Mapping the Reactions in a Single Zero-Valent Iron Nanoparticle

Nov 17, 2017 - (38-40) These surface groups may function as reactive sites to bind with contaminants in the solution, in a mode similar to surface ads...
1 downloads 9 Views 1MB Size
Subscriber access provided by READING UNIV

Article

Mapping the Reactions in a Single Zero-valent Iron Nanoparticle Lan Ling, Xiao-yue Huang, Meirong Li, and Wei-Xian Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b02233 • Publication Date (Web): 17 Nov 2017 Downloaded from http://pubs.acs.org on November 18, 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 28

Environmental Science & Technology

Mapping the Reactions in a Single Zero-valent Iron Nanoparticle

Lan Ling*, Xiaoyue Huang, Meirong Li, 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: +86-21-18501682197, Fax: +86-21-6598 0041.

ACS Paragon Plus Environment

Environmental Science & Technology

1

Abstract

2

Nanoscale zero-valent iron (nZVI) posseses unique functionalities for metal-metalloid

3

removal and sequestration. So far, direct evidence on the heavy metal-iron reactions in

4

the solid phase is still limited due to low concentration of heavy metals and small size

5

of nanoparticles. In this work, angstrom-resolution spectral mappings on the reactions

6

of nZVI with chromate, arsenate, nickel, silver, cesium, and zinc ions are presented.

7

This work was achieved with spherical aberration-corrected scanning transmission

8

electron microscopy integrated with high-sensitivity X-ray energy-dispersive

9

spectroscopy scanning transmission electron microscopy (XEDS-STEM). Results

10

confirm that iron nanoparticles have a core-shell structure. In addition, the removal

11

mechanism significantly depends on the standard potential E0 (E0 is standard potential

12

w.r.t. standard hydrogen electrode at 25 °C when free ion activity is 1.). For strong

13

oxidizing agents, such as Cr(VI), the removal mechanism is diffusion and

14

encapsulation in the core area of the nZVI particle. For moderate oxidizers, such as

15

As(V) with E0 more positive than that of iron, the removal mechanism is adsorption at

16

the surface, followed by diffusion and encapsulation into the particle between the core

17

and the shell. For metal cations with an E0 close to or more negative than that of iron,

18

such as Cs(I) and Zn(II), the removal mechanism is sorption or surface-complex

19

formation. For metal cations with E0 much more positive than that of iron, such as

20

Ag(I), the removal mechanism is rapid reduction on the surface of nZVI. Meanwhile,

21

metals with E0 slightly more positive than that of iron, such as Ni(II), can be

22

immobilized at the nanoparticle surface via sorption and reduction. The synergetic

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology

23

effects of sorption, reduction, and encapsulation mechanisms of nZVI lead to rapid

24

reactions and high efficiency for treatment and immobilization of many toxic heavy

25

metals. Results also demonstrate that the XEDS-STEM technique is a powerful tool

26

for studying reactions in individual nanoparticles and is particularly valuable for

27

mapping trace-level elements in environmental media.

ACS Paragon Plus Environment

Environmental Science & Technology

Graphical Table of Contents

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology

28

Introduction

29

Heavy metal-metalloid ions, such as Cr(VI), As(V), Ni(II), Zn(II) and Ag(I) are

30

nonbiodegradable, bioaccumulative and toxic, which can accumulate in living

31

organisms and most of those are known to be highly toxic or carcinogenic.1,2 A major

32

barrier preventing the fundamental understanding of transport, and reaction

33

mechanisms of trace metal contaminants (80%) arsenic is on the outer shell (~20% volume) of nZVI, and that 60% of

212

the total arsenic deposits between the iron core and the iron-oxide shell (~10%

213

volume).

214

After reacting with Cr(VI), oxidation of the iron nanoparticle is much more extensive

215

than that with As(V), as indicated by the increase in oxygen passing through the iron

216

core of the particle (Figure 2). Meanwhile, Cr and O are entrenched deeper in the iron

217

particle whereas As amasses mostly at the Fe(0)–oxide interface sitting on the surface

218

of the Fe(0) core. In situ time-dependent x-ray absorption spectroscopy has shown

219

that the oxyanions undergo two stages of transformation upon adsorption at the nZVI

220

surface. The first stage corresponds to breaking metal/metalloid–O bonds at the

221

particle surface, and the second stage involves further reduction and diffusion of As

222

and Cr across the thin oxide layer enclosing the nanoparticle.40 The distinctive

39,40

XEDS quantification further demonstrates that

ACS Paragon Plus Environment

Environmental Science & Technology

223

reactive pathways for Cr(VI) and arsenate(V) are expected from their different

224

standard redox potentials [E0 Cr(VI) = +1.36 eV and As(V) = +0.56 V]20,21,40 (see

225

Table S1 in the supporting information), which results in chromium diffusion into the

226

core of nZVI particles whereas the enriched arsenic at the surface of the Fe(0) core

227

region has limited mobility into the interior of the metal core.

