Removal Mechanisms of Phosphate by Lanthanum Hydroxide

Oct 16, 2017 - In spite of the high efficiency of La in phosphate removal and the relatively low acute toxicity,(12) the ecological safety in regard t...
3 downloads 10 Views 2MB Size
Subscriber access provided by Eastern Michigan University | Bruce T. Halle Library

Article

Removal Mechanisms of Phosphate by Lanthanum Hydroxide Nanorods: Investigations using EXAFS, ATRFTIR, DFT and Surface Complexation Modeling Approaches Liping Fang, Qiantao Shi, Jessica Nguyen, Baile WU, Zimeng Wang, and Irene M. C. Lo Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b03803 • Publication Date (Web): 16 Oct 2017 Downloaded from http://pubs.acs.org on October 16, 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 34

Environmental Science & Technology

1

Removal Mechanisms of Phosphate by Lanthanum Hydroxide

2

Nanorods: Investigations using EXAFS, ATR-FTIR, DFT and Surface

3

Complexation Modeling Approaches

4

Liping Fang,† Qiantao Shi,‡ Jessica Nguyen,§ Baile Wu,† Zimeng Wang,§ Irene M.C.

5

Lo*,†

6



7

Science and Technology, Hong Kong, China

8



9

Jersey 07030, USA.

Department of Civil and Environmental Engineering, The Hong Kong University of

Center for Environmental Systems, Stevens Institute of Technology, Hoboken, New

10

§

11

Rouge, LA 70803, USA.

12

*Corresponding Author:

13

Fax: (852) 23581534; email: [email protected].

14

Notes

15

The authors declare no competing financial interest.

Department of Civil and Environmental Engineering, Louisiana State University, Baton

1 ACS Paragon Plus Environment

Environmental Science & Technology

16

ABSTRACT

17

Lanthanum-based materials are effective for sequestering phosphate in water, however

18

their removal mechanisms remain unclear, and the effects of environmentally relevant

19

factors have not yet been studied. Hereby, this study explored the mechanisms of

20

phosphate removal using La(OH)3 by employing extended X-ray absorption spectroscopy

21

(EXAFS), attenuated total reflectance Fourier transform infrared spectroscopy (ATR-

22

FTIR), density functional theory (DFT) and chemical equilibrium modeling. The results

23

showed that surface complexation was the primary mechanism for phosphate removal

24

and in binary phosphate configurations, namely diprotonated bidentate mononuclear

25

(BM-H2) and bidentate binuclear (BB-H2), coexisting on La(OH)3 in acidic conditions.

26

By increasing the pH to 7, BM-H1 and BB-H2 were the two major configurations

27

governing phosphate adsorption on La(OH)3, while BB-H1 was the dominant

28

configuration of phosphate adsorption at pH 9. With increasing phosphate loading, the

29

phosphate configuration of on La(OH)3 transforms from binary BM-H1 and BB-H2 to

30

BB-H1. Amorphous Ca3(PO4)2 forms in the presence of Ca, leading to enhanced

31

phosphate removal at alkaline conditions. The contributions of different mechanisms to

32

the overall phosphate removal were successfully simulated by a chemical equilibrium

33

model that was consistent with the spectroscopic results. This study provides new

34

insights into the molecular-level mechanism of phosphate removal by La(OH)3.

2 ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34

35

Environmental Science & Technology

INTRODUCTION

36

Conventional biological processes have poor ability to remove phosphate from sewage,

37

causing excessive discharge of phosphorus (P) into natural water bodies. As a

38

consequence, this can lead to eutrophication with algal blooms and red tides, damaging

39

the water quality and eventally threatening wildlife and human health.1, 2 Adsorption is an

40

economical and efficient solution for removing phosphate from water,3 and metal

41

(hydr)oxide-based adsorbents are deemed promising for phosphate removal because of

42

their strong affinity for phosphate.4-6 In the past decades, extensive studies have shown

43

that the mechanism of phosphate adsorption on typical minerals like Fe/Al (hydr)oxides

44

is likely through the formation of inner-sphere surface complexes;5,

45

substitute for either one or two hydroxyl groups on the surface of metal (hydr)oxides to

46

form monodentate (MD), bidentate (BD) or coexisting MD/BD binding surface

47

complexes depending on the pH and surface loading.7-9 Structural information (e.g.

48

surface-specific bond distance) obtained from interpreting the surface configuration of

49

phosphate on Fe/Al (hydr)oxides at the molecular-scale, can be useful in predicting the

50

phosphate adsorption with metal (hydro)oxides under different environmental

51

conditions.8

7

phosphate can

52

As one of most promising adsorbents, besides commercialized lanthanum modified

53

bentonite clay (i.e. Phoslock®), lanthanum hydroxide (i.e. La(OH)3) is receiving

54

increasing attention for its high potential in effectively removing phosphate even at trace

55

levels.10,

56

adsorption capacities have been developed to tackle the problem of phosphate-

57

contaminated water.10,

11

To date, a number of lanthanum-based adsorbents with high phosphate

12, 13

Although tremendous efforts have been made to develop 3 ACS Paragon Plus Environment

Environmental Science & Technology

58

lanthanum-based adsorbents for phosphate removal, there has been less attention paid on

59

developing a full understanding of the mechanisms of phosphate adsorption on La(OH)3

60

at the molecular level. Using Fourier transform infrared spectroscopy (FTIR) and

61

macroscopic batch methods, Xie and colleagues reported that phosphate ions form

62

deprotonated MD complexes on La(OH)3 at pH > 8.2.13 Using X-ray diffraction (XRD)

63

and 31P solid-state nuclear magnetic resonance (31P NMR), Zhang et al.10 and Dithmer et

64

al.14 observed from their experiments that a new crystal phase of LaPO4·xH2O was found

65

after phosphate adsorption. However, the exact structural information of phosphate

66

coordination with La(OH)3 remains unclear. In addition, current knowledge about the

67

mechanisms of phosphate adsorption on La(OH)3 has been obtained under different

68

experimental conditions. As pH and surface loading may critically influence the local

69

atomic environment of phosphate,9, 15 a systematic investigation across a wide range of

70

water chemistry parameters is warranted.

