Insights into Antimony Adsorption on {001} TiO2: XAFS and DFT Study

May 17, 2017 - Antimony (Sb) contamination poses an emerging environmental risk, whereas its removal remains a contemporary challenge due to the lack ...
0 downloads 9 Views 2MB Size
Subscriber access provided by Warwick University Library

Article

Insights into antimony adsorption on {001} TiO2: XAFS and DFT study Li Yan, Jiaying Song, Ting-Shan Chan, and Chuanyong Jing Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 17 May 2017 Downloaded from http://pubs.acs.org on May 17, 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 27

Environmental Science & Technology

Insights into antimony adsorption on {001} TiO2: XAFS and DFT study

1 2

Li Yan†,‡, Jiaying Song†, Tingshan Chan§, Chuanyong Jing†,‡,*

3 4 5



6

for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China

State Key Laboratory of Environmental Chemistry and Ecotoxicology, Research Center

7 ‡

8

University of Chinese Academy of Sciences, Beijing 100049, China

9 10 11

§

National Synchrotron Radiation Research Center, 101 Hsin-Ann Road, Hsinchu Science Park, Hsinchu 30076, Taiwan

12 13

Tel: +86 10 6284 9523; Fax: +86 10 6284 9523

14

E-mail: [email protected]

15 16

1 ACS Paragon Plus Environment

Environmental Science & Technology

17

Abstract

18 19

Antimony (Sb) contamination poses an emerging environmental risk, whereas its

20

removal remains a contemporary challenge due to the lack of knowledge in its surface

21

chemistry and efficient adsorbent. In this study, self-assembly {001} TiO2 was examined

22

for its effectiveness in Sb removal, and the molecular level surface chemistry was studied

23

with X-ray absorption spectroscopy and density functional theory calculations. The

24

kinetics results show that Sb adsorption followed the pseudo-second order reaction, and

25

the Langmuir adsorption capacity was 200 mg/g for Sb(III) and 156 mg/g for Sb(V). The

26

PZC of TiO2, which was 6.6 prior to the adsorption experiment, shifted to 4.8 and 5 days) of the Sb(III) oxidation to Sb(V),16, 19-21 since Sb(V) is more

75

mobile and difficult to remove. Therefore, developing novel adsorbents with high

76

adsorption capacity and chemical stability becomes an urgent requirement for Sb

77

remediation. TiO2 is a promising material due to above advantages. Generally, the

78

adsorption capacity of TiO2 is determined by its specific surface area.22 However, recent

79

studies demonstrate that TiO2 adsorption performance largely depend on its exposed

80

crystal facets and corresponding surface energy.23, 24 The surface energy of three low-

81

index anatase TiO2 facets, which was widely studied for water and dye adsorption,

82

follows the order {001} (0.90 J/m2) > {100} (0.53 J/m2) > {101} (0.44 J/m2).25-27 The

83

equilibrium shape of an anatase crystal according to the Wulff construction and surface

84

energy is a slightly truncated bipyramid enclosed by about 94% {101} and 6% {001}

85

facets.25 However, water and dye molecules prefer to bind to high-energy {001} rather

86

than {101} facets.26-28 Therefore, a reasonable surmise is that TiO2 material with large

87

percentage of exposed {001} facet with high surface energy and large surface area should

88

exhibit an excellent performance for Sb removal.

Although the sequestration of Sb(III) by iron (oxyhydr)oxides is

89

The objective of this study was to investigate the molecular level mechanisms of

90

Sb removal using a novel high-energy {001}-faceted TiO2 adsorbent. The surface

91

reactions were explored using multiple complementary techniques including macroscopic

92

wet chemistry, X-ray absorption fine structure (XAFS) spectroscopy, and density

93

function theory (DFT) calculations. The insights gained from this study further our

94

understanding of Sb chemistry at the mineral-water interface.

5 ACS Paragon Plus Environment

Environmental Science & Technology

95

Materials and methods

96

Materials. KSbC4H4O7·1/2H2O, K2H2Sb2O7·4H2O, and titanium (IV) isopropoxide

97

(TTIP) were purchased from Sinopharm Chemical Reagent Co., Ltd. Sb(III) and Sb(V)

98

stock solutions of 500 mg/L were prepared by dissolving KSbC4H4O7·1/2H2O and

99

K2H2Sb2O7·4H2O in deionized (DI) water, respectively. A Sb contaminated river water

100

sample was collected from a Sb mining site at Xikuangshan area (27°47′ N latitude,

101

111°29′E longitude, Hunan, China) and the detail water analysis is shown in the

102

Supporting Information (SI). TiO2 was prepared with hydrothermal method by using

103

TTIP as precursor, and the synthesis and characterization are detailed in the SI.

104

Adsorption Experiments. Dosage experiments were performed to evaluate the

105

effectiveness of TiO2 and widely-used goethite on Sb removal from a contaminated river

106

water. To the river water sample, increasing adsorbent dosage from 0 to 10 g/L was

107

added. After the samples were mixed on a rotator for 24 h, the suspensions were filtrated

108

through a 0.22-µm membrane filter for analysis.

