Low Molecular Weight Organic Acid Complexation Affects Antimony(III

Apr 10, 2019 - The CD-MUSIC adsorption model indicates that this was solely caused by strong competition from tartrate complexation in solution, which...
0 downloads 0 Views 947KB Size
Subscriber access provided by UNIV OF LOUISIANA

Remediation and Control Technologies

Low Molecular Weight Organic Acid Complexation Affects Antimony(III) Adsorption by Granular Ferric Hydroxide Xiaochen Li, Tatiana Reich, Michael Kersten, and Chuanyong Jing Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06297 • Publication Date (Web): 10 Apr 2019 Downloaded from http://pubs.acs.org on April 16, 2019

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 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 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.

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 30

Environmental Science & Technology

119x137mm (300 x 300 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

1

Low Molecular Weight Organic Acid Complexation Affects

2

Antimony(III) Adsorption by Granular Ferric Hydroxide

3

Xiaochen Li1,2, Tatiana Reich2, Michael Kersten2,*, Chuanyong Jing1,3**

4

1State

5

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

6

2Geosciences

7

3University

Key Laboratory of Environmental Chemistry and Ecotoxicology, Research Center for

Institute, Johannes Gutenberg University, Mainz 55099, Germany

of Chinese Academy of Sciences, Beijing 100049, China

8

9

ABSTRACT: Antimony(III) mobility in natural aquatic environments is generally enhanced

10

by dissolved organic matter. Tartaric acid is often used to form complexes with and stabilize

11

dissolved Sb(III) in adsorption studies. However, competition between such low molecular

12

organic acid complexation and adsorption of Sb(III) has received little attention, which

13

prompted us to measure Sb(III) adsorption by iron oxyhydroxide adsorbents commonly used in

14

water treatment plants. Sb K-edge X-ray absorption fine structure (EXAFS) spectra gave Sb–

15

O and Sb–Fe distances and coordinations compatible with a bidentate binuclear inner-sphere

16

complex with trigonal Sb(O,OH)3 polyhedra sharing corners with Fe(O,OH)6 octahedra, and a

17

bidentate mononuclear inner-sphere complex but with Sb(O,OH)4 tetrahedra at alkaline pH.

18

Experimental batch titration data were fitted using the charge-distribution multi-site surface

19

complexation (CD-MUSIC) model, constrained by the EXAFS molecular level information and

20

taking competitive effects by the organic ligand into consideration. The proportion adsorbed at

21

acid–neutral pH decreased as the Sb(III) concentration increased. The CD-MUSIC adsorption

22

model indicates that this was solely caused by strong competition from tartrate complexation ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30

Environmental Science & Technology

2 23

in solution, which leads to adsorption constants higher than those derived without taking this

24

competition into account. The adsorption model results allow for calculating a pe-pH

25

predominance diagram using the USGS PhreePlot code. The study provides consistent surface

26

complexation stability constants, allowing the new database to be used also to model reliably

27

adsorption of toxic oxyanions in anoxic aqueous environments, for example to accurately

28

simulate competition between Sb(III) and As(III).

29 30

INTRODUCTION

31

Antimony (Sb) and its compounds are important environmental contaminants of emerging

32

importance.1 Like As, Sb is primarily found in the +3 valence state (i.e., Sb(III)) and the +5

33

valence state (i.e., Sb(V)), under natural reducing and oxidizing conditions, respectively. Sb(V)

34

is the dominant species in oxygenated aqueous systems, but multiple studies have found Sb(III)

35

in marine and fresh water, groundwater, and soil porewater.1 In sulfide-free solutions, Sb(III)

36

undergoes hydrolysis to give neutral antimonyl (Sb(OH)30) and antimonite oxyanion

37

(Sb(OH)4−). There is evidence that Sb(III) may be a human skin carcinogen, as is the case for

38

As(III).2 The recommended maximum total Sb concentrations in drinking water (US

39

Environmental Protection Agency 6 g L−1, EU 5 g L−1, Japan 2 g L−1) are even lower than

40

the recommended maximum total As concentrations. Total dissolved Sb is 10 mol L−1 (the solubility of Sb2O3 in pure water at ~pH 7) to be used. Laboratory adsorption

106

studies often use adsorbate concentrations much higher than expected in natural aqueous media.

