetland sediments as a source of arsenic release to ground water in Asia. Nature 2008, 454,

309

505−509.

310

9. Yavuz, C. T.; Mayo, J. T.; Yu, W. W.; Prakash, A.; Falkner, J. C.; Yean, S.; Cong, L.;

311

Shipley, H. J.; Kan, A. T.; Tomson, M. B.; Natelson, D.; Colvin, V. L. Low-field magnetic

312

separation of monodisperse Fe3O4 Nanocrystals. Science 2006, 314, 964−967.

313

10. Roberts, L. C.; Hug, S. J.; Ruettimann, T.; Billah, M.; Khan, A. W.; Rahman, M. T. Arsenic

314

removal with iron(II) and iron(III) waters with high silicate and phosphate concentrations.

315

Environ. Sci. Technol. 2004, 38, 307−315.

ACS Paragon Plus Environment

20

Page 21 of 25

Environmental Science & Technology

316

11. Ona-Nguema, G.; Morin, G.; Wang, Y.; Foster, A. L.; Juillot, F.; Calas, G.; Brown Jr, G. E.

317

XANES evidence for rapid arsenic(III) oxidation at magnetite and ferrihydrite surfaces by

318

dissolved O2 via Fe2+-mediated reactions. Environ. Sci. Technol. 2010, 44, 5416–5422.

319

12. Su, C.; Puls, R. W. Arsenate and arsenite removal by zerovalent iron: kinetics, redox

320

transformation, and implications for in situ groundwater remediation. Environ. Sci. Technol.

321

2001, 35, 1487–1492.

322

13. Kanel, S. R.; Manning, B.; Charlet, L.; Choi, H. Removal of arsenic(III) from groundwater

323

by nanoscale zero-valent iron. Environ. Sci. Technol. 2005, 39, 1291–1298.

324

14. Leupin, O. X.; Hug, S. J. Oxidation and removal of arsenic (III) from aerated groundwater by

325

filtration through sand and zero-valent iron. Water Res. 2005, 39, 1729−1740.

326

15. Yan, W. L.; Vasic, R.; Frenkel, A. I.; Koel B. E. Intraparticle reduction of arsenite (As(III))

327

by nanoscale zerovalent iron (nZVI) investigated with in situ X-ray absorption spectroscopy.

328

Environ. Sci. Technol. 2012, 46, 7018−7026.

329

16. Yan, W. L.; Ramos, M. A. V.; Koel, B. E.; Zhang, W. X. As(III) sequestration by iron

330

nanoparticles: study of solid-phase redox transformations with X-ray photoelectron spectroscopy.

331

J. Phys. Chem. C 2012, 116, 5303–5311.

332

17. Ling, L.; Zhang, W. X. Environ. Sequestration of arsenate in zero-valent iron nanoparticles:

333

visualization of intraparticle reactions at angstrom resolution. Sci. Technol. Lett. 2014, 1,

334

305−309.

335

18. Watanabe, M.; Ackland, D. W.; Burrows, A.; Kiely, C. J.; Williams, D. B.; Krivanek, O.L.;

336

Dellby, N.; Murfitt, M.F.; Szilagyi, Z. Improvements in the X-ray analytical capabilities of a

337

scanning transmission electron microscope by spherical aberration correction. Microsc.

338

Microanal. 2006, 12, 515–526.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 25

339

19. Kim, Y.; Tao, R.; Klie, R. F.; Seidman, D. N. Direct Atomic-Scale Imaging of Hydrogen and

340

Oxygen Interstitials in Pure Niobium Using Atom-Probe Tomography and Aberration-Corrected

341

Scanning Transmission Electron Microscopy. ACS Nano 2013, 7, 732–739.

342

20. Jiang, Y.; Wang, Y.; Sagendorf, J.; West, D.; Kou, X.; Wei, X.; He, L.; Wang, K. L.; Zhang,

343

S.; Zhang, Z. Direct atom-by-atom chemical identification of nanostructures and defects of

344

topological insulators. Nano Lett. 2013, 13, 2851–2856.

345

21. Ling, L.; Pan, B.; Zhang, W. X. Removal of selenium from water with nanoscale zero-valent

346

iron: mechanisms of intraparticle reduction of Se(IV). Water. Res. 2015, 71, 274−281.

347

22. Ling, L.; Zhang, W. X. Enrichment and encapsulation of uranium with iron nanoparticle. J.

348

Am. Chem. Soc. 2015, 137, 2788−2791.

349

23. Ling, L.; Zhang, W. X. Environ. Reactions of Nanoscale Zero-Valent Iron with Ni(II):

350

Three-Dimensional Tomography of the “Hollow Out” Effect in a Single Nanoparticle. Environ.

351

Sci. Technol. Lett. 2014, 1, 209−213.

352

24. Feitz, A. J.; Joo, S. H.; Guan, J.; Sun, Q.; Sedlak, D. L.; Waite, T. D. Oxidative

353

transformation of contaminants using colloidal zero-valent iron. Colloids and Surfaces A:

354

Physicochem. Eng. Aspects. 2005, 265, 88–94.

355

25. Joo, S. H.; Feitz, A. J.; Sedlak, D. L.; Waite, T D. Quantification of the oxidizing capacity of

356

nanoparticulate zero-valent iron. Environ. Sci. Technol. 2005, 39, 1263–1268.

357

26. Bishop, E. J.; Fowler, D. E.; Skluzacek, J. M.; Seibel, E. M.; Mallouk, T. E. Anionic

358

homopolymers efficiently target zero valent iron particles to hydrophobic contaminants in sand

359

columns. Environ. Sci. Technol. 2010, 44, 9069−9074.