109

Sb(III) and Sb(V) adsorption isotherms were conducted at pH 7.0 ± 0.2 by adding

110

10 mg of TiO2 into 100 mL 0.04 M NaCl solution with initial Sb concentration ranging

111

from 5 to 500 mg/L. The adsorption kinetics on 0.1 g/L TiO2 were studied at pH 7.0 ±

112

0.2 in 150 mL of 0.04 M NaCl solution with 40 mg/L Sb(III/V). The adsorption pH

113

envelope experiments with 5 mg/L Sb(III/V) and 0.1 g/L TiO2 were conducted in

114

triplicate to determine the adsorption edge. To study the competitive adsorption of

115

Sb(III/V) with anions including As(III/V), PO43-, SO42-, NO3-, and F-, similar pH edge

116

experiments were conducted with Sb to anion molar ratio at 1:1 and 1:5. The suspension

6 ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

Environmental Science & Technology

117

was adjusted to desired pH values in the range 2 to 11 with HCl and NaOH. After the

118

samples were reacted for 24 h, the final pH were measured and the concentraitons were

119

detected.

120

Zeta (ζ) Potential Measurements. Zeta (ζ) potential measurements were conducted

121

using a Zetasizer Nano ZS (Malvern Instrument Ltd., UK). All samples were purged with

122

N2 to eliminate CO2 from the system. The pH of the suspension containing 0.1 g/L TiO2

123

and 5 mg/L Sb(III/V) in 0.04 M NaCl was adjusted to desired values using HCl and

124

NaOH. Suspension samples were placed on a rotator for 24 h and the final pH was

125

measured. The reported ζ potential value was the average of three measurements.

126

Analysis. The concentrations of dissolved Sb were measured using an inductively

127

coupled plasma optical emission spectroscopy (ICP-OES, Optima 2000 DV, Perkin Elmer

128

Co., USA) with a detection limit of 0.032 mg/L. The Sb speciation was determined using

129

a high performance liquid chromatograph (HPLC) coupled with a hydride generation-

130

atomic fluorescence spectrometer (HG-AFS, Jitian, P.R. China) with a detection limit of

131

0.6 µg/L for Sb(III) and 1.2 µg/L for Sb(V).

132

XAFS Study and DFT Calculations. EXAFS spectroscopy was employed to

133

characterize the local coordination environment of adsorbed Sb/As on TiO2. The K-edge

134

spectra of Sb (30,491 eV) and As (11,867 eV) were collected on beamline 01C1 at the

135

National Synchrotron Radiation Research Center (NSRRC), Taiwan. The Sb L-edge

136

(LIII=4,132 eV, LII=4,380 eV, and LI=4,698 eV) X-ray absorption near edge structure

137

(XANES) spectra were collected on beamline 16A1 at the NSRRC. Details of sample

138

preparation and data analysis are reported in the SI. The DFT calculation was performed

139

using Castep package in Materials Studio 7.0 (Accelrys, San Diego, CA),29 with model

7 ACS Paragon Plus Environment

Environmental Science & Technology

140

building and parameter set up are detailed in the SI.

141

Surface Complexation Modeling. The charge distribution multi-site complexation (CD-

142

MUSIC) model with 1-pK TPM adsorption option was employed to describe the Sb

143

adsorption behaviors on TiO2 under the constraint of XAFS and DFT results. The

144

calculation was performed using the chemical equilibrium program MINTEQ to simulate

145

the adsorption and the aqueous reactions with a fixed ionic strength at 0.04 M NaCl. The

146

surface parameters and species used in the model are shown in the SI.

147

Results and discussion

148

Characterization of {001} TiO2. The field-emission scanning electron microscopy (FE-

149

SEM) and high resolution transmission electron microscopy (HR-TEM) characterization

150

demonstrate that the TiO2 spheres by self-assembly 5 nm thick nanosheets exposed nearly

151

100% {001} facet with a lattice fringe of 1.9 Å (inset to Figure 1a).30 The average size of

152

TiO2 spheres was 870 ± 25 nm with a porous structure (Figure 1a). Such a porous

153

structure resulted in a high specific surface area of 205 m2/g (Figure S1a). The Barrett–

154

Joyner–Halenda (BJH) pore size distribution was ranged from 35 to 60 nm with the peak

155

centered at 45 nm (Figure S1b). Raman spectrum revealed the characteristic peaks of

156

anatase TiO2 at 144 (Eg), 397 (B1g), 516 (A1g), and 639 cm-1 (Eg) (Figure S1c).31 The

157

XRD pattern indicated that all diffraction peaks of TiO2 can be indexed to anatase TiO2

158

(JCPDS: 21-1272) (Figure S1d).32

159

TiO2 Performance in Sb-laden Water Treatment. The river sample from a Sb mining

160

site with Sb(V) (5.7 mg/L) at pH 7.6 was collected (Table S2). The effectiveness of

161

{001} TiO2 for Sb removal was evaluated and compared with that of goethite, a

162

commonly used adsorbent for metals.11, 18, 33 As shown in Figure 1b, the residue Sb(V)