71

As an element-specific technique, extended X-ray absorption fine structure (EXAFS)

72

spectroscopy has shown advantages in directly probing the surface configuration of

73

oxyanions (e.g. phosphate, arsenate and sulfate) on mineral surfaces.9, 16-19 Although two

74

research groups both characterized the structure of phosphate complexes on La-bearing

75

adsorbents by applying La K-edge EXAFS,12, 14 it is more straightforward to investigate

76

the mechanisms of phosphate removal by a pure phase of La(OH)3 using P K-edge

77

EXAFS, as this can could give more accurate information on the surface configuration of

78

P with La. Moreover, EXAFS is not able to reflect the protonation states of phosphate on

79

the surface of La(OH)3 under different pHs, without the complementation of other

4 ACS Paragon Plus Environment

Page 4 of 34

Page 5 of 34

Environmental Science & Technology

80

techniques such as attenuated total reflectance (ATR)-FTIR spectroscopy with density

81

functional theory (DFT).20

82

To address the aforementioned research gaps, the objectives of this study were to

83

determine the molecular configurations and protonation states of phosphate in association

84

with La(OH)3 as a function of pH, P loading and co-existing ion concentrations by

85

combining batch sorption experiments, EXAFS, ATR-FTIR, DFT calculation and

86

chemical equilibrium modeling techniques. La(OH)3 were used as the model in this study.

87

Batch studies were first undertaken to investigate how background ions (e.g. Cl, NO3-,

88

HCO3-, SO42-, Ca and Mg), natural organic matter (NOM) and pH influence phosphate

89

adsorption on La(OH)3. ATR-FTIR spectroscopy was used to identify the phosphate-

90

La(OH)3 complexes that formed under different pH values and phosphate loadings, and

91

DFT calculations were used to predict the IR spectra of phosphate after adsorption on

92

La(OH)3 and to compare the predictions compared with experimental IR data. Further,

93

the local structural environment of the phosphate was directly probed under different

94

environmental conditions by means of P K-edge EXAFS. Finally, a generic double layer

95

model incorporating the surface configurations of phosphate and possible precipitation of

96

calcium phosphate based on the multiple techniques was developed, successfully

97

predicting the phosphate removal by La(OH)3 under different environmental conditions.

98

99

MATERIALS AND METHODS

100

Material Synthesis and Characterization. La(OH)3 nanorods were synthesized by

101

precipitating an appropriate amount of analytical-grade La(NO3)3 in NaOH, and

5 ACS Paragon Plus Environment

Environmental Science & Technology

102

hydrothermally treating the suspension at 180 °C for 12 h.21 The white product was

103

washed and dried prior to further use. A detailed description of the synthesis of La(OH)3

104

nanorods is found in the Supporting Information (Text S1). All chemicals used in this

105

study were of analytical grade.

106

The morphology and particle size of the as-synthesized La(OH)3 were determined by

107

transmission electron microscopy (TEM; JEOL JEM-2010, Japan) at an accelerating

108

voltage of 200 kV, and by scanning electron microscopy (SEM; JEOL JSM-6700F,

109

Japan). The structure of La(OH)3 was confirmed by X-ray diffraction (XRD; PANalytical

110

X'pert Pro, Netherlands).

111 112

Batch Experiments. The synthesized La(OH)3 nanorods have an excellent adsorption

113

capacity and a high La use efficiency (Figure S1 and Table S1). Batch studies were

114

conducted to examine the effects of background ions (e.g. Cl-, HCO3-, SO42-, NOM, Ca2+

115

and Mg2+) and pH on the adsorption of phosphate on the as-synthesized La(OH)3. An

116

appropriate amount of phosphate stock solution was spiked into 0.025g/L of the La(OH)3

117

suspension, to achieve an initial phosphate concentration of 0.16 mM. The concentration

118

of background inorganic anions ranged from 0 to 16.7 mM, and the concentration of

119

NOM ranged from 0 to 2.5 mM as C. The pH of the solution was kept at 7.0 ± 0.1, with

120

either NaOH or HCl. The effects of cations (i.e. Ca and Mg) on the adsorption of

121

phosphate were investigated at pH values of 7.0 and 9.0. The effects of pH were

122

investigated with and without the addition of Ca in a pH range of 3.0 to 10.0. The ionic

123

strength of the bath studies was maintained using 0.01 M NaCl. The concentration of

6 ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34

Environmental Science & Technology

124

phosphate was determined using the ammonium molybdate method with a UV/vis

125

spectrometer (Lambda 25, Perkin-Elmer, USA).3

126

EXAFS Analysis. To probe the local coordination environment of phosphorus on the

127

surface of La(OH)3, the extended X-ray fine structure (EXAFS) spectra were obtained in

128

the fluorescence mode at beamline 4B7A of the Beijing Synchrotron Radiation Facility

129

(BSRF; Beijing, China). A Si (111) crystal was used as the monochromator during

130

measurement, and the energy of the electrons was 2.5 GeV in the storage ring. The

131

scanned energy ranged between -200 and 1000 eV from the P K-edge, with minimum

132

energy steps of 0.2 eV and a dwell time of 1 s per point. The EXAFS samples were

133

prepared using a similar protocol to the batch experiments with an initial P concentration

134

of 0.04 mM (0.08 mM for the samples with a higher P loading or in the presence of Ca)

135

in a La(OH)3 dosage of 0.025 g/L at the desired pH values. The obtained samples were

136

dried and directly pressed into tablets and pasted onto a sample holder made from

137

stainless steel. The EXAFS data analysis was performed using Athena and Artemis.17

138

Details of the analysis procedures are given in the Supporting Information (Text S2).

139 140

ATR-FTIR Experiments. In order to obtain information on the protonation state of

141

the loaded phosphate on La(OH)3, the ATR-FTIR spectra of the phosphate-loaded

142

La(OH)3 were recorded on a Bruker Vertex 70 Hyperion 1000 spectrometer (Bruker,

143

Germany) equipped with a platinum diamond ATR crystal accessory and MCT-A

144

detector cooled by liquid N2. The samples of the phosphate-loaded La(OH)3 under

145

different experimental conditions (i.e. pH and surface loading) were prepared using the

146

same protocol as the batch experiments, and their ATR-FTIR spectra were immediately 7 ACS Paragon Plus Environment

Environmental Science & Technology

147

collected at 256 scans per spectrum with a 4 cm-1 resolution. The ATR-FTIR spectra of

148

all samples were subtracted from their corresponding IR spectra of filtrate solution. For

149

the ATR-FTIR spectra of the aqueous phosphate at pH values ranging from 3.0 to 10.0, a

150

drop (16.1 mM) of phosphate solution at given pH conditions was added to the ATR

151

crystal and the IR spectra were collected as the average of 256 scans at a 4 cm-1 resolution.

152

The water spectrum was used as the background and subtracted from each individual

153

spectrum. The second-derivative was used to locate the IR peak position of the phosphate

154

during curve fitting.