228

After reacting with Ni(II), the spherical nZVI particles are devoid of the Fe(0) core,

229

leaving behind doughnut-like or horseshoe-like structures with large cavities in their

230

interior (Figures 3a–3e). The reacted nZVI consists of a bright circlet containing

231

metallic iron Fe(0) in the middle and two outer lower-intensity shells of iron oxides

232

(Figure 3a). The iron mapping (Figure 3b) confirms the disappearance of the Fe signal

233

in the core area, and the Fe footprint takes on a doughnut shape with two dimmer

234

exterior layers (approximately 3–5 nm thick). The two exterior layers consist of

235

oxygen and iron. The Ni mapping and elemental overlay (Figures 3d and 3e)

236

reconfirm that nickel distributes over the surface of the spent particle and accumulates

237

predominantly on the surface of the iron core just inside the outer shell (surface), with

238

some Ni nanoparticle enriched inside the “hollow” area. The Ni nanoparticles thus

239

formed range in size from 2 to 4 nm. In our previous work, electron-energy-loss

240

spectroscopy demonstrated that the Ni nanoparticles contain mainly elemental Ni.42 In

241

the present work, we quantify the relative abundance of Fe, O, and Ni over the spent

242

particle. The Fe:O:Ni atomic ratio over the entire spent particle is 43.4:55.1:1.5,

243

whereas the ratio Fe:O:Ni is 29.8:68.4:1.8 in the outer shell and 50.8:48.3:0.9 in the

244

inner shell. The results of the batch experiments show that 94% of the Ni(II) is

ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Environmental Science & Technology

245

removed within 10 min and 100% in less than 3 h, whereas the “hollowed-out” effect

246

appears after 5 min. The reduced nickel on the nZVI surface can alter the electronic

247

properties of iron and accelerate the oxidization of Fe(0).42 Because the electron

248

transfer from Fe(0) is often suggested as the rate-limiting step for the overall reaction,

249

defects on the particle surface may offer potential breakthrough conduits for the

250

nickel ions to quickly “attack” the core area, which is filled with Fe(0)—the electron

251

donor for Ni(II) reduction. The close standard redox potential with iron and the

252

relatively slow reaction rate may facilitate Ni diffusion into the core area for more

253

Fe(0). Furthermore, the Kirkendall effect may be a mechanism for the formation of

254

hollow nanoparticles.

255

fast-moving Fe ions through the oxide layer and a balancing inward flow of vacancies

256

in the redox reaction.43 The net directional flow of matter is balanced by an opposite

257

flow of vacancies, which condense into pores. The coexistence of multiple-structure

258

spent nZVI suggests that significant variations in the particle lifespan may be the case

259

at the microscopic level. Factors such as physical dimension, grain size, crystallinity,

260

and other structural attributes of the Fe(0) core and the oxide layer may influence the

261

rate of iron oxidation.44

262

After reacting with Ag(I), the characteristics of the nZVI core-shell structure become

263

less noticeable, and the shape of the nZVI nanoparticles is completely distorted with

264

asymmetrical boundaries and dendritic structures similar to a product of

265

electrochemical reduction (Figure 3f). The iron mapping (Figure 3g) shows the

266

footprint of iron, which consists of a small but dense spherical core enclosed by a

43

The voids develop because of the outward transport of

ACS Paragon Plus Environment

Environmental Science & Technology

267

dimmer outer layer. Ag mapping (Figure 3h) shows that the Ag outer layer condenses

268

and completely covers the metallic iron particle, corresponding to the bright dendritic

269

structures in the HAADF image. It may be a mixture of metallic Ag and a small

270

amount of oxygen resulting from the reduction reactions. The oxygen is widely

271

distributed over the whole particle although it is less concentrated in the Ag-rich area

272

(Figures 3i and 3j). The relative abundance of Fe, O, and Ag over the spent particle

273

gives an Fe:O:Ag atomic ratio of 89.7:1.9:8.4 in the dense core and of 3.0:5.6:91.4 in

274

the outer layer. Quantification further confirms that the iron nanoparticles are covered

275

by pure metallic Ag nanostructures with trace iron oxide.

