Enhanced adsorption of p-arsanilic acid from water ... - ACS Publications

modified UiO-67 as examined using EXAFS, XPS and DFT. 2 calculations. 3 ... Restoration in Industry Clusters (Ministry of Education), Guangdong Engine...
0 downloads 10 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Article

Enhanced adsorption of p-arsanilic acid from water by amine modified UiO-67 as examined using EXAFS, XPS and DFT calculations Chen Tian, Jian Zhao, Xinwen Ou, Jieting Wan, Yuepeng Cai, Zhang Lin, Zhi Dang, and Baoshan Xing Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05761 • Publication Date (Web): 31 Jan 2018 Downloaded from http://pubs.acs.org on January 31, 2018

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 36

Environmental Science & Technology

1

Enhanced adsorption of p-arsanilic acid from water by amine

2

modified UiO-67 as examined using EXAFS, XPS and DFT

3

calculations

4

Chen Tian, †,§ Jian Zhao, ‡,§ Xinwen Ou, † Jieting Wan, † Yuepeng Cai, ‖ Zhang Lin, *,†

5

Zhi Dang, † and Baoshan Xing

6



7

Restoration in Industry Clusters (Ministry of Education), Guangdong Engineering and Technology

8

Research Center for Environmental Nanomaterials, South China University of Technology,

9

Guangzhou 510006, China



School of Environment and Energy, The Key Laboratory of Pollution Control and Ecosystem

10



11

Science and Ecology (Ministry of Education), Ocean University of China, Qingdao 266100, China

12



13 14

College of Environmental Science and Engineering, Key Laboratory of Marine Environmental

School of Chemistry and Environment, South China normal University, Guangzhou 510006, China



Stockbridge School of Agriculture, University of Massachusetts, Amherst, MA 01003, USA

§Equal

contribution

15 16 17

* Email: [email protected] (Dr. Zhang Lin); phone: 86-20-39380503; fax: 86-20-

18

39380508

19

1

ACS Paragon Plus Environment

Environmental Science & Technology

20 21

TOC

22

2

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36

Environmental Science & Technology

23

ABSTRACT

24

p-Arsanilic acid (p-ASA) is an emerging organoarsenic pollutant comprising both

25

inorganic and organic moieties. For the efficient removal of p-ASA, adsorbents with

26

high adsorption affinity are urgently needed. Herein, amine modified UiO-67 (UiO-67-

27

NH2) metal-organic frameworks (MOFs) were synthesized, and their adsorption

28

affinities towards p-ASA were 2 times higher than that of the pristine UiO-67. Extended

29

X-ray absorption fine structure (EXAFS), X-ray photoelectron spectroscopy (XPS) and

30

density functional theory (DFT) calculation results revealed the adsorption through a

31

combination of As-O-Zr coordination, hydrogen bonding, and π-π stacking, among

32

which As-O-Zr coordination was the dominant force. Amine groups played a

33

significant role in enhancing the adsorption affinity through strengthening the As-O-Zr

34

coordination and π-π stacking, as well as forming new adsorption sites via hydrogen

35

bonding. UiO-67-NH2s could remove p-ASA at low concentrations (< 5 mg L-1) in

36

simulated natural and waste waters to an arsenic level lower than that in the drinking

37

water standard of World Health Organization (WHO) and the surface water standard of

38

China, respectively. This work provided an emerging and promising method to increase

39

the adsorption affinity of MOFs towards pollutants containing both organic and

40

inorganic moieties, via modifying functional groups based on the pollutant structure to

41

achieve synergistic adsorption effect.

42

3

ACS Paragon Plus Environment

Environmental Science & Technology

43

INTRODUCTION

44

As a typical phenylarsonic acid compound, p-arsanilic acid (p-ASA) is an emerging

45

micropollutant, since its concentrations in water and soil were reported to be ranged

46

from 0.5 to 5000 μg L-1 and 0.2 to 1000 μg kg-1, respectively.1-3 p-ASA is widely used

47

as animal feed additives, due to its broad antimicrobial and nontoxic properties.4,5

48

However, after excreted in manure, the water-soluble p-ASA would be degraded into

49

high-toxic inorganic arsenic species (such as arsenate) in 30 days via the biotic and

50

abiotic routes, thus leading to severe arsenic pollution in soil and groundwater.6,7 For

51

example, in the Pearl River Delta of southern China, the concentration of p-ASA was

52

detected up to 771 μg kg-1 in the surface soils surrounding swine farms, and the arsenic

53

content was found to be far beyond the local background level.2 Therefore, the removal

54

of p-ASA is crucial for maintaining the quality and safety of water and soil.