155 156

DFT Calculation. To verify the experimental IR spectra of the phosphate with

157

La(OH)3, theoretical calculations of the IR frequency of the surface complexes of the

158

phosphate on the La(OH)3 proceeded using the generalized gradient approximation (GGA)

159

with the function parameterized by Perdew, Burke and Enzerhof (PBE) in the DMol3 of

160

Material Studio 2017 (BIOVIA Inc.). The double numerical plus polarization (DNP)

161

function basis set (basis file: 3.5) was employed with the effective core potentials used as

162

lanthanum valence electron wave functions. The use of the DNP function basis set has

163

been suggested to be more accurate than that of the Gaussian 6-31+G (d,p) basis set.22, 23

164

A conductor-like screening model (COSMO) was applied to simulate the water solvent

165

environment. A simple cluster model of a lanthanum hydroxide unit cell containing an

166

edge-sharing seven-coordinated hexagonal cluster for calculating possible configurations

167

with phosphate was built (Figure S2).17 The size of the cluster in the model has been

168

proven to sufficiently satisfy the basic demands for the calculations of the interfacial

169

configurations and IR frequency of small oxyanions like phosphate complexation on 8 ACS Paragon Plus Environment

Page 8 of 34

Page 9 of 34

Environmental Science & Technology

24

170

minerals.

This approach is also suggested to be more convincing than that using

171

periodic slab models that cannot calculate the vibrational spectra of surface complexes on

172

minerals or metal oxides very well.8, 24

173

Chemical Equilibrium Modeling. A chemical equilibrium model including

174

adsorption and precipitation reactions was developed to simulate different phosphate

175

removal mechanisms. A generic double layer surface complexation model was used to

176

establish the equilibrium between phosphate species in the aqueous phases and on the

177

surface of the La(OH)3. Based on the findings of EXAFS, ATR-FTIR and the DFT

178

calculations, bidentate surface complexes with two different protonation levels (H2 and

179

H1) were employed in the model.25 However, the model doesn’t differentiate

180

mononuclear and binuclear coordination to avoid over fitting the data with excessive

181

fitting parameters. Possible precipitation of calcium and magnesium phosphate and the

182

aqueous speciation reactions were also calculated in the model. The logKs values for the

183

two surface complexation reactions were optimized by minimizing the sum of the

184

residual squares between the experimental and simulated data for both 0.06 and 0.16 mM

185

total P adsorption experiments. The log Ks values for the two precipitation reactions were

186

subsequently optimized by fitting the experimental results with the addition of Ca and

187

Mg, and the obtained constants were close to the values in the literature. The model was

188

implemented by MINEQL+ 5.0 26, with the optimization executed by MINFIT. 27

189 190

RESULTS AND DISCUSSION

191

Characterization of the Synthesized La(OH)3 Nanorods. Figure 1 shows the XRD

192

pattern of the synthesized material, where a highly crystalline structure was observed in

9 ACS Paragon Plus Environment

Environmental Science & Technology

193

line with that for a typical La(OH)3 (JCPDS 83-2034), and is significantly distinct from

194

that of LaPO4. The XRD reflections can be indexed to a hexagonal phase [P63/m(no.176)]

195

of La(OH)3.28 Electron microscopy study confirms the nanorod-like structure of the

196

synthesized La(OH)3. Detailed discussion is given in the Supporting Information (Figure

197

S3).

198

Batch Studies. The competitive adsorption experiment is an intuitive approach to

199

examine the selectivity of phosphate adsorption on La(OH)3, which can indirectly reflect

200

whether inner-sphere complexation occurs when phosphate is adsorbed on La(OH)3,

201

since the adsorption through the outer-sphere complexation is generally affected by

202

background ions like sulfate and bicarbonate in water.

203

presence of different anions such as SO42-, Cl-, HCO3- or natural organic matter (NOM)

204

exhibits a negligible effect on the adsorption of phosphate on La(OH)3 at pH 7.0 ± 0.1,

205

even when the molar ratio of the respective background anion to phosphate is up to 35 for

206

Cl-, 40 for SO42-, 100 for HCO3- and 2.5 for NOM (as C), respectively. It should be noted

207

that the given initial phosphate concentration (i.e. 0.16 mM) exceeds the estimated

208

adsorption capacity of La(OH)3 nanorods (Figure S1). The highly selective adsorption of

209

La(OH)3 for phosphate indicates that dominance of inner-sphere complexes of phosphate

210

with La(OH)3 is likely. 13

29

As shown in Figure S4, the

211

Possible effects of cations (e.g. Ca2+ and Mg2+) on phosphate adsorption were also

212

investigated in batch studies. Figure 2A shows that the presence of Ca can significantly

213

enhance the adsorption of phosphate at pH 9.0 ± 0.1, while a negligible effect was

214

observed for the case of Mg. In addition, the effect of Ca was further investigated

215

together with the pH condition, and the findings are given in Figure 2B. The results show 10 ACS Paragon Plus Environment

Page 10 of 34

Page 11 of 34

Environmental Science & Technology

216

that phosphate adsorption gradually decreases from 6.45 to 1.5 mmol/g with the pH

217

increasing from 4 to 10 when more phosphate than can be removed by La(OH)3 nanorods

218

is present in the water (e.g. 0.16 mM). The presence of Ca (2.5 mM) shows insignificant

219

effects in acidic to neutral conditions, while a dramatic increase of phosphate adsorption

220

was found at pH > 7.5 (Figure 2B). This is in line with the findings reported by Lin and

221

colleagues, where notably enhanced adsorption of phosphate on zirconium oxide was also

222

observed.30 They suggested that this was due to the occurrence of co-precipitation of

223

CaHPO4.

224

It is well known that Ca can precipitate phosphate from neutral to alkaline conditions

225

when the saturation index (SI) of calcium and phosphate-based minerals exceeds 0.

226

Hence, a simple calculation was carried out to predict the possible precipitation under our

227

experimental conditions using MINEQL+ 5.0.

228

SI values of several calcium-based minerals, including hydroxyapatite, amorphous

229

Ca3(PO4)2 and β- Ca3(PO4)2, exceed 0 at pH ≥ 7, amorphous Ca3(PO4)2 appears to have a

230

high possibility of being present in our system (Figure S5). This is also confirmed by

231

XRD analysis, showing that no new crystalline minerals such as hydroxyapatite or

232

brushite formed at pH 9 (Figure S6).31 In addition, we also found about 4.6% of total Ca

233

lost from aqueous solution after the reaction of phosphate with La(OH)3 at pH 9 (Figure

234

2B). XPS analysis indicates that the Ca 2p spectrum of the solid sample obtained after

235

sorption experiments at pH 9 has a comparably binding energy of 347.6 eV, which is in

236

line with that for amorphous Ca3(PO4)2 (Figure S7).32 By combining this evidence, we

237

conclude that the enhanced sorption of phosphate in the presence of Ca in alkaline

238

conditions is likely due to the co-precipitation of amorphous Ca3(PO4)2.