8 ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

Environmental Science & Technology

163

concentration was 3.9 µg/L with a TiO2 dosage at 2 g/L, which is well below the MCL of

164

Sb (6 µg/L). In contrast, goethite did not reduce Sb to its MCL even up to a 10 g/L

165

dosage. The specific surface area (SBET) normalized Sb(V) adsorption capacity was 3.76

166

molecules/nm2 for {001} TiO2, which was slightly lower than goethite (4.31), suggesting

167

that the large surface area of {001} TiO2 contributed to its excellent adsorption

168

performance in batch experiments. The promising results motivate our study on Sb

169

chemistry on {001} TiO2 surfaces using integrated macroscopic, spectroscopic and DFT

170

study.

171

Macroscopic Characterization of Sb(III/V) Adsorption. The adsorption isotherms of

172

Sb(III/V) on {001} TiO2 conformed to the Langmuir model (R2 > 0.98, Figure 1c). The

173

maximum adsorption capacity was 200 mg/g for Sb(III) and 156 mg/g for Sb(V), which

174

is considerably higher than that of previously used adsorbents (0.6-198 mg/g for Sb(III),

175

0.2-139 mg/g for Sb(V), Table S1). The SBET normalized adsorption capacity was 4.82

176

and 3.76 molecules/nm2 for Sb(III) and Sb(V), respectively, which was in the range of

177

reported values (0.028-235.85 molecules/nm2 for Sb(III), 0.025-44.89 for Sb(V), Table

178

S1). Notably, Sb adsorption on {001} TiO2 is higher than other two {100}- and {101}-

179

faceted TiO2 samples (Figure S2-4, Table S3), indicating that both high surface area and

180

surface energy of {001} TiO2 determine its high adsorption capacity for Sb, the detailed

181

discussion is shown in the SI.

182

The kinetics of Sb(III/V) adsorption on {001} TiO2 followed the pseudo-second

183

order reaction (R2>0.99, Figure 1d), suggesting that chemical adsorption is the rate-

184

controlling step.34 The adsorption was fast in the first 20 min, and then reached

185

equilibrium after 40 min. The rate constant, k, for Sb(III) adsorption was 0.114 g/mgh,

9 ACS Paragon Plus Environment

Environmental Science & Technology

186

which is higher than that of Sb(V) (0.075 g/mgh, inset to Figure 1d). The faster kinetics

187

of Sb(III) may be attributed to its smaller molecular size and stronger adsorption affinity

188

with TiO2 {001} facet compared with Sb(V).

189

pH Dependence of Sb Adsorption with Coexisting Anions. Figure 2 shows the effect of

190

pH on Sb(III/V) adsorption on {001} TiO2. The results show that pH had a negligible

191

effect on Sb(III) adsorption as evidenced by its consistent 87% removal in the pH range 2

192

to 11. On the other hand, Sb(V) adsorption was favorable at acidic pH with 45-64%

193

adsorption when pH PO43- > SO42- > F- > NO3-. Notably, Sb(V) adsorption was inhibited at high pH due

214

to its anionic adsorption characteristics, and this pH effect was pronounced in the

215

presence of competing ions.

216

The surface chemistry of Sb and As was further compared due to their coexistence

217

and carcinogenicity.1, 8 The results in Figure 2c showed that the Sb(III) adsorption was

218

preferential to As(III). For instance, 89% of Sb(III), compared with 45% of As(III), were

219

adsorbed with the same initial concentrations at 0.041 mM at pH 7. This observation was

220

attributed to the stronger Lewis base property of Sb(III) than As(III),16 exhibiting a

221

stronger binding affinity with Lewis acid of Ti5c atoms on TiO2 surface. Nevertheless,

222

Sb(V) adsorption was slightly less than As(V), as evidenced by only 25% of Sb(V),

223

compared with 30% of As(V), was adsorbed (Figure 2d). This is understandable due to a

224

larger octahedral Sb(OH)6- structure than AsO43-, attributing to a greater steric hindrance.

225

This surmise was justified by our DFT calculations as detailed in the SI. Based on the

226

following XAFS and DFT results, the surface species including Ti2O2SbO-5/3 (Sb(III)-

227

TiO2), Ti2O2Sb(OH)4-5/3 (Sb(V)-TiO2), Ti2O2AsO-5/3 (As(III)-TiO2), and Ti2O2AsO2-5/3

228

(As(V)-TiO2) were used in the CD-MUSIC model to calculate the surface complexation

229

reaction (Table S4). With the inclusion of these surface species, the modeling results

230

agreed well with the experimental adsorption edge curves (Figure 2).