107

This is because surface complexation models require adsorption–pH curves to be fitted, and

108

there are no such curves to fit if ~100% of adsorbate is adsorbed over a broad pH range at low

109

concentrations. The experiments were therefore performed with higher concentrations of 0.5,

110

1.0 and 1.5 mmol L−1, where a pH dependence of the curves becomes obvious.

111

A 30 mL aliquot of a diluted Na tartrate or Sb(III) tartrate stock solution acidified with

112

0.1 M HNO3 to pH 3 was added to a low-density polyethylene centrifuge tube, then the sorbent

113

was added (1.7 g L−1 fresh GFH based on a pore water content of 40% by weight). Between 10

114

and 20 experiments at different pH values were performed for each background electrolyte

115

concentration (10, 50 and 100 mM NaNO3) by adding increasing amount of 1 M NaOH to a

116

series of tubes. The tubes were shaken for 24 h on a horizontal shaker in a dark room. In

117

preliminary adsorption experiments, an equilibration time of 24 h was found to be sufficient

118

even for the very porous GFH (particle size 4. The precision and accuracy of the

128

analytical procedure were assessed by analysing the NIST 1643e certified reference material

129

(US NIST, Gaithersburg, MD, USA) and the analytical uncertainty was less than 5%. Sb

130

speciation in selected solutions was determined by anodic stripping voltammetry and the results

131

indicated no oxidation to Sb(V) as shown in Figure S3 (Supporting Information). The dissolved

132

organic carbon (tartrate) concentrations in the supernatants were determined using a High TOC

133

II carbon analyser with a combustion tube and CO2 detector (Elementar Analysensysteme,

134

Hanau, Germany) using the German DIN EN 1484 (DEV H3) method.

135

X-ray Absorption Analysis. Incident X-ray energy was scanned across the Sb K-edge (30491

136

eV) using the Swiss Light Source SuperXAS beamline (PSI-Villigen, Switzerland). When the

137

measurements were made, the beamline optics consisted of a water-cooled Si(311) double

138

crystal monochromator between two Rh-coated harmonic rejection mirrors, one for beam

139

collimation and the other for focusing. The monochromator was calibrated using a Sb(0) foil.

140

A sample was mounted on the stage and then cooled to 100 K using a nitrogen cryostream

141

manipulator (Cryojet, Oxford Instruments, High Wycombe, UK) to improve data quality and

142

avoid Sb(III) photo-oxidation. Take care of the air humidity when using a cryojet device.

143

Spectra for the akaganéite samples were collected in fluorescence mode. Spectra of Sb(III) and

144

Sb(V) reference minerals were also acquired but in transmission mode at room temperature.

145

The natural reference minerals schafarzikite (FeSbIII2O4) and tripuhyite (FeSbVO4) were ground

146

to fine powders, diluted with polyethylene powder (Merck) and pressed into pellets using a 13

ACS Paragon Plus Environment

Environmental Science & Technology

7 147

mm die. A number of scans for each sample (three scans for a reference mineral and six scans

148

for an akaganéite sample) were averaged to improve the signal-to-noise ratio. The XAS data

149

were automatically deadtime-corrected. The X-ray absorption near-edge structure (XANES)

150

parts of the first and last spectra were compared to ensure that no photon-induced redox

151

reactions had occurred while a sample was being analysed.