26

Despite the prediction showing that the

11 ACS Paragon Plus Environment

Environmental Science & Technology

239

Phosphorus K-edge EXAFS Analysis. EXAFS was employed to directly probe the

240

local coordination environment of P at the surface of La(OH)3 nanorods under different

241

environmental conditions, such as pH and surface loadings of phosphate. The Fourier

242

transform (FT) EXAFS spectra of P K-edge for the phosphate-loaded La(OH)3 are given

243

in Figure 4 and the corresponding fitted parameters are listed in Table 1.

244

The EXAFS fitting results show a significant oscillation at a distance of 1.52 to 1.57 Å

245

for all samples, and this can be assigned to the first shell of P coordination with 4.0

246

oxygen atoms. This P-O bonding distance is closely in line with previous findings by P

247

K-edge EXAFS fitting (i.e. 1.51 – 1.52 Å),9 and DFT calculation (i.e. 1.56 – 1.57 Å). 8, 33

248

Two P-La distances were obtained from the EXAFS analysis for the La(OH)3 sample

249

after phosphate adsorption, i.e. a distance of 2.73 to 2.89 Å, and 3.17 – 3.21 Å,

250

respectively (Figure 3 and Table 1). Interestingly, the coordination of P-La with two

251

different distances only appears at pH values ranging from 3 to 7 when the surface

252

loading of P is 1.6 mmol/g, however, at pH 9 only the coordination of P-La with a

253

distance of 3.19 Å from P is present (Figure 3 and Table 1). Besides, the value of the

254

coordination number (CN) can well reflect the surface configuration of P with La(OH)3.

255

As shown in Table 1, both bidentate binuclear and bidentate mononuclear inner-sphere

256

complexes of P on La (OH)3 exist at pH values from 3 to 7, but only the bidentate

257

binuclear (BB) surface complex forms at pH 9. The shortest P-La distances of 2.73 to

258

2.89 Å are probably due to the bidentate mononuclear (BM; 1V/2E) configuration that

259

forms at pH between 3 and 7 with a P loading of 1.6 mmol/g, since these P-La distances

260

are too short for the monodentate mononuclear configuration. The BM configuration of P

261

on La(OH)3 is expected to be reasonable as similar surface complexation has been 12 ACS Paragon Plus Environment

Page 12 of 34

Page 13 of 34

Environmental Science & Technology

262

frequently observed for P and As(V) (analogue of P) on metal (hydr)oxides.9, 34, 35 Our

263

results are partially in agreement with Xu and colleagues

264

EXAFS finding of formation of BB surface complexes of phosphate with

265

lanthanum/alumina. Further, our EXAFS data also do not support the formation of

266

lanthanum phosphate since the reported value of 4.2 Å for the La-P distance of LaPO4 is

267

too large compared with that in our study (Table 1). In addition, the results also suggest

268

that the increase of P loading on La(OH)3 from 2.5 to 5 mM leads to a BB configuration

269

of P with La at pH 7 (Figure 3 and Table 1), showing a similar pattern for the P

270

configuration with iron hydroxide reported by Abdala and colleagues.9 Moreover, the

271

EXAFS spectra show that the distance of La-P decreases from 3.19 to 2.76 Å after the

272

addition of Ca at pH 9, and the CN value changes from 1.9 to 1.3, which indicates the

273

possible transformation of the surface complexation of P with La(OH)3 from a BB to a

274

BM configuration (Figure 3). This is likely due to the co-precipitation of amorphous

275

Ca3(PO4)2 by Ca2+, leading to a lower P loading on the La(OH)3 (Figure 2B). Oscillation

276

accounting for Ca-P was not observed in the sample, and is probably due to the weaker

277

backscattering of Ca than La.

12

regarding the La L-edge

278

279

ATR-FTIR Analysis and DFT Calculation. The protonation status of the P

280

complexation on La(OH)3 is further crucial information for predicting P adsorption

281

behavior.

282

analysis is a powerful tool to complement such limitation. The ATR-FTIR spectra of the

283

dissolved phosphate and the phosphate-loaded La(OH)3 under different pH conditions are

284

depicted in Figure S8 and Figure 4A, where typical P-O fingerprint features are observed

17, 36

However, EXAFS is unable to identify these data, while ATR-FTIR

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 34

285

in the frequency range of 900 – 1200 cm-1. Frequencies below 900 cm-1 are not

286

considered due to overlap with the IR background of La(OH)3.

287

The IR spectra of the P-La(OH)3 are markedly different from those of the dissolved P

288

species (i.e. H3PO4, H2PO4- and HPO42-) (Figure 4A, Figures S8 and S9), indicating the

289

formation of inner-sphere complexes for P-La(OH)3.24 With the increase of pH from 3 to

290

9, a significant peak shift to a lower frequency is observed, due to the change of each

291

characteristic band as a result of protonation/deprotonation change of the loaded P on

292

La(OH)3.17 Each band in the P-O fingerprint range of 900 – 1200 cm-1 can be resolved

293

through curve fitting and then compared with the calculated IR frequencies. As shown in

294

Figure 5A the fitted ATR-FTIR spectrum exhibits vibrational modes at 1102, 1073, 1046,

295

1018, 991, 956 and 918 cm-1, for the P-La(OH)3 at a pH of 3. A further increase of pH to

296

5 results in one distinct group of vibrational modes at 1078, 1045, 1008, 962 and 917 cm-

297

1

298

relative peak intensity at 1006 to 1046 cm-1 is significantly enhanced at pH 7 (Figure 4A).

299

At pH 9, the fitted IR spectrum exhibits distinct vibrational modes at 1075, 1048, 1017,

300

988 cm-1 and very weak bands at 955 and 918 cm-1 (Figure 4A). In the meantime, there is

301

no data on the characteristic peak positions for P adsorption on La(OH)3, so direct

302

assignment of these vibrational modes for P-La(OH)3 is difficult. Hence, DFT calculation

303

was further applied to predict the IR frequencies of P-La(OH)3 with different

304

configurations. On the basis of the EXAFS findings (Figure 3 and Table 1), six surface

305

configurations

306

mononuclear (BM-H0, BM-H1 and BM-H2), deprotonated, monoprotonated and

307

diprotonated bidentate binuclear (BB-H0, BB-H1 and BB-H2)] are considered in the DFT

, while similar vibrational modes appear at pH 7. In comparison to that at pH 5, the

[i.e.

deprotonated,

monoprotonated

and

14 ACS Paragon Plus Environment

diprotonated

bidentate

Page 15 of 34

Environmental Science & Technology

308

calculation (Figure S10). As shown in Figure S11, DFT calculation suggests that none of

309

the individual surface complex configurations can reproduce all the observed IR peaks

310

for the P-La(OH)3 at pH 3 to 7 (R2 ranges from 0.958 to 0.994). The results indicate that

311

a combination of BM-H2 with BB-H2 satisfactorily correlates with the observed IR

312

frequencies for the P-La(OH)3 at pH 3, and the combination of BM-H1 with BB-H2

313

complexes matches with that at pH 5 and 7. The formation of the BB-H2 configuration in

314

the pH range of 3 to 7 is reasonable, and is in line with a previous study by Shi et al.17

315

who reported a similar finding for As(V) (analogue of phosphate) adsorption on La(OH)3.