231

XAFS Study. Figure 3a-c present the k2-weighted EXAFS spectra and corresponding

11 ACS Paragon Plus Environment

Environmental Science & Technology

232

Fourier transform (FT) for Sb(III) adsorption samples, and the fitting parameters are

233

listed in Table S5. The results showed that the first Sb coordination shell consisted of

234

three oxygen atoms at a distance of 1.98 Å, suggesting a trigonal pyramidal geometry of

235

Sb(OH)3. The second FT peak was ascribed to a Sb-Ti single scattering (SS) path at the

236

distance of 3.50 Å with a coordination number (CN) of 0.5-0.8. XAFS-derived bond

237

lengths have a smaller relative error (~ 0.01-0.02 Å), compared to the coordination

238

number (10-20% error).38 Therefore, the fact that the measured bond length of 3.50 Å, is

239

consistent with a bidentate binuclear (BB) adsorption complex, which is also observed on

240

iron and aluminum oxides.10, 12, 13 In fact, it is not the first case where the structural

241

configuration derived from atomic distance does not match well with that from

242

coordination number,39-41 considering the monodentate mononuclear (MM) complex with

243

hydrogen bonding to an adjacent singly-coordinated hydroxyl group.40 In this study, the

244

thermodynamically stable BB structure of Sb(III) on TiO2 {001} facet was confirmed by

245

DFT calculations (Figure 4a), with a comparable Sb-O (2.01 Å) and Sb-Ti (3.47 Å)

246

distances with EXAFS results.

247

For Sb(V) adsorption samples, the k2-weighted EXAFS spectra are shown in

248

Figure 3d-f. The first Sb coordination shell was resolved by six O atoms at 1.98 Å,10, 11, 13

249

and the second shell resulted in a Sb-Ti SS path at 3.59-3.72 Å (Table S6). These

250

distances were within the range of BB configuration with an average Sb-O distance at

251

2.05 Å and Sb-Ti distance at 3.70 Å as elucidated by our DFT results (Figure 4b).

252

Notably, a Sb-Sb shell at 4.00 Å with CN of 1.0-2.2 was also observed, indicative

253

of surface precipitation.10, 13 The formation of surface precipitation was attributed to the

254

high surface density over 60 mg/g with initial Sb concentration of 40 mg/L. However,

12 ACS Paragon Plus Environment

Page 12 of 27

Page 13 of 27

Environmental Science & Technology

255

precipitation was a slow dynamic process. For instance, only 2.7% and 2.2% loss in

256

dissolved Sb(III) and Sb(V) concentrations were observed within 24 h, respectively, and

257

less than 8.2% in four days with an initial concentration of 213 mg/L for Sb(III) and 182

258

mg/L for Sb(V) at pH 7 (Figure S7). The formation of precipitates should be assisted by

259

TiO2 surfaces, which provide “a seed” to promote the nucleation process of

260

precipitation.39, 42

261

The normalized Sb K-edge XANES spectra demonstrated that no oxidation

262

occurred during adsorption and XAFS data collection (Figure S8). Meanwhile, Sb K-

263

edge EXAFS spectra exhibited no significant surface structural difference between single

264

Sb and Sb-As coexisting samples (Figure 3 and Table S5-6), suggesting that the existence

265

of As does not affect the adsorption configuration of Sb on TiO2. In turn, As K-edge

266

EXAFS spectra indicated that the presence of Sb have negligible influence on the As BB

267

adsorption structure (Figure S9 and Table S7).39 In addition, the change of pH condition

268

exhibited no effect on adsorption configurations as evidenced by EXAFS analysis (Figure

269

3 and S9, Table S5-7).

270

Sb L-edge spectra were examined to study the distribution of density of states

271

(DOS) upon surface complexation formation, and the spectra and their corresponding

272

first derivatives are shown in Figure S10. The LI edge spectrum exhibited no change

273

upon adsorption due to the negligible DOS change of p states in the vicinity of Fermi

274

level (Figure 5).43 In contrast, the LII/III edges resolved discrepancies upon adsorption

275

(Figure S10a-b, bottom). Notably, the transition of 2p to pd hybridization states (peak 7

276

in Figure S10b) move towards higher energy, suggesting charge redistribution upon

277

adsorption as detailed in the SI. Our DFT calculations demonstrated that the adsorption

13 ACS Paragon Plus Environment

Environmental Science & Technology

278

imparts 0.17 and 0.12 electrons to Sb(OH)3 and Sb(OH)6- molecules, respectively (Figure

279

4). The acquisition of electrons by Sb was coupled with the formation of new bonding

280

and DOS redistribution. Some p character states of Sb were elevated to above Fermi level

281

of 2.5 eV (Figure 5), which hybrid with d character states of Ti to form pd hybridization

282

states. Therefore, the peak 7 in Figure S10 was shifted to high energy upon adsorption. In

283

sum, the change in orbital energy derived from orbital hybridizing of adsorbed molecules

284

on surface should be expected, and this is the driving force that underlines the surface

285

chemistry for Sb adsorption.