276

The distinctive reactive pathways for Cr(VI), As(V), Ni(I) and Ag(I) may result from

277

their different standard redox potentials, reaction speed, different ionic forms and

278

coordination.2,16,18,21,23,27–30 The oxyanions undergo two stages of transformation upon

279

adsorption at the nZVI surface, including Cr-O and As-O bonds breaking, further

280

reduction and diffusion across the oxide layer.40 The standard redox potential of Ni(II)

281

is slightly more positive than that of Fe(0). Our previous work with high-resolution

282

x-ray photoelectron spectroscopy demonstrates that Ni(II) is immobilized at the

283

nanoparticle surface by both sorption and reduction and that the oxidized iron

284

nanoparticles cannot reduce the Ni(II),45,46 thereby facilitating the diffusion of Ni into

285

the Fe(0) core and accelerating the “hollowing out” of the Fe(0) core. In comparison,

286

Ag(I) has a standard redox potential that is significantly more positive than that of Fe,

287

so Ag(I) can be reduced by Fe(II). The surface and morphological characteristics of

288

nZVI change with the progression of Ag+ electrochemical reduction, metallic Ag

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology

289

deposition, and iron ionization, dissolution, and hydroxide precipitation. As a strong

290

oxidizing agent, Ag is quickly reduced to Ag(0) and deposits onto the surface of the

291

nZVI particles. Meanwhile, nZVI is oxidized to Fe(II) and Fe(III), and the spent

292

particle is covered with iron oxide.

293

Sorption and precipitation of nZVI. After reacting with Cs(I), the spent nZVI

294

particles become brighter in HAADF images and still preserve the core-shell structure

295

with a bright core and a dim shell (Figure 4f). Of the three major elements in the

296

system, Cs has the largest atomic number of 55, O = 8, and Fe = 26, so it is expected

297

to be brighter than the surrounding area in HAADF images. Cs, Fe, and O mappings

298

(Figures 4b–4d) show the footprints of Fe, O, and Cs, respectively. The metallic core

299

of the spent nZVI particle is much denser whereas the signal intensity for iron gets

300

dimmer near the exterior surface. From Figure 4d, the ring of oxygen matches the

301

lower-intensity area of iron, reconfirming the remaining core-shell structure. The

302

volume occupied by cesium is slightly greater than that occupied by oxygen, which

303

may suggest the sorption or surface complex of cesium, which sits just above the

304

particle surface. Cs mapping (Figures 4c and 4e) and the color overlay of the three

305

elements illustrate that Cs is uniformly distributed over the whole zero-valent iron

306

core just on the iron oxide shell, which makes the particle appear brighter in HAADF

307

images.

308

After reacting with Zn(II), HAADF images show that the spent nZVI particle retains

309

its core-shell structure and is decorated with tiny agglomerates. In the system, Zn

310

(atomic number 30) is in the same row of the periodic table as Fe, and so has a similar

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 28

311

contrast for the same thickness in HAADF images. Thus, the agglomerates do not

312

appear brighter than the nZVI particle. Fe and O mappings (Figures 4h and 4i) show

313

the footprints of iron and oxygen and illustrate the retained core-shell structure. Zn

314

mapping (Figure 4g) and the color overlay of the three elements (Figure 4j) show

315

clearly that Zn forms clumps and sticks loosely to the zero-valent iron surface. This

316

may explain why the maximum Zn(II) removal capacity is reportedly more than one

317

order of magnitude greater than the theoretical uptake capacity afforded by surface

318

adsorption.47

319

Cesium is considered the most active metal in the alkali family, which consists of the

320

elements in group 1 (IA) of the periodic table. No reduction is involved in the

321

sequestration process, as expected from the significantly more negative standard

322

redox potential of cesium (E0 Cs = −2.923 eV) compared with iron (E0 Fe=−0.41 V).43

323

Cesium exists in solution as hydrated Cs+ cations, and its speciation is independent of

324

pH and redox potential.48 Cesium does not hydrolyze or form precipitates or sparingly

325

soluble

326

hydrosphere-complexing agents nor intrinsic colloids. Sorption of cesium is rather

327

inefficient owing to the large size and low charge of the Cs ion. The maximum Cs(I)

328

removal rate with nZVI is 38.6% at pH 6. Cs(I) uptake is caused by nonspecific

329

sorption or surface complexation with the iron oxide shell.

330

In addition, no reduction is involved in the Zn(II) sequestration process with nZVI

331

because the standard redox potential of zinc (E0 Zn = −0.76 eV) is more negative than

332

that of iron (E0 Fe=−0.41 V).41 Zn(II) adsorption is strongly influenced by the pH of

compounds,

and

neither

does

it

form

complexes

ACS Paragon Plus Environment

with

typical

Page 19 of 28

Environmental Science & Technology

333

the solution. At elevated pH, bulk precipitation of zinc hydroxide may occur. At the

334

early stage of the reaction (