55

To date, studies on p-ASA removal mainly focused on two types of techniques

56

including degradation and adsorption.8,9 Comparing with the former technique,

57

adsorption exhibited the advantages of higher efficiency, simpler operation, and lower

58

risk of releasing inorganic arsenic into the environment. Several types of adsorbents

59

have been studied for p-ASA adsorption, including ferric and manganese binary

60

oxide,10 iron humate,11 and metal-organic frameworks (MOFs).12 Among these

61

adsorbents, MOFs are attracting more attentions because of their high porosity,

62

controllable structure, and high adsorption capacity to p-ASA (up to 791 mg g-1).13

63

However, the adsorption affinity of MOFs to p-ASA is relatively weak, since the

64

adsorption is mainly depended on the electrostatic attraction and pore filling. Therefore, 4

ACS Paragon Plus Environment

Page 4 of 36

Page 5 of 36

Environmental Science & Technology

65

to achieve efficient and quick removal of p-ASA, especially at environmentally relevant

66

concentrations (e.g., below 5 mg L-1), specific adsorption sites in MOF structures

67

towards p-ASA are required.

68

Our previous work revealed that roxarsone (ROX) could be efficiently adsorbed by

69

Fe3O4-graphene nanocomposites through coordination, hydrogen bonding, and π-π

70

stacking.14 Considering the similar structure of p-ASA with ROX, new strategies are

71

expected to increase the adsorption affinity through the mentioned interactions. Such

72

strategies should point to utilizing MOFs which contain metal clusters for the As-O-M

73

(M stands for the metal clusters in MOFs) coordination, organic linkers with benzene

74

rings for π-π stacking, and functional groups for forming hydrogen bonds with p-ASA.

75

UiO-67 is a Zr-based MOF which comprises Zr6O4(OH)4 clusters as 12-connected

76

nodes and linear 4,4’-biphenyldicarboxylic acid (BPDC) ligands.15 UiO-67 showed

77

high application potential in practical water treatments due to its simple synthesis

78

method and superior chemical stability.16 It consists of substantial octahedral (16 Å)

79

and tetrahedral (12 Å) cages,17,18 and consequently can expose abundant active sites for

80

p-ASA adsorption. Considering the -NH2 and -OH groups in p-ASA molecules, the

81

incorporation of -NH2 groups to the BPDC ligands of UiO-67 (denote as UiO-67-NH2)

82

is expected to enhance the adsorption affinity through hydrogen bonding.19-21 Moreover,

83

other interactions are also expected between UiO-67-NH2 and p-ASA. It was reported

84

that the missing-linkers in UiO-66 could induce defects in the Zr nodes, which could

85

provide more Zr-OH groups and greatly enhance the capture of ROX via the stronger

86

As-O-Zr coordination.22,23 Similarly, this effect is also predicted between the Zr nodes 5

ACS Paragon Plus Environment

Environmental Science & Technology

87

with defects in UiO-67-NH2 and p-ASA. In addition, the BPDC ligands in UiO-67-NH2

88

are also expected to bind p-ASA through π-π stacking.24,25

89

Therefore, we expect the defective UiO-67-NH2 to be a powerful and advanced

90

material for the efficient adsorption of p-ASA from aqueous media. Herein, we aimed

91

to: (1) determine the adsorption capacity, affinity and rate of defective UiO-67-NH2

92

towards p-ASA in comparison with the pristine UiO-67, to confirm the enhanced

93

adsorption of UiO-67-NH2; (2) analyze the contribution of As-O-Zr coordination, π-π

94

stacking and hydrogen bonding to the adsorption via extended X-ray absorption fine

95

structure (EXAFS), x-ray photoelectron spectroscopy (XPS), and density functional

96

theory (DFT) calculations; and (3) investigate the mechanism of amine modification on

97

the enhancement of adsorption affinity. This study provides useful information on the

98

design of novel and efficient materials for p-ASA removing via understanding the

99

adsorption mechanism, and offers a new alternative to increase the adsorption affinity

100

via the synergistic effect between the introduced functional groups.