286

Surface Chemistry for Sb Adsorption. When molecules react with surfaces, the

287

interactions can be generally classified into the following three types:44 (i) the highest

288

occupied molecular orbital (HOMO) of molecules bond with the conduction band (CB)

289

of surface; (ii) the lowest unoccupied molecular orbital (LUMO) of molecules bond with

290

the valence band (VB) of surface; and (iii) the HOMO of molecules bond with the VB of

291

surface (Figure 6). These interactions bind the molecule onto the surface, and electrons

292

are generally transferred from bonding orbitals in one component to antibonding orbitals

293

in the other. Thus, a new bond is formed between the adsorbed molecule and the surface,

294

inducing a stabilization effect. Then, the fundamental question is where the electrons go

295

and at what energy the bonding occurs during the Sb adsorption process? Based on the

296

molecular orbital (MO) theory and projected density of states (PDOS) analysis, the MO

297

energy level diagram for Sb(III) and Sb(V) adsorption on TiO2 was constructed as shown

298

in Figure 6.

299

Our DOS results shown in Figure 5 demonstrated that for both TiO2 surface and

300

Sb(III) molecule, the antibonding exists above the Fermi level with an energy range 2.5

14 ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27

Environmental Science & Technology

301

to 4.8 eV because of the interaction (iii) that electrons occupied the antibonding orbitals

302

between Sb and surface (Figure 6, red line). The formation of chemical bond during

303

adsorption was further elucidated by the overlap of PDOS as shown in Figures S11-12.

304

The 2p orbitals of O1 and O2 from Sb(OH)3 overlap, respectively, with 3d orbitals of Ti34

305

and Ti14 (Figure S11), forming a Sb-O-Ti surface complex (Figure 4a). By comparing the

306

PDOS of uncoordinated O3 with the other two coordinated O1 and O2 atoms for Sb(III)

307

adsorption, the energy of coordinated O1 and O2 atoms is lowered, and their p orbital

308

energy is located between -5.2 to -0.2 eV, in the same energy range as the Ti-d orbitals

309

(Figure S11). Thus, the major contribution to the newly formed Ti-O bonds is due to the

310

electron sharing between O-2p and Ti-3d orbitals,45 which corresponds to the interaction

311

(i) (Figure 6, purple line). When Sb(III) forms a BB complex on TiO2 surface,

312

deprotonation occurs with two H atoms fall off from Sb(OH)3 and bond with the O2c

313

atoms on TiO2 surface. The bonding of surface O2c and H atoms can be justified by their

314

overlap in PDOS. The results indicated that compared with the undissipated H3 atom, the

315

dissociated H1 and H2 atoms reduce their s orbital energy to the range of -6.0 to -4.8 eV

316

to match the lowered 2p orbital energy of surface O29 and O69 (Figure S11). The

317

dissociative adsorption and surface reconstruction upon adsorption significantly decrease

318

the surface energy due to the interaction (ii) (Figure 6, green line), leading to a stable

319

adsorption configuration.

320

Sb(V) adsorption also exhibited the DOS redistribution (Figure 5), and PDOS

321

overlap for newly formed Ti14-O2 and Ti34-O5 bonds (Figure S12). Similarly, dissociative

322

adsorption and formation of surface hydroxyl O29-H2 and O69-H5 were observed for

323

Sb(V) molecule (Figure 4b). However, the hybrid orbital energy for Ti-O bonds located at

15 ACS Paragon Plus Environment

Environmental Science & Technology

324

high energy range of -5.3 to 0.1 eV, indicating that these bonds are less stable than those

325

at low energy (-5.2 to -0.2 eV) for Sb(III) adsorption. The bonding resulted in a lower

326

adsorption energy of Sb(III) (-4.99 eV), compared with that of Sb(V) (-4.71 eV, Figure

327

4), indicating a favorable adsorption of Sb(III), which is in line with our macroscopic

328

adsorption results and other independent studies.34, 46

329

The understanding of Sb surface chemistry enables us to explain why TiO2

330

exposed with high-energy {001} facet exhibited favorable Sb adsorption. Our DFT

331

calculations indicated that the surface reconstruction exists on {001} facets in contact

332

with Sb molecules.23 With respect to the pristine {001} facet with 100% Ti5c and 100%

333

O2c atoms, surface reconstruction reduces the amount of O2c atoms by the formation of

334

O2c-H bond, where H is contributed by the dissociation of Sb(OH)3 or Sb(OH)6-. Both

335

surface reconstruction and dissociation adsorption of Sb molecules contribute to a

336

favorable adsorption configuration with a low adsorption energy.28 Therefore, the {001}

337

TiO2 is capable of immobilizing Sb, and this molecular mechanism may be generalizable

338

and applicable to other metal oxide surfaces.

339 340

Acknowledgements

341

We acknowledge the financial support of the National Basic Research Program of

342

China (2015CB932003, 2016YFA0203102), the Strategic Priority Research Program of

343

the Chinese Academy of Sciences (XDB14020201), and the National Natural Science

344

Foundation of China (41373123, 41425016, and 21321004). The XAFS spectra were

345

acquired at the National Synchrotron Radiation Research Center (NSRRC) BL01C1,

346

Taiwan.

16 ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

Environmental Science & Technology

347 348

Supporting Information

349

Details of TiO2 synthesis, BET, XRD, Raman characterization, CD-MUSIC modeling, Sb

350

adsorption comparison, Competitive adsorption, XAFS results, DFT calculations, PDOS

351

analysis, and additional figures and tables. This material is available free of charge via

352

the Internet at http:// pubs.acs.org.