152

The energy scales of the experimental data were recalibrated, then spline fitting was

153

performed using the AUTOBK program.18 The background correction, extraction of the

154

EXAFS parts of the spectra and harmonic analysis were performed using the EXAFSPAK

155

software package.19 The Sb K-edge k3-weighted (k) spectra for the reference minerals had

156

very good signal-to-noise ratios out to k values ~15 Å−1 (Figure 1). The resulting EXAFS (k)

157

spectra were then converted to R-space by taking the Fourier transform of (k). Least-squares

158

refinement to the schafarzikite structure was achieved using the OPT mode in the EXAFSPAK

159

program for (k) spectra of the pH 4 sample in the range 2.4–11.4 Å−1, and of the pH 7 and pH

160

10 samples in the range 2.4–12.4 Å−1. Fitting of the data to the EXAFS equation was performed

161

to determine the degeneracy N, half-path length R, and mean-square displacement 2 of the

162

backscatterers. Theoretical scattering phases and amplitudes were obtained using FEFF8.2.20 A

163

124-atom cluster based on the schafarzikite crystal structure was used to define the Hedin–

164

Lundqvist self-energy potential for FEFF calculations of the theoretical effective pathways. All

165

the interactions were modelled using single scattering (SS) interactions derived from previously

166

published structural refinements.21,22 There is no evidence of MS paths in our data. Sb K-edge

167

data processing was achieved using three paths (a SS Sb–O shell, a SS Sb–Fe shell, and a SS

168

Sb–Sb shell), which were found to be significant contributors to the EXAFS signal. Several

169

measures were taken to prevent the number of degrees of freedom exceeding the number of

170

parameters allowed to vary when fitting the data for the reference minerals. E0 was allowed

171

to vary only as a global parameter for each fitting procedure. The amplitude reduction factor ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30

Environmental Science & Technology

8 172

S02 was set at 1.0 to assure agreement between the resulting structural parameters of reference

173

samples with their XRD data. The structural results (coordination numbers N, real backscatterer

174

distances r, and Debye-Waller factors 2) of the least-squares-fitting process using the SS shell

175

contributions to the reference mineral spectra are summarized in Tables S2 and S3 (Supporting

176

Information), and for the adsorption samples in Table 1. The fit quality was defined as χ2res =

177

∑[χidata − χimodel(x)] / ∑[χidata]2, where χ is the magnitude of EXAFS oscillation, and x is the set

178

of variables to be refined. Our EXAFS agreed well with previously published XRD data for the

179

reference minerals, as shown in Tables S2 and S3 (Supporting Information). Notably, in

180

addition to the schafarzikite coordination shells from previously published XRD refinements,

181

a weak shell of 0.5 S atoms at a distance of ~2.5 Å was detected, indicative of some stibnite

182

(Sb2S3) impurity of the reference as supported by the results of the XRD phase-purity analysis.

183

RESULTS AND DISCUSSION

184

Sb K-edge XAFS data. The K-edge shape (a more or less pronounced white line) and the first

185

inflection point (30496–30497 and 30491–30493 eV for Sb(V) and Sb(III), respectively) were

186

taken from the XANES spectra. The XANES spectra together with their first derivatives of the

187

Sb(III) reference minerals and samples are shown in Figure 1. The spectra confirm that the

188

added Sb remained as Sb(III) and was not oxidized to Sb(V). The Sb K-edge height and shape

189

were identical in the samples and Sb(III) reference material. These results corroborate the

190

voltammetric aqueous speciation analysis results for the batch equilibration experiment

191

solutions (Figure S3, Supporting Information). The EXAFS spectra fits are shown in Figure 2,

192

and the structural results of the least-squares-fitting process for the akaganéite samples loaded

193

with Sb(III) at the three different pH values are summarized in Table 1. The results indicate that

194

the nearest neighbours of the Sb atoms in the samples from the adsorption experiments were O

195

atoms, each at a distance of 1.96 ± 0.02 Å, as expected for Sb(III).23

ACS Paragon Plus Environment

Environmental Science & Technology

9 196

The measured coordination numbers N were in the range of 3.4 – 4.4. These were a little

197

higher than the theoretical N of 3 expected for Sb coordinated to three O atoms (Table 1), but a

198

similar result was found previously for antimonyl surface complexes.24 On the other hand,