101 102

EXPERIMENTAL SECTION

103

Synthesis and characterization of UiO-67 and UiO-67-NH2s

104

The defective UiO-67 and its -NH2 modified derivatives were prepared using

105

biphenyl-4,4’-dicarboxylic acid (BPDC) as ligand by following previously reported

106

methods.22,26 The missing-linker-induced defects in the MOFs were produced by adding

107

concentrated HCl during preparation. Two types of UiO-67-NH2s with different amine

108

contents were obtained. The as-prepared UiO-67-NH2 that theoretically had one amine 6

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36

Environmental Science & Technology

109

group per BPDC ligand was abbreviated as UiO-67-NH2(1), while the other one that

110

had two -NH2 groups per BPDC ligand as UiO-67-NH2(2).

111

The crystal structure, surface morphology, specific surface area and surface charge

112

of UiO-67 and UiO-67-NH2s were analyzed using X-ray diffraction (XRD), scanning

113

electron microscope (SEM), Brunauer-Emmett-Teller (BET) N2 adsorption/desorption

114

method and zeta potential analyzer, respectively. Zeta potentials of the materials at pHs

115

3-10 were determined using a NanoBrook Omni zeta potential analyzer (Brookhaven,

116

USA). The concentration of the tested MOFs was 1 g L-1, and the pH was adjusted by

117

HNO3 or KOH solution. Details of the characterization are provided in Section S1.

118

Adsorption experiments

119

For adsorption kinetic experiments, 0.045 g adsorbents were added to 300 mL p-ASA

120

(Aladdin Bio-Chem Technology, China) solution at a concentration of 50 mg L-1, and

121

shaken at pH 4.0 ± 0.1 and 25 C with a speed of 200 rpm for 1 min to 24 h. The

122

adsorption kinetics curves were fitted with the pseudo-second-order non-linear kinetic

123

model (SI Section S1). Adsorption isotherm studies were conducted with the p-ASA

124

concentrations from 1 to 100 mg L-1, while the adsorbent concentration was kept at 0.15

125

g L-1 in 30 mL solution (pH 4.0 ± 0.1). The mixtures were shaken under 25 C with a

126

speed of 200 rpm for 24 h. After centrifugation separation, the residual p-ASA

127

concentrations in the supernatants were measured by a UV-vis spectrometer (UV-2600,

128

Shimadzu) at 252 nm. The adsorption isotherms were fitted with Langmuir and

129

Freundlich models (SI Section S1). The effect of pH on adsorption was investigated by

130

conducting adsorption experiments within the pH range of 3.0-10.0 at the p-ASA 7

ACS Paragon Plus Environment

Environmental Science & Technology

131

concentration of 50 mg L-1 and the adsorbent dosages of 0.15 g L-1.

132

To detect the reusability of UiO-67 and UiO-67-NH2s, each adsorbent (5 mg) was

133

firstly added to 30 mL p-ASA solution (pH 4.0 ± 0.1). After shaking for 24 h (25 C,

134

200 rpm), the obtained adsorbents after adsorption were moved to 30 mL desorption

135

solution containing 0.5 mol L-1 NaOH/ethanol, and shaking with a speed of 200 rpm at

136

25 C for 1 h. The regenerated adsorbents were dried in a vacuum oven at 110 °C for

137

12 h before the next adsorption cycle. The removal of p-ASA by UiO-67-NH2s in

138

simulated natural and waste waters was studied in the presence of dissolved organic

139

matter (DOM, 0-16 mg C L-1) and in swine manure lixivium (added p-ASA of 1-5 mg

140

L-1). Experimental details were displayed in Section S1.