316

All of this evidence indicates that pH is a decisive factor on the protonation status of

317

interfacial P configurations. In addition, the presence of two P-La(OH)3 surface

318

complexes at pH from 3 to 7 is also in line with our EXAFS findings (Figure 3 and Table

319

1), which is not unusual and has been extensively reported in the literature for the P

320

configuration on iron oxides.8, 24, 37 Nevertheless, at pH 9, the calculated IR frequencies

321

of the BM-H1 or BB-H1 configuration is found highly correlated with the fitted IR

322

frequencies derived from the experimental IR spectra of the samples (R2=0.941, Figure

323

S11). In combination with our EXAFS finding, our data suggest that BB-H1 likely

324

governs the surface configuration of P on La(OH)3 at pH 9 (Figure 3 and Table 1).

325

The effect of P loading on the P configuration was also investigated by ATR-FTIR,

326

and the corresponding results are depicted in Figures 4 and S12. The results show that the

327

peak intensity increases with increasing P loading, which agrees with the finding of the

328

batch studies (Figure S1). Further, the fitted IR spectrum of the lowest P loading (i.e. 0.4

329

mmol/g) exhibits more distinct vibrational modes at 1020 and 988 cm-1 than those with

330

higher P loadings (i.e. 1007 – 1009 cm-1, Figure 4B). The unique IR vibrational modes at 15 ACS Paragon Plus Environment

Environmental Science & Technology

331

1020 and 988 cm-1 likely correspond to the BB-H1 configuration because of their high

332

correlation with the calculated IR frequencies of BB-H1 (Figure S13). With the increase

333

of P loading, the vibrational modes at 1020 and 988 cm-1 shift to a single vibrational

334

mode at about 1008 cm-1 that may be due to the formation of a BB-H2 or BM-H1

335

configuration (Figure 4B). Interestingly, Abdala et al. 9 reported a similar finding, where

336

P formed a BB configuration on goethite at low P loading and then transformed to a BM

337

and subsequently, a BB configuration with increase of P loading. These findings are also

338

in accordance with the above EXAFS results, showing that the BB configuration of

339

phosphate formation occurred with an increase of P loading (Figure 3 and Table 1). In

340

order to examine the possible transformation of the crystal structure reported by other

341

groups,10, 14 XRD analysis was carried out to confirm the crystal structure of La(OH)3

342

after P adsorption at pH 9. As shown in Figure S14, the nanorod-like structure of

343

La(OH)3 still remains after loading different amounts of P on the surface La(OH)3 and the

344

crystal structure of P-La(OH)3 is markedly different from that of LaPO4 (Figure 1).38

345

With the increase of Ca concentration from 0 to 5 mM, the ATR-FTIR spectra also show

346

significantly enhanced peak intensities within the range of 900 – 1200 cm-1 (Figure S15),

347

due to the enhanced P removal.39 The IR spectra of the P-La(OH)3 samples became

348

broader after the introduction of Ca, which is likely due to the presence of an amorphous

349

Ca3(PO4)2 phase.40 This has also been demonstrated by our theoretical calculation and

350

XPS study (Figures S5 and S7).

351

Chemical Equilibrium Modeling. The surface complexation model successfully

352

predicts the variation of surface complex speciation of phosphate on La(OH)3 as a

353

function of pH, P loading and the absence and presence of Ca (Figure 2B and Table S2). 16 ACS Paragon Plus Environment

Page 16 of 34

Page 17 of 34

Environmental Science & Technology

354

The predicted protonation levels of the surface complexes from acid, neutral to alkaline

355

pH, as well as at excessively high phosphate concentration, followed the

356

spectroscopically and computationally determined results (Figure S16). The model

357

predicted the formation of amorphous Ca3(PO4)2 when the combination of pH and the

358

concentration of Ca enable its saturation (Figure 2, Figures S5 and S16). Precipitation

359

enhances the overall removal of the dissolved phosphate ions and may also outcompete

360

the adsorption contribution, leading to the transformation of a BB to BM configuration as

361

shown in our EXAFS data (Table 1).41 The model also shows a fine prediction of

362

phosphate removal by La(OH)3 in the presence of Ca and Mg at pH 9.0 ± 0.1 (Figure 2A);

363

magnesium phosphate has higher solubility than calcium phosphate, and the model only

364

predicted minor precipitation for the highest (i.e. 8 mM) Mg concentration. Overall, the

365

proposed mechanisms of adsorption and precipitation for phosphate removal are

366

quantitatively verified by the surface complexation model. Our results indicate that the

367

model has the potential to accurately extrapolate the removal of phosphate by La(OH)3

368

under different environmental conditions, such as pH, P loading and the presence of Ca

369

and Mg (Figure S16).

370

371

Environmental Significance. Given the important role of La-based sorbents for

372

managing phosphate in eutrophic lake and sewage effluents ,12, 14 this work provides new

373

insights into the molecular-level removal mechanism of phosphate by means of batch

374

studies, EXAFS, ATR-FTIR and DFT calculations (Figure 6). The combination of

375

multiple spectroscopic techniques with theoretical calculations (e.g. DFT) allows us to

376

build the entire picture of surface complexation of phosphate with La(OH)3, and provides 17 ACS Paragon Plus Environment

Environmental Science & Technology

377

a reliable approach in future efforts to probe the interfacial chemistry of other

378

oxyanions.36 Our findings on the P configurations on La(OH)3 and possible precipitation

379

of phosphate in the presence of Ca or Mg give concrete evidence for developing surface

380

complexation models, which can be used to predict phosphate removal using La(OH)3

381

during lake restoration or sewage treatment. The prediction of phosphate removal should

382

consider specific environmental variables such pH values (e.g. pH from 5 to 9 for natural

383

lakes, and about 7.5 for sewage),3, 42 surface P loading, and the presence of Ca and Mg at

384

high pH values (e.g. pH 9). In spite of the high efficiency of La in phosphate removal and

385

the relatively low acute toxicity,12 the ecological safety in regard to the extensive

386

application of La-based sorbents requires caution, in particular, the long-term effects on

387

aquatic organisms should be carefully evaluated in future work.