353 354

Notes

355

The authors declare no competing financial interest.

17 ACS Paragon Plus Environment

Environmental Science & Technology

356

References

357

(1) Okkenhaug, G.; Zhu, Y.-G.; He, J.; Li, X.; Luo, L.; Mulder, J. Antimony (Sb) and arsenic (As) in Sb mining impacted paddy soil from Xikuangshan, China: Differences in mechanisms controlling soil sequestration and uptake in rice. Environ. Sci. Technol. 2012, 46, 3155-3162. (2) Multani, R. S.; Feldmann, T.; Demopoulos, G. P. Antimony in the metallurgical industry: A review of its chemistry and environmental stabilization options. Hydrometallurgy 2016, 164, 141-153. (3) Filella, M.; Belzile, N.; Chen, Y. W. Antimony in the environment: a review focused on natural waters I. Occurrence. Earth-Sci. Rev. 2002, 57, 125-176. (4) Adra, A.; Morin, G.; Ona-Nguema, G.; Menguy, N.; Maillot, F.; Casiot, C.; Bruneel, O.; Lebrun, S.; Juillot, F.; Brest, J. Arsenic scavenging by aluminum-substituted ferrihydrites in a circumneutral pH river impacted by acid mine drainage. Environ. Sci. Technol. 2013, 47, 12784-12792. (5) Bolanz, R. M.; Wierzbicka-Wieczorek, M.; Čaplovičová, M.; Uhlík, P.; Göttlicher, J.; Steininger, R.; Majzlan, J. Structural incorporation of As5+ into hematite. Environ. Sci. Technol. 2013, 47, 9140-9147. (6) Yan, L.; Hu, S.; Jing, C. Recent progress of arsenic adsorption on TiO2 in the presence of coexisting ions: A review. J. Environ. Sci. 2016, 49, 74-85. (7) Dong, G.; Huang, Y.; Yu, Q.; Wang, Y.; Wang, H.; He, N.; Li, Q. Role of nanoparticles in controlling arsenic mobilization from sediments near a realgar tailing. Environ. Sci. Technol. 2014, 48, 7469-7476. (8) Wilson, S. C.; Lockwood, P. V.; Ashley, P. M.; Tighe, M. The chemistry and behaviour of antimony in the soil environment with comparisons to arsenic: A critical review. Environ. Pollut. 2010, 158, 1169-1181. (9) Filella, M.; Williams, P. A.; Belzile, N. Antimony in the environment: knowns and unknowns. Environ. Chem. 2009, 6, 95-105. (10) Scheinost, A. C.; Rossberg, A.; Vantelon, D.; Xifra, I.; Kretzschmar, R.; Leuz, A. K.; Funke, H.; Johnson, C. A. Quantitative antimony speciation in shooting-range soils by EXAFS spectroscopy. Geochim. Cosmochim. Acta 2006, 70, 3299-3312. (11) Guo, X. J.; Wu, Z. J.; He, M. C.; Meng, X. G.; Jin, X.; Qiu, N.; Zhang, J. Adsorption of antimony onto iron oxyhydroxides: Adsorption behavior and surface structure. J. Hazard. Mater. 2014, 276, 339-345. (12) Mitsunobu, S.; Takahashi, Y.; Terada, Y.; Sakata, M. Antimony(V) incorporation into synthetic ferrihydrite, goethite, and natural iron oxyhydroxides. Environ. Sci. Technol. 2010, 44, 3712-3718. (13) Ilgen, A. G.; Trainor, T. P. Sb(III) and Sb(V) sorption onto Al-rich phases: Hydrous Al oxide and the clay minerals kaolinite KGa-1b and oxidized and reduced nontronite NAu-1. Environ. Sci. Technol. 2012, 46, 843-851. (14) Ramadugu, S. K.; Mason, S. E. DFT study of antimony(V) oxyanion adsorption