141

EXAFS and XPS analyses

142

EXAFS and XPS were used to investigate the interaction between p-ASA and UiO-

143

67s.27 The samples on p-ASA (50 mg L-1) adsorbed UiO-67 and UiO-67-NH2s (0.15 g

144

L-1) were prepared at pH 4.0, in a manner similar to the adsorption experiments. The K-

145

edge EXAFS spectra of As in p-ASA-loaded samples were acquired at the beamline

146

4W1B of the Beijing Synchrotron Radiation Facility (BSRF). All EXAFS data were

147

collected in the fluorescence mode at room temperature, using a Lytle detector equipped

148

with Soller slits and Ge filter for screening scattering and fluorescence background. The

149

maximum absorption edge As was set to 11875 eV. The EXAFS data analysis was

150

performed using the ATHENA and ARTEMIS interfaces to the IFEFFIT version 1.2.11

151

program package.28 Details of the EXAFS data collection and analysis are provided in

152

Section S1. 8

ACS Paragon Plus Environment

Page 8 of 36

Page 9 of 36

Environmental Science & Technology

153

XPS spectra were analyzed using an Axis Ultra DLD instrument (Kratos Analytical,

154

U.K.) with an A1 Kα X-ray source. Spectra were recorded at a pass energy of 160 eV

155

for survey scans and 40 eV for high-resolution scans. High-resolution scans were

156

carried out in an energy range of 295-280 eV for C 1s, and 405-395 eV for N 1s XPS

157

spectra.

158

Computational calculation

159

DFT calculations were performed using the Vienna ab initio simulation package

160

(VASP). The Perdew-Burke-Ernzerhof (PBE) functional, and the potential projector

161

augmented wave (PAW) pseudo-potentials were used for all the calculations. The plane-

162

wave cutoff energy was set to 500 eV, which has been previously tested and shown to

163

be appropriate for UIO-67.29 Scalar relativistic effects were incorporated into the

164

effective core potentials via explicit mass-velocity and Darwin corrections. Detailed

165

calculation methods are provided in Section S1.

166 167

RESULTS AND DISCUSSION

168

Characterization of the as-prepared MOFs

169

The as-synthesized UiO-67 and UiO-67-NH2s exhibited typical characteristic peaks

170

of Fm-3m symmetric space groups in the XRD patterns (Figure S1), consistent with the

171

reported UiO-67 topology.30 The identical XRD peaks of UiO-67-NH2s with the

172

pristine UiO-67 implied that introducing -NH2 groups did not change the framework

173

structure. FTIR spectra of UiO-67-NH2s showed C-N and N-H stretching vibrations at

174

1257 and 3300-3500 cm-1, respectively (Figure S2). The mole ratios of -NH2:Zr 9

ACS Paragon Plus Environment

Environmental Science & Technology

175

calculated from XPS data were 0.88 and 1.80 for UiO-67-NH2(1) and UiO-67-NH2(2),

176

respectively (Table S1), which were similar to their theoretical values (1.0 for UiO-67-

177

NH2(1) and 2.0 for UiO-67-NH2(2)). These results confirmed the successful amine

178

modification of UiO-67. The structural illustrations of UiO-67, UiO-67-NH2(1) and

179

UiO-67-NH2(2) are shown in Figure 1. The BET surface areas of UiO-67, UiO-67-

180

NH2(1) and UiO-67-NH2(2) were 1871, 750, and 465 m2 g-1, respectively (Figure S3,

181

and Table S1). The lower surface areas of UiO-67-NH2s compared with the pristine

182

UiO-67 were probably due to the partial blocking of the pores by -NH2 groups on the

183

surface (Figure 1a-1c).

184

The as-prepared MOFs are thermally stable up to 320 °C as detected by the TG-MS

185

(Figure S4) and temperature-dependent XRD (Figure S5). Therefore, based on the TGA

186

data on the weight loss at 320-500 °C (Table S2, Figure S6), the numbers of the missing-

187

linker induced defects for each unit of UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2)

188

were calculated as 4, 3.2 and 3.2, respectively. These defects would provide more

189

binding centers towards p-ASA through the external Zr-OH groups,31 thus facilitating

190

the p-ASA adsorption.

191

Adsorption kinetics

192

Adsorption kinetics of p-ASA by UiO-67 and UiO-67-NH2s was investigated and the

193

results are shown in Figure 2a. All the kinetic curves were fitted using the pseudo-

194

second order model, which was commonly used to describe adsorption kinetics in

195

which chemisorption controlled the adsorption rate and the number of active sites

196

determined the adsorption capacity.32 The obtained kinetic parameters are shown in 10

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36

Environmental Science & Technology

197

Table S3. Good fits of the tested MOFs with the pseudo-second order model (R2 > 0.99)

198

implied the adsorption of p-ASA on the tested MOFs was controlled by chemical

199

interactions. The increased adsorption rates (k2) of UiO-67-NH2s relative to the pristine

200

UiO-67 implied the enhancement of chemisorption between p-ASA and the amine

201

modified MOFs. Moreover, in comparison with other porous MOFs (such as meso-ZIF-

202

8 in ref.13), substantially enhanced adsorption rates were observed for UiO-67 and UiO-

203

67-NH2s, which was possibly due to the stronger chemisorption between UiO-67s and

204

p-ASA molecules.

205

Adsorption isotherms

206

Adsorption isotherms of UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2) towards p-

207

ASA are shown in Figure 2b. All isotherms were fitted well with Langmuir and

208

Freundlich33,34 models (Table S4), but the R2 values of Langmuir model were higher.

209

Therefore, the following discussion was mainly based on Langmuir-fitting results of all

210

the MOFs. Maximum adsorption capacities (Qm) calculated from Langmuir model

211

followed an order of UiO-67 > UiO-67-NH2(1) > UiO-67-NH2(2) (Figure 2b, Table S4).

212

However, this order changed to an opposite sequence after normalizing by the surface

213

area (Qm/Asurf in Figure 2c), indicating higher adsorption capacities of UiO-67-NH2s

214

than the pristine UiO-67 deducting the effect of surface area. This observation could be

215

explained by the steric hindrance of the introduced -NH2 groups situating in the aperture

216

of the nanopores (Figure 1 and Table S1). The entry of p-ASA (with a molecule

217

diameter of 6.625 Å13) to the internal part of UiO-67-NH2s would be obstructed because

218

of their narrower apertures than the pristine UiO-67. 11

ACS Paragon Plus Environment

Environmental Science & Technology

219

It was reported that materials with a small Qm and large affinity coefficient (KL) could

220

have a larger value of qe when compared with materials with a large Qm and small KL,

221

at low pollutant concentrations.35 Therefore, considering p-ASA as a micropollutant,

222

the KL value was more significant than Qm. As shown in Table S4, KL values of UiO-

223

67-NH2s were much higher than that of the pristine UiO-67, and increased with

224

increasing -NH2 content. In order to further evaluate the adsorption affinity at different

225

p-ASA concentrations, single point sorption coefficients (K)36 based on the Langmuir

226

fitting results were calculated when Ce = 0.03 (K0.03) and 0.3 (K0.3) mg L-1. As shown

227

in Table S4, both of K0.03 and K0.3 followed the order of UiO-67-NH2(2) > UiO-67-

228

NH2(1) > UiO-67, indicating that UiO-67-NH2s had higher affinities towards p-ASA.

229

The higher affinities made UiO-67-NH2s having higher qe compared with the pristine

230

UiO-67 at low concentrations of p-ASA, as shown in the insert of Figure 2b. This is

231

very important for the practical application of UiO-67-NH2s, considering that the

232

concentrations of p-ASA in environment are usually very low.

233

The comparison of adsorption capacity and affinity of the present UiO-67s with other

234

adsorbents in the published literature is listed in Table S5. The Qm value of UiO-67

235

(454.54 mg g-1) was higher than most of the reported absorbents. Despite that

236

mesoporous ZIF-8 had the highest Qm (791 mg g-1) and Qm/Asurf (0.70 mg m-2), UiO-67-

237

NH2(2) was more superior for p-ASA capture in practical applications due to its much

238

higher adsorption capacity and affinity toward p-ASA (qe0.03 = 4.50 mg g-1 and K0.03 =

239

150 L g-1) than ZIF-8 (qe0.03 = 0.76 mg g-1 and K0.03 = 25.3 L g-1) at low concentrations.

240

Reusability 12

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36

Environmental Science & Technology

241

The regeneration efficiency of UiO-67s is shown in Figure S7. Despite the adsorption

242

capacities slightly decreased after the first recycle, UiO-67 and UiO-67-NH2(2) still

243

exhibited considerable adsorption capacities after four cycles regeneration (over 85%

244

of the original capacity). FTIR analysis showed the -NH and C-N vibration bands still

245

present in UiO-67-NH2(2) after the first adsorption cycle (Figure S8), suggesting that

246

the -NH2 moieties remained stable during the repeated adsorption process. Meanwhile,

247

XRD results showed that the crystal structure of UiO-67 and UiO-67-NH2(2) did not

248

change after regeneration (Figure S9). These results revealed that the UiO-67 and UiO-

249

67-NH2(2) could be regenerated and reused for multiple times, thus confirming the

250

practical value and application of UiO-67s as the adsorbents of p-ASA.

251

Mechanism of p-ASA adsorption on UiO-67-NH2s

252

To get insight into the adsorption mechanism of UiO-67-NH2s towards p-ASA,

253

electrostatic interaction, As-O-Zr coordination, hydrogen bonding, and π-π interaction

254

were taken into consideration.37 The effects of pH on the adsorption of p-ASA were

255

studied to investigate the contribution of electrostatic interaction. Figure S10a showed

256

that UiO-67 and UiO-67-NH2s exhibited the highest adsorption capacity at pH 4.0, and

257

then sharply decreased with the increasing pHs. The isoelectric points of the three UiO-

258

67s were determined to be between pH 4 and 5, and thus their zeta potentials were quite

259

low at pH 4.0 (Figure S10b). Moreover, the aqueous dissociation constants (pKa) of p-

260

ASA molecules were 1.91, 4.13 and 9.19, respectively (Figure S11). Therefore, the

261

arsenic group in p-ASA is electroneutral at pH 4.0.38 These results indicated that the

262

electrostatic attraction was not the dominant force for the p-ASA adsorption at pH 4.0. 13

ACS Paragon Plus Environment

Environmental Science & Technology

263

It has been well-documented that Zr-based materials could efficiently capture

264

arsenate groups due to the bidentate binuclear and monodentate mononuclear

265

complexes forming between the Zr-OH groups and As.22,39 For the as-prepared UiO-67

266

and UiO-67-NH2s, these Zr-OH groups mainly originated from the defects present in

267

Zr-O clusters.23,40 To further reveal the role of As-O-Zr coordination in p-ASA

268

adsorption, x-ray adsorption spectroscopy of As was determined. The As-O and As-Zr

269

shells could be obviously isolated in the k3-weighted As K-edge EXAFS spectra and

270

the corresponding Fourier-transforms (Figure 3), indicating the formation of As-O-Zr

271

inner-sphere complexes after adsorption. In all the samples, the As-O first-neighbor

272

contributions were fit with 3.6-3.9 oxygen atoms at 1.69 Å (Table S6), which agreed

273

with the DFT calculation in the previous study.39

274

The second-neighbor contributions were fitted with the As-Zr atom distances of 3.44,

275

3.41, and 3.40 Å for UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2), respectively (Table

276

S6). The As-Zr distances were consistent with that in the bidentate binuclear complex

277

obtained from the DFT calculation (see details in the DFT Calculation section), which

278

was also observed in other Zr-based MOFs.41 Moreover, the As-Zr distances in UiO-

279

67-NH2(1) and UiO-67-NH2(2) were shorter than that in UiO-67, indicating that the

280

amine modification of UiO-67 shortened the As-Zr atom distance, and thus formed

281

stronger inner-sphere complexes. The enhanced As-O-Zr coordination was one of the

282

explanations for the higher adsorption affinity of UiO-67-NH2s towards p-ASA than

283

the pristine UiO-67.

284

The higher coordination number (CN) of the As-Zr contribution in UiO-67-NH2s 14

ACS Paragon Plus Environment

Page 14 of 36

Page 15 of 36

Environmental Science & Technology

285

than the pristine UiO-67 (Table S6) also indicated stronger As-O-Zr coordination. Such

286

enhancement could be ascribed to the synergistic effect of other interactions, such as π-

287

π stacking and hydrogen bonding. Moreover, according to the Langmuir fitting results,

288

the maximum molecule uptake of p-ASA by UiO-67 was found to be 5.3:1 (p-ASA:Zr6

289

in Figure 4a), which exceeded the stoichiometric p-ASA:Zr6 value calculated by As-O-

290

Zr coordination (4:1 of p-ASA:Zr6 based on the bidentate binuclear complex forming

291

between p-ASA and the defective UiO-67 unit). Considering the absence of -NH2

292

groups in UiO-67 ligands, the hydrogen bonds between the pristine UiO-67 and p-ASA

293

were perceived to be relatively weak. Thus, the higher p-ASA uptake than the

294

stoichiometric value was possibly due to the π-π stacking. Further indication was

295

conducted with the high resolution C 1s XPS spectra (Figure 4b and 4c). For UiO-67s

296

before p-ASA adsorption, three peaks were observed at 284.6 eV, 285.9 eV and 288.5

297

eV, which belonged to the sp2 C=C, C-N, and O-C=O groups in the BPDC linkers.42,43

298

The characteristic peak of π-π* component appeared at 291.2 eV44 after adsorption,

299

suggesting the contribution of π-π stacking for p-ASA adsorption. Meanwhile, the π-π*

300

proportions of UiO-67-NH2(1) and UiO-67-NH2(2) were 2.11% and 7.78%,

301

respectively, which were both higher than that of the pristine UiO-67, suggesting the

302

enhancement of π-π interaction after the -NH2 modification.

303

Considering the higher adsorption affinity of UiO-67-NH2s than the pristine UiO-67,

304

additional mechanism other than As-O-Zr coordination and π-π interaction should also

305

contribute to the adsorption by UiO-67-NH2s. As mentioned above, the good fitting of

306

the adsorption kinetic to the pseudo-second order model suggested that the adsorption 15

ACS Paragon Plus Environment

Environmental Science & Technology

307

capacities of UiO-67s were determined by the number of active sites on the

308

adsorbents.32 Since the surface area of UiO-67-NH2(2) was only 60% of UiO-67-NH2(1)

309

(Table S1), the active sites in UiO-67-NH2(2) could also conclude to be ~60% of those

310

in UiO-67-NH2(1). Therefore, the adsorption capacity of UiO-67-NH2(2) should be

311

about 60% of UiO-67-NH2(1) theoretically. However, the Langmuir-fitted adsorption

312

capacity and the molecular uptake of p-ASA:Zr6 by UiO-67-NH2(2) were calculated to

313

be 178 mg g-1 and 2.3:1, respectively, which were much higher than 60% of the p-ASA

314

amount adsorbed by UiO-67-NH2(1) (Table S4 and Figure 4a). This result suggested

315

that the additional -NH2 groups provided more active sites, thus enhancing the

316

adsorption capacity and affinity of UiO-67-NH2(2). Considering the previous

317

discussion, hydrogen bonds between UiO-67-NH2s and p-ASA were expected to be the

318

key role for the enhancement. XPS N 1s results provided strong evidence for the

319

hydrogen bond formation (Figure 4d and 4e). Before adsorption, the symmetrical peaks

320

locating at 399.4 eV were attributed to the free -NH245 in the BPDC linkers of UiO-67-

321

NH2s (Figure 4d). No signals of H-bonded -NH2 were observed in the spectra. After

322

adsorption, although no -NH2 groups existing on the pristine UiO-67 structure, the peak

323

for free -NH2 was still observed (Figure S12), which belonged to the adsorbed p-ASA.

324

Moreover, a new peak assigned to H-bonded -NH2 at 400.2 eV45 was observed in the N

325

1s spectra of UiO-67-NH2s after adsorption (Figurer 4e), and the peak ratio increased

326

with the amine content in UiO-67-NH2s. This result indicated that these hydrogen bonds

327

increased the sites for p-ASA adsorption on UiO-67-NH2s.

328 16

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36

Environmental Science & Technology

329

DFT calculations

330

To further elucidate the roles of -NH2 groups in UiO-67-NH2s for p-ASA adsorption,

331

the possible interaction scenarios of UiO-67 and UiO-67-NH2(2) with p-ASA were

332

further investigated at the molecular level through DFT calculation. For UiO-67, it was

333

obvious that the As-O-Zr coordination had the highest interaction energy (-181.5 kJ

334

mol-1 in Figure S13a) and was considered to be the most favorable adsorption force.

335

The optimal configurations of the Zr nodes and p-ASA molecules suggested the

336

formation of bidentate binuclear As-