388

389

Supporting Information

390

Detailed synthesis of La(OH)3 nanorods, EXAFS data collection and analysis,

391

adsorption of phosphate on La(OH)3, comparison of La use efficiencies of La-based

392

sorbents, cluster structure of La(OH)3, electron microscopic images of La(OH)3,

393

competitive adsorption of background anions with phosphate, calculated saturation

394

indexes of calcium phosphate species, X-ray diffractograms of La(OH)3, Ca 2p XPS

395

spectra, ATR-FTIR spectra of aqueous phosphate and phosphate loaded La(OH)3,

396

phosphate species distribution, configurations of P on La(OH)3, correlations of the

397

calculated and experimental IR frequencies of phosphate, reaction equations for surface

18 ACS Paragon Plus Environment

Page 18 of 34

Page 19 of 34

Environmental Science & Technology

398

complexation modelling, comparison of experimental results with surface complexation

399

modelling results, are provided in the Supporting Information.

400

401

ACKNOWLEDGEMENTS

402

This work was supported financially by the Research Grants Council of Hong Kong

403

(GRF16207916; T21-711/16-R-1). The authors are thankful for the technical support

404

from the Advanced Engineering Material Facility of the Hong Kong University of

405

Science and Technology (AEMF-HKUST). EXAFS beam time was granted by the 4B7A

406

endstation of the Beijing Synchrotron Radiation Facility at the Institute of High Energy

407

Physics, Chinese Academy of Sciences. Dr. Lei Zheng from 4B7A beamline is gratefully

408

acknowledged for his support in the measurements and data reduction. Jessica Nguyen

409

was supported by a Huel Perkins Diversity Fellowship.

19 ACS Paragon Plus Environment

Environmental Science & Technology

410

REFERENCES

411 412

1.

Mayer, B. K.; Baker, L. A.; Boyer, T. H.; Drechsel, P.; Gifford, M.; Hanjra, M. A.;

413

Parameswaran, P.; Stoltzfus, J.; Westerhoff, P.; Rittmann, B. E., Total Value of

414

Phosphorus Recovery. Environ Sci Technol 2016, 50, (13), 6606-6620.

415

2.

416

removal in activated sludge systems. Fems Microbiol Rev 2003, 27, (1), 99-127.

417

3.

418

ZrO2@Fe3O4 with enhanced phosphate recovery from sewage: Performance and

419

adsorption mechanism. Chem Eng J 2017, 319, 258-267.

420

4.

421

upscaled synthesis of layered iron oxide nanosheets and their application in phosphate

422

removal. J Mater Chem A 2015, 3, (14), 7505-7512.

423

5.

424

Hydroxide Coprecipitates. Environ Sci Technol 2011, 45, (15), 6283-6289.

425

6.

426

by Fresh and Aged Aluminum Hydroxide. Consequences for Lake Restoration. Environ

427

Sci Technol 2008, 42, (17), 6650-6655.

428

7.

429

Spectroscopic Study of Phosphate Sorption Mechanisms on Aluminum (Hydr)oxides.

430

Environ Sci Technol 2013, 47, (15), 8308-8315.

431

8.

432

Pierre-Louis, A.-M.; Strongin, D. R., ATR–FTIR and Density Functional Theory Study

Seviour, R. J.; Mino, T.; Onuki, M., The microbiology of biological phosphorus

Fang, L.; Wu, B.; Lo, I. M. C., Fabrication of silica-free superparamagnetic

Fang, L.; Huang, L.; Holm, P. E.; Yang, X.; Hansen, H. C. B.; Wang, D., Facile

Liu, Y.-T.; Hesterberg, D., Phosphate Bonding on Noncrystalline Al/Fe-

de Vicente, I.; Huang, P.; Andersen, F. Ø.; Jensen, H. S., Phosphate Adsorption

Li, W.; Feng, X.; Yan, Y.; Sparks, D. L.; Phillips, B. L., Solid-State NMR

Kubicki, J. D.; Paul, K. W.; Kabalan, L.; Zhu, Q.; Mrozik, M. K.; Aryanpour, M.;

20 ACS Paragon Plus Environment

Page 20 of 34

Page 21 of 34

Environmental Science & Technology

433

of the Structures, Energetics, and Vibrational Spectra of Phosphate Adsorbed onto

434

Goethite. Langmuir 2012, 28, (41), 14573-14587.

435

9.

436

orthophosphate surface complexation at the goethite/water interface as examined by

437

extended X-ray Absorption Fine Structure (EXAFS) spectroscopy. J Colloid Interf Sci

438

2015, 437, 297-303.

439

10.

440

Nanosized Hydrated La(III) Oxide Confined in Cross-linked Polystyrene Networks.

441

Environ Sci Technol 2016, 50, (3), 1447–1454.

442

11.

443

arsenate ions on lanthanum impregnated silica gel. Water Environ Res 1996, 68, (3), 295-

444

300.

445

12.

446

Novel Lanthanum/Aluminum-Hydroxide Composite: Implication for Eutrophication

447

Control. Environ Sci Technol 2017, 51, (6), 3418–3425.

448

13.

449

recovery of phosphate from water by lanthanum hydroxide materials. Chem Eng J 2014,

450

254, 163-170.

451

14.

452

G., Characterization of Phosphate Sequestration by a Lanthanum Modified Bentonite

453

Clay: A Solid-State NMR, EXAFS, and PXRD Study. Environ Sci Technol 2015, 49, (7),

454

4559-4566.

Abdala, D. B.; Northrup, P. A.; Arai, Y.; Sparks, D. L., Surface loading effects on

Zhang, Y.; Pan, B.; Shan, C.; Gao, X., Enhanced Phosphate Removal by

Wasay, S. A.; Haron, J.; Tokunaga, S., Adsorption of fluoride, phosphate, and

Xu, R.; Zhang, M.; Mortimer, R. J. G.; Pan, G., Enhanced Phosphorus Locking by

Xie, J.; Wang, Z.; Lu, S. Y.; Wu, D. Y.; Zhang, Z. J.; Kong, H. N., Removal and

Dithmer, L.; Lipton, A. S.; Reitzel, K.; Warner, T. E.; Lundberg, D.; Nielsen, U.

21 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 34

455

15.

Luengo, C.; Brigante, M.; Antelo, J.; Avena, M., Kinetics of phosphate adsorption

456

on goethite: Comparing batch adsorption and ATR-IR measurements. J Colloid Interf Sci

457

2006, 300, (2), 511-518.

458

16.

459

Degradation

460

Peroxymonosulfate: Synergistic Effects and Mechanisms. Environ Sci Technol 2016, 50,

461

(6), 3119-3127.

462

17.