358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395

18 ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435

Environmental Science & Technology

on alpha-Al2O3(1-102). J. Phys. Chem. C 2015, 119, 18149-18159. (15) Mason, S. E.; Trainor, T. P.; Goffinet, C. J. DFT study of Sb(III) and Sb(V) adsorption and heterogeneous oxidation on hydrated oxide surfaces. Comput. Theor. Chem. 2012, 987, 103-114. (16) Leuz, A. K.; Monch, H.; Johnson, C. A. Sorption of Sb(III) and Sb(V) to goethite: Influence on Sb(III) oxidation and mobilization. Environ. Sci. Technol. 2006, 40, 72777282. (17) Vithanage, M.; Rajapaksha, A. U.; Dou, X. M.; Bolan, N. S.; Yang, J. E.; Ok, Y. S. Surface complexation modeling and spectroscopic evidence of antimony adsorption on iron-oxide-rich red earth soils. J. Colloid Interface Sci. 2013, 406, 217-224. (18) Fan, J. X.; Wang, Y. J.; Fan, T. T.; Dang, F.; Zhou, D. M. Effect of aqueous Fe(II) on Sb(V) sorption on soil and goethite. Chemosphere 2016, 147, 44-51. (19) Xu, C.; Zhang, B.; Zhu, L.; Lin, S.; Sun, X.; Jiang, Z.; Tratnyek, P. G. Sequestration of antimonite by zerovalent iron: Using weak magnetic field effects to enhance performance and characterize reaction mechanisms. Environ. Sci. Technol. 2016, 50, 1483-1491. (20) Fan, J. X.; Wang, Y. J.; Fan, T. T.; Cui, X. D.; Zhou, D. M. Photo-induced oxidation of Sb(III) on goethite. Chemosphere 2014, 95, 295-300. (21) Kong, L. H.; Hu, X. Y.; He, M. C. Mechanisms of Sb(III) oxidation by pyriteinduced hydroxyl radicals and hydrogen peroxide. Environ. Sci. Technol. 2015, 49, 34993505. (22) Xu, Z.; Meng, X. Size effects of nanocrystalline TiO2 on As(V) and As(III) adsorption and As(III) photooxidation. J. Hazard. Mater. 2009, 168, 747-752. (23) Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53229. (24) Liu, G.; Yang, H. G.; Pan, J.; Yang, Y. Q.; Lu, G. Q.; Cheng, H. M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014, 114, 9559-9612. (25) Lazzeri, M.; Vittadini, A.; Selloni, A. Structure and energetics of stoichiometric TiO2 anatase surfaces. Phys. Rev. B 2001, 63, 155409. (26) Zhao, Z.; Li, Z.; Zou, Z. A theoretical study of water adsorption and decomposition on the low-index stoichiometric anatase TiO2 surfaces. J. Phys. Chem. C 2012, 116, 7430-7441. (27) De Angelis, F.; Vitillaro, G.; Kavan, L.; Nazeeruddin, M. K.; Graetzel, M. Modeling ruthenium-dye-sensitized TiO2 surfaces exposing the (001) or (101) faces: A first-principles investigation. J. Phys. Chem. C 2012, 116, 18124-18131. (28) Vittadini, A.; Selloni, A.; Rotzinger, F. P.; Gratzel, M. Structure and energetics of water adsorbed at TiO2 anatase (101) and (001) surfaces. Phys. Rev. Lett. 1998, 81, 29542957. (29) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M. I. J.; Refson, K.; Payne, M. C. First principles methods using CASTEP. Z. Kristallogr. 2005, 220, 56719 ACS Paragon Plus Environment

Environmental Science & Technology

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

570. (30) Chen, J. S.; Tan, Y. L.; Li, C. M.; Cheah, Y. L.; Luan, D. Y.; Madhavi, S.; Boey, F. Y. C.; Archer, L. A.; Lou, X. W. Constructing hierarchical spheres from large ultrathin anatase TiO2 nanosheets with nearly 100% exposed (001) facets for fast reversible lithium storage. J. Am. Chem. Soc. 2010, 132, 6124-6130. (31) Ohsaka, T.; Izumi, F.; Fujiki, Y. Raman spectrum of anatase, TiO2. J. Raman Spectrosc. 1978, 7, 321-324. (32) Yan, L.; Du, J.; Jing, C. How TiO2 facets determine arsenic adsorption and photooxidation: Spectroscopic and DFT study. Catal. Sci. Technol. 2016, 6, 2419-2426. (33) Yang, Y.; Yan, W.; Jing, C. Dynamic adsorption of catechol at the goethite/aqueous solution interface: A molecular-scale study. Langmuir 2012, 28, 1458814597. (34) Luo, J. M.; Luo, X. B.; Crittenden, J.; Qu, J. H.; Bai, Y. H.; Peng, Y.; Li, J. H. Removal of antimonite (Sb(III)) and antimonate (Sb(V)) from aqueous solution using carbon nanofibers that are decorated with zirconium oxide (ZrO2). Environ. Sci. Technol. 2015, 49, 11115-11124. (35) Filella, M.; Belzile, N.; Chen, Y. W. Antimony in the environment: a review focused on natural waters II. Relevant solution chemistry. Earth-Sci. Rev. 2002, 59, 265285. (36) Wang, X.; Li, X.; Zhang, X.; Qian, S. Speciation analysis of antimony in water samples via combined nano-sized TiO2 colloid preconcentration and AFS analysis. J. Anal. At. Spectrom. 2014, 29, 1944-1948. (37) Xi, J.; He, M.; Lin, C. Adsorption of antimony(III) and antimony(V) on bentonite: Kinetics, thermodynamics and anion competition. Microchem. J. 2011, 97, 85-91. (38) Kelly, S.; Hesterberg, D.; Ravel, B. Analysis of soils and minerals using X-ray absorption spectroscopy. Methods of soil analysis. Part 5 2008, 387-463. (39) Hu, S.; Yan, L.; Chan, T.; Jing, C. Molecular insights into ternary surface complexation of arsenite and cadmium on TiO2. Environ. Sci. Technol. 2015, 49, 59735979. (40) Mikutta, C.; Kretzschmar, R. Spectroscopic evidence for ternary complex formation between arsenate and ferric iron complexes of humic substances. Environ. Sci. Technol. 2011, 45, 9550-9557. (41) Ona-Nguema, G.; Morin, G.; Juillot, F.; Calas, G.; Brown, G. E. EXAFS analysis of arsenite adsorption onto two-line ferrihydrite, hematite, goethite, and lepidocrocite. Environ. Sci. Technol. 2005, 39, 9147-9155. (42) Stumm, W., Chemistry of the solid-water interface. Wiley-Interscience: New York, 1999. (43) Sun, F. H.; Wu, F. C.; Liao, H. Q.; Xing, B. S. Biosorption of antimony(V) by freshwater cyanobacteria Microcystis biomass: Chemical modification and biosorption mechanisms. Chem. Eng. J. 2011, 171, 1082-1090. 20 ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