199

Sb(V) is coordinated to six O atoms, as shown for the reference mineral tripuhyite in Table S3

200

(Supporting Information) and in previous EXAFS studies of Sb(V) adsorption by goethite.25 As

201

shown in Table 1, the second coordination shell of Sb(III) sorbed to the pure akaganéite samples

202

had an extra feature at R 3.2 Å (not corrected for phase shift), which fitted satisfactorily to a

203

single Sb–Fe path. The Fe backscatterer was fitted to a real distance of r 3.56–3.60 Å with N of

204

3 (Table 1). We therefore conclude that our EXAFS data indicate only one main inner-sphere

205

surface complex in acidic solution. The Sb(III)–Fe path distance, 3.58 ± 0.02 Å, could be

206

interpreted to indicate a bidentate binuclear corner-sharing (2C) complex bridging two edge-

207

sharing FeO3(OH)3 octahedra with one pyramidal Sb(O,OH)3 molecule, as previously found

208

also for ferrihydrite and goethite.24 Schematics of the structural model of this surface species is

209

presented in Figure 3. Although the GFH adsorber material is comprised of the two different

210

solid phases ferrihydrite and akaganéite,16 this does thus not imply that the Sb(III) surface

211

complexes formed are different, at least not in the acidic to circumneutral pH range.

212

Least-squares fitting indicated that the pH 10 samples had another weak Sb–Fe

213

backscattering path at 3.11 ± 0.02 Å (corrected for phase shift) with a N of only ~0.7. The Sb–

214

Fe path distance, commonly interpreted as indicating an inner-sphere bidentate but

215

mononuclear edge-sharing 2E5 surface complex, was found to be required to describe the second

216

coordination shell features of the akaganétite samples conditioned at pH 10. The first-shell

217

oxygen coordination number of four indicates that tetrahedral antimonite-type Sb(OH)4− was

218

the adsorbing species in this additional binding mode. This is in agreement with the antimonite-

219

type complex dominating the dissolved Sb(III) speciation in the alkaline pH range. A schematic

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30

Environmental Science & Technology

10 220

of the structural model of this unique 2E5 surface species is presented in Figure S4 (Supporting

221

Information).

222

Sb(III) Adsorption and Speciation. The adsorption isotherms of Sb(III) on GFH conformed

223

to the Langmuir model (Figure S5, Supporting Information). The maximum adsorption capacity

224

was 120 mg g-1, and the BET surface normalized adsorption capacity was 2.0 molecules per

225

nm2. This adsorption capacity is just half of that of self-assembly {001} TiO2 studied

226

previously,26 and 2/3 of that for pure ferrihydrite of the same nominal (i.e., dried BET) specific

227

surface area.27 Figure 4 shows the effect of pH on Sb(III) adsorption by GFH for different Sb

228

and background electrolyte concentrations. Within the pH range studied, it had a negligible

229

effect on Sb(III) adsorption as evidenced by its consistent top 98% removal in the pH range

230

310 at the lowest Sb concentration series, making it one of the most strongly adsorbing

231

oxyanions on Fe oxyhydroxides. The background electrolyte was found not to markedly affect

232

the adsorption curves for this series. Sb(III) must therefore have been bound predominantly as

233

strong inner-sphere complexes with Fe surface sites, as also suggested by the EXAFS results.

234

The pH does not affect Sb(III) sorption because the sorbate moiety is neutral. Unlike the case

235

with As(III), Sb(III) adsorption even do not decline in the alkaline pH region. This could be

236

attributed to the stronger Lewis base behavior of Sb(III) than As(III), having a higher

237

deprotonation pKa value (pKa1 = 9.17 of H3AsO3, pKa2 = 11.82 of Sb(OH)3). Moreover, as

238

indicated by the EXAFS results, the deprotonated antimonite species Sb(OH)4- occurring at

239

high pH values is adsorbed as well. On the other hand, markedly less adsorption occurred at pH

240