463

Impregnated Activated Alumina: Spectroscopic and DFT Study. Acs Appl Mater Inter

464

2015, 7, (48), 26735-26741.

465

18.

466

spectroscopic quantification and speciation modeling of sulfate adsorption on ferrihydrite

467

surfaces. Environ Sci Technol 2016, 50, (15), 8067-8076.

468

19.

469

Catalano, J. G., Effect of phosphate on U (VI) sorption to montmorillonite: Ternary

470

complexation and precipitation barriers. Geochim Cosmochim Ac 2016, 175, 86-99.

471

20.

472

with phosphate and sulfate on iron oxide surfaces. Geochim Cosmochim Ac 2015, 158,

473

130-146.

474

21.

475

photoluminescence of lanthanum hydroxide nanorods by a simple route at room

476

temperature. Nanotechnology 2009, 20, (34), 345602-345609.

Feng, Y.; Wu, D.; Deng, Y.; Zhang, T.; Shih, K., Sulfate Radical-Mediated of

Sulfadiazine

by

CuFeO2

Rhombohedral

Crystal-Catalyzed

Shi, Q.; Yan, L.; Chan, T.; Jing, C., Arsenic Adsorption on Lanthanum-

Gu, C.; Wang, Z.; Kubicki, J. D.; Wang, X.; Zhu, M., X-ray absorption

Troyer, L. D.; Maillot, F.; Wang, Z.; Wang, Z.; Mehta, V. S.; Giammar, D. E.;

Hinkle, M. A.; Wang, Z.; Giammar, D. E.; Catalano, J. G., Interaction of Fe (II)

Mu, Q. Y.; Chen, T.; Wang, Y. D., Synthesis, characterization and

22 ACS Paragon Plus Environment

Page 23 of 34

Environmental Science & Technology

477

22.

Zhang, M.; He, G.; Pan, G., Structure and stability of arsenate adsorbed on α-

478

Al2O3 single-crystal surfaces investigated using grazing-incidence EXAFS measurement

479

and DFT calculation. Chem Geol 2014, 389, 104-109.

480

23.

481

energies of hydrogen bonded complexes: Evidence of small basis set superposition error

482

compared to Gaussian basis sets. J Comput Chem 2008, 29, (2), 225-232.

483

24.

484

Aspartate Adsorption on Goethite and Competition with Phosphate. Environ Sci Technol

485

2016, 50, (6), 2938-2945.

486

25.

487

surface complexation modeling: Theory and practice. Environ Sci Technol 2013, 47, (9),

488

3982-3996.

489

26.

490

system, version 5.0, Environmental Research Software: Hallowell, ME, 2016.

491

27.

492

parameter estimation in an equilibrium speciation software program. Environ. Sci.

493

Technol. 2016, 50, (20), 11112-11120.

494

28.

495

Single-Crystal Nanowires. Angewandte Chemie International Edition 2002, 41, (24),

496

4790-4793.

497

29.

498

C.; Sextl, G.; Franzreb, M.; Steinmetz, H., Phosphate recovery from wastewater using

Inada, Y.; Orita, H., Efficiency of numerical basis sets for predicting the binding

Yang, Y.; Wang, S.; Xu, Y.; Zheng, B.; Liu, J., Molecular-Scale Study of

Wang, Z.; Giammar, D. E., Mass action expressions for bidentate adsorption in

Schecher, W. D.; McAvoy, D. C. MINEQL+: A chemical equilibrium modeling

Xie, X.; Giammar, D. E.; Wang, Z., MINFIT: A spreadsheet-based tool for

Wang, X.; Li, Y., Synthesis and Characterization of Lanthanide Hydroxide

Drenkova-Tuhtan, A.; Mandel, K.; Paulus, A.; Meyer, C.; Hutter, F.; Gellermann,

23 ACS Paragon Plus Environment

Environmental Science & Technology

499

engineered superparamagnetic particles modified with layered double hydroxide ion

500

exchangers. Water Res 2013, 47, (15), 5670-5677.

501

30.

502

calcium ion on phosphate adsorption onto hydrous zirconium oxide. Chem Eng J 2017,

503

309, 118-129.

504

31.

505

precipitation in aerobic granular sludge process. Water Res 2011, 45, (12), 3776-3786.

506

32.

507

Tech 1995, 55, (3), 311-314.

508

33.

509

Complex Structures of Phosphates to Iron Hydroxides:  Calculation of Vibrational

510

Frequencies and Adsorption Energies. Langmuir 2004, 20, (21), 9249-9254.

511

34.

512

Analysis of Arsenite Adsorption onto Two-Line Ferrihydrite, Hematite, Goethite, and

513

Lepidocrocite. Environ Sci Technol 2005, 39, (23), 9147-9155.

514

35.

515

Adsorption and Surface Speciation at the Hematite−Water Interface. Environ Sci Technol

516

2004, 38, (3), 817-824.

517

36.

518

for Phosphorus Speciation in Soils and Other Environmental Systems All rights reserved.

519

No part of this periodical may be reproduced or transmitted in any form or by any means,

520

electronic or mechanical, including photocopying, recording, or any information storage

Lin, J.; Zhan, Y.; Wang, H.; Chu, M.; Wang, C.; He, Y.; Wang, X., Effect of

Mañas, A.; Biscans, B.; Spérandio, M., Biologically induced phosphorus

Demri, B.; Muster, D., XPS study of some calcium compounds. J Mater Process

Kwon, K. D.; Kubicki, J. D., Molecular Orbital Theory Study on Surface

Ona-Nguema, G.; Morin, G.; Juillot, F.; Calas, G.; Brown, G. E., EXAFS

Arai, Y.; Sparks, D. L.; Davis, J. A., Effects of Dissolved Carbonate on Arsenate

Kizewski, F.; Liu, Y.-T.; Morris, A.; Hesterberg, D., Spectroscopic Approaches

24 ACS Paragon Plus Environment

Page 24 of 34

Page 25 of 34

Environmental Science & Technology

521

and retrieval system, without permission in writing from the publisher. J Environ Qual

522

2011, 40, (3), 751-766.

523

37.

524

Phillips, B. L., Molecular level investigations of phosphate sorption on corundum (α-

525

Al2O3) by 31P solid state NMR, ATR-FTIR and quantum chemical calculation. Geochim

526

Cosmochim Ac 2013, 107, 252-266.

527

38.

528

over inorganic adsorbents derived from lanthanum metal organic frameworks. Chem Eng

529

J 2017, 326, (Supplement C), 1086-1094.

530

39.

531

magnetite; rate of adsorption, surface speciation, and effect of calcium ions. J Colloid

532

Interf Sci 2009, 333, (1), 27-32.

533

40.