476 477 478 479 480 481 482 483 484 485 486

Environmental Science & Technology

(44) Hoffmann, R., Solids and surfaces: A chemist's view of bonding in extended structures. Wiley-VCH: New York, 1988. (45) Yan, L.; Hu, S.; Duan, J.; Jing, C. Insights from arsenate adsorption on Rutile (110): Grazing-incidence X-ray absorption fine structure spectroscopy and DFT+U study. J. Phys. Chem. A 2014, 118, 4759-4765. (46) Zou, J. P.; Liu, H. L.; Luo, J.; Xing, Q. J.; Du, H. M.; Jiang, X. H.; Luo, X. B.; Luo, S. L.; Suib, S. L. Three-dimensional reduced graphene oxide coupled with Mn3O4 for highly efficient removal of Sb(III) and Sb(V) from water. ACS Appl. Mater. Interfaces 2016, 8, 18140-18149.

21 ACS Paragon Plus Environment

Environmental Science & Technology

487

488 489 490 491 492 493 494 495

Figure 1. (a) FE-SEM image of {001} TiO2 with inset HR-TEM characterization. (b) Residue Sb(V) concentrations in river water as a function of dosage of TiO2 and goethite, initial Sb(V) concentration was 5.7 mg/L, pH = 7.6. (c) Isotherm and (d) kinetics for Sb(III) and Sb(V) adsorption on 0.1 g/L TiO2 in 0.04 M NaCl solution at pH 7. Symbols are experimental data, and solid lines represent the Langmuir model (c) and pseudo-2nd kinetics model simulation (d). Inset tables in c and d shows model parameters.

22 ACS Paragon Plus Environment

Page 22 of 27

Page 23 of 27

Environmental Science & Technology

496

497 498 499 500 501 502 503 504 505

Figure 2. (a) Experimental pH adsorption edges (symbols) and charge distribution multisite complexation (CD-MUSIC) models (lines) of Sb(III) and Sb(V) adsorption on TiO2. (b) Zeta potential of 0.1 g/L TiO2 as blank (green), 0.041 mM Sb(III) (red), and 0.041 mM Sb(V) (blue) adsorption samples as a function of pH in 0.04 M NaCl solution. Error bars represent the standard deviation (n=3). Adsorption pH edges of coexisting (c) Sb(III) and As(III), (d) Sb(V) and As(V) on TiO2. Both the initial Sb and As concentrations were 0.041 mM; adsorbent dosage was 0.1 g/L.

23 ACS Paragon Plus Environment

Environmental Science & Technology

506

507 508 509 510 511 512 513

Figure 3. Observed (red circles) and fitted (black lines) (a, d) k2-weighted Sb K-edge EXAFS spectra and their corresponding (b, e) FT magnitude and (c, f) real parts. The samples were prepared by reacting 0.33 mM Sb, with 0.33 mM As for co-adsorption, on 0.1 g/L TiO2 in 0.04 M NaCl at pH 4, 7, 10. The fitting parameters are shown in Table S5 and S6.

514 515 516 517

24 ACS Paragon Plus Environment

Page 24 of 27

Page 25 of 27

Environmental Science & Technology

518

519 520 521 522

Figure 4. Density functional theory (DFT) optimized bidentate surface configuration for (a) Sb(III) and (b) Sb(V) adsorption on TiO2 {001} facet.

25 ACS Paragon Plus Environment

Environmental Science & Technology

523

524 525 526 527 528

Figure 5. Density of states (DOS) analysis for Sb(III) and Sb(V) adsorption on TiO2. The left and right ones indicated the DOS of TiO2 and Sb molecule before adsorption, respectively. The middle columns suggested the DOS of TiO2 and Sb molecule after adsorption.

529 530 531 532

26 ACS Paragon Plus Environment

Page 26 of 27

Page 27 of 27

Environmental Science & Technology

533 534

535 536 537 538 539

Figure 6. Molecular orbital energy level diagram for Sb(III) and Sb(V) adsorption on TiO2. Three types of interactions between Sb and TiO2 surface are represented with dash lines in (i) purple, (ii) green, and (iii) red.

540 541 542

27 ACS Paragon Plus Environment