ACS Paragon Plus Environment

22

Page 23 of 25

Environmental Science & Technology

360

27. Kim, J. Y.; Lee, C.; Love, D. C.; Sedlak, D. L.; Yoon, J.; Nelson, K. L. Inactivation of MS2

361

Coliphage by Ferrous Ion and Zero-Valent Iron Nanoparticles. Environ. Sci. Technol. 2011, 45,

362

6978–6984.

363

28. Schrick. B.; Hydutsky. B. W.; Blough. J. L.; Mallouk. T. E. Delivery vehicles for zerovalent

364

metal nanoparticles in soil and groundwater. Chem. Mater. 2004, 16, 2187–2193.

365

29. Liu. Y. Q.; Lowry. G. V. Effect of particle age (Fe-O content) and solution pH on NZVI

366

reactivity: H-2 evolution and TCE dechlorination. Environ. Sci Technol. 2006, 40, 6085–6090.

367

30. Tee, Y. H.; Bachas, L.; Bhattacharyya, D. Degradation of trichloroethylene and

368

dichlorobiphenyls by iron-based bimetallic nanoparticles. J. Phys. Chem. C 2009, 113, 9454–

369

9464.

370

31. Imran, A. New generation adsorbents for water treatment. Chem. Rev. 2012, 112, 5073–5091.

371

32. Leupin, O. X.; Hug, S. J. Oxidation and removal of arsenic (III) from aerated groundwater by

372

filtration through sand and zero-valent iron. Water Res. 2005, 39, 1729−1740.

373

33. Wang, C. M.; Baer, D. R.; Amonette, J. E.; Engelhard, M. H.; Antony, J.; Qiang, Y.

374

Morphology and electronic structure of the oxide shell onthe surface of iron nanoparticles. J. Am.

375

Chem. Soc. 2009, 131, 8824–8832.

376

34. Arslan, I.; Yates, T. J. V.; Browning, N. D.; Midgley, P. A. Embedded nanostructures

377

revealed in three dimensions. Science 2005, 309, 2195–2198.

378

35. Van Aert, S.; Batenburg, K. J.; Rossell, M. D.; Erni, R.; Van Tendeloo, G. Three-

379

dimensional atomic imaging of crystalline nanoparticles. Nature 2011, 470, 374–377.

380

36. Que, E. L.; Bleher, R.; Duncan, F. E.; Kong, B. Y.; Gleber, S. C. et al. Quantitative mapping

381

of zinc fluxes in the mammalian egg reveals the origin of fertilization-induced zinc sparks. Nat.

382

Chem. 2015, 7, 130–139.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 25

383

37. Genc, A.; Kovarik, L.; Gu, M.; Cheng, H.; Plachinda, P.; Pullan, L.; Freitag, B.; Wang, C. M.

384

XEDS STEM tomography for 3D chemical characterization of nanoscale particles.

385

Ultramicroscopy 2013, 131, 24–32.

386

38. Zhang, B.; Su, D. S. Electron Tomography: three-dimensional imaging of real crystal

387

structures at atomic resolution. Angew. Chem. Int. Edit. 2013, 52, 8504–8506.

388

39. Prévot, G.; Nguyen, N. T.; Alloyeau, D.; Ricolleau, C.; Nelayah, J. Ostwald-Driven Phase

389

Separation in Bimetallic Nanoparticle Assemblies. ACS Nano 2016, 10, 4127–4133.

390

40. Wang, Y., Morin, G., Ona-Nguema, G., Juillot, F.,Guyot, F., Calas, G., Brown Jr., G. E.

391

Evidence for different surface speciation of arsenite and arsenate on green rust: an EXAFS and

392

XANES study. Environ. Sci. Technol. 2010, 44, 109−115.

393

41. Shan, Z.W.; Stach, E. A.; Wiezorek, J. M. K.; Knapp, J. A.; Follstaedt, D. M.; Mao, S. X.

394

Grain boundary–mediated plasticity in nanocrystalline nickel. Science 2004, 305, 654-657.

395

42. Kumar, K.S., Van Swygenhoven, H. & Suresh, S. Mechanical behavior of nanocrystalline

396

metals and alloys. Acta Mater. 2003, 51, 5743–5774.

397

43. Parent, L. R.; Robinson, D. B.; Woehl, T. J.; Ristenpart, W. D.; Evans, J. E.; Browning, N.

398

D.; Arslan, I. Direct in Situ Observation of Nanoparticle Synthesis in a Liquid Crystal Surfactant

399

Template. ACS Nano 2012, 6, 3589–3596.

400

44. Krivanek, O. L.; Chisholm, M. F.; Nicolosi, V.; Pennycook, T. J.;

401

N.; Murfitt, M. F.; Own, C. S.; Szilagyi, Z. S.; Oxley, M. P.; Pantelides, S. T.; Pennycook S. J.

402

Atom-by-atom structural and chemical analysis by annular dark-field electron microscopy.

403

Nature 2010, 464, 571−574.

404

45. Stumm, W.; Morgan, J. J. Aquatic Chemistry: Chemical Equilibria and Rates in Natural

405

Waters. (John Wiley & Sons, Inc. Hoboken, 1996).

ACS Paragon Plus Environment

24

Corbin, G. J.; Dellby,

Page 25 of 25

Environmental Science & Technology

406

46. Morel, F. M. M.; Hering, J. G. Principles and Applications of Aquatic Chemistry. (John

407

Wiley & Sons, Inc. Hoboken, 1993).

ACS Paragon Plus Environment

25