534

phases formed during induction of mineralization by collagenase-released matrix vesicles

535

in vitro. Journal of Biological Chemistry 1988, 263, (27), 13718-13724.

536

41.

537

U (IV) adsorption to metal oxide minerals. Environmental Science & Technology Letters

538

2015, 2, (8), 227-232.

539

42.

540

71, (4), 449-482.

Li, W.; Pierre-Louis, A.-M.; Kwon, K. D.; Kubicki, J. D.; Strongin, D. R.;

Zhang, X.; Sun, F.; He, J.; Xu, H.; Cui, F.; Wang, W., Robust phosphate capture

Roonasi, P.; Holmgren, A., An ATR-FTIR study of sulphate sorption on

Sauer, G.; Wuthier, R., Fourier transform infrared characterization of mineral

Wang, Z.; Ulrich, K.-U.; Pan, C.; Giammar, D. E., Measurement and modeling of

Khan, F. A.; Ansari, A. A., Eutrophication: An ecological vision. Bot Rev 2005,

541 542

25 ACS Paragon Plus Environment

Environmental Science & Technology

543

TOC/graphic abstract

544

26 ACS Paragon Plus Environment

Page 26 of 34

Page 27 of 34

545

Environmental Science & Technology

Table 1. Structure parameters derived from P K-edge EXAFS analysis. Samples Path CN a R (Å) b σ2 (Å2) c ∆E0 (eV) d

(a) pH 3

(b) pH 5

(c) pH 7_1i

(d) pH 7_2i

(e) pH 9.0

(f) pH 9+Ca

546 547 548 549

a

P-O

4.0f

1.52 (3)

0.005 (1)

P-Lag

1.4 (7)

2.73 (13)

0.014 (8)

P-Lah

2.0 (1)

3.21 (18)

0.025 (9)

P-O

4.0 f

1.52 (2)

0.003 (2)

P-Lag

1.0 (4)

2.89 (13)

0.012 (7)

P-Lah

2.0 (4)

3.18 (16)

0.013 (6)

P-O

4.0 f

1.57 (2)

0.004 (3)

1.2 (4)

2.85 (14)

0.006 (3)

P-Lah

2.0 (1)

3.21 (5)

0.006 (5)

P-O

4.0 f

1.52 (1)

0.002 (1)

P-Lag

-

-

-

P-Lah

1.9 (3)

3.17 (4)

0.004 (3)

P-O

4.0 f

1.53 (2)

0.003 (1)

P-Lag

-

-

-

P-Lah

1.9 (4)

3.19 (5)

0.013 (5)

P-O

4.0 f

1.53 (2)

0.007 (1)

P-Lag

1.3 (4)

2.76 (9)

0.014 (3)

P-Lah

-

-

-

P-Lag

b

c

R-factor e

2.5

0.018

0.7

0.027

13.3

0.016

2.4

0.023

9.1

0.030

3.6

0.029

coordination number. interatomic distance. Debye-Waller factor. dthreshold energy shift. egoodness-of-fit parameter: R-factor = Σ(χdata – χfit)2/Σ(χdata)2. fthe parameters were fixed during fitting. Parentheses: the estimated parameter uncertainties are listed in parentheses, representing the errors in the last digit. gP-La distance represents a bidentate 27 ACS Paragon Plus Environment

Environmental Science & Technology

550 551 552

mononuclear configuration (BM). hP-La distance represents a bidentate binuclear configuration (BB). i pH 7_1 represents the sample with lower P loading of 1.6 mmol/g, while pH 7_2 represents the sample with a higher P loading of 3.2 mmol/g.

553

28 ACS Paragon Plus Environment

Page 28 of 34

Page 29 of 34

Environmental Science & Technology

554 555 556

Figure 1. X-ray diffraction (XRD) pattern of the as-synthesized La(OH)3 before and after loading different amounts of phosphate at pH 7.0.

557

29 ACS Paragon Plus Environment

Environmental Science & Technology

558

559 560 561 562 563 564 565 566

Figure 2. The effect of Mg and Ca on the removal of phosphate by La(OH)3 at pH 9.0 ± 0.1 (A). The initial concentration of phosphate was 0.16 mM, and the dosage of La(OH)3 was 0.025 g/L. The band area represents the simulated phosphate removal by La(OH)3 at pH 9.0 ± 0.1 in the presence of different concentrations of Ca or Mg; the experimental results and surface complexation modeling results for the effects of pH on the phosphate removal by La(OH)3 in the presence and absence of Ca (B). The initial phosphate concentration was 0.08 and 0.16 mM, and the dosage of La(OH)3 was 0.025 g/L.

567

30 ACS Paragon Plus Environment

Page 30 of 34

Page 31 of 34

Environmental Science & Technology

568

569 570 571 572 573 574 575

Figure 3. Normalized k2-weighted experimental (dots) and their fitted (lines) Phosphorus (P) K-edge EXAFS spectra of phosphate removal by La(OH)3 (A), the corresponding Fourier transformed magnitude (B), and real parts of the Fourier transform. Data sets represent the samples of P-La(OH)3 with P loading of 1.6 mmol/g at pH 3 (a), pH 5 (b), pH 7 (c), pH 7 with a higher P loading of 3.2 mmol/g (d), pH 9 (e), and in the presence of Ca (5 mM) at pH 9.0 (f). The corresponding parameters are listed in Table 1.

576

31 ACS Paragon Plus Environment

Environmental Science & Technology

577

578 579 580

Figure 4. ATR-FTIR spectra of the phosphate loaded La(OH)3 at pH 3, pH 5, pH 7 and pH 9 (A), and the phosphate loaded La(OH)3 with a phosphate loading of 0.4, 1.6, 3.2 32 ACS Paragon Plus Environment

Page 32 of 34

Page 33 of 34

581 582 583

Environmental Science & Technology

and 5.5 mmol/g at pH 7 (B). Slurries were obtained by performing phosphate adsorption experiments at a La(OH)3 dosage of 0.025 g/L and an initial phosphate concentration of 0.16 mM.

584

`

585 586 587 588 589

Figure 5. ATR-FTIR spectra of the phosphate loaded La(OH)3 at pH 9.0 in the presence of 5 mM and 2.5 mM (A-B), and the absence (C) of Ca. Slurries were obtained by performing phosphate adsorption experiments at a La(OH)3 dosage of 0.025 g/L and an initial phosphate concentration of 0.16 mM at pH 9.0.

590

33 ACS Paragon Plus Environment

Environmental Science & Technology

591

592 593 594

Figure 6. A schematic diagram of a potential surface complexation of phosphate on La(OH)3 under different pH conditions.

595 596

34 ACS Paragon Plus Environment

Page 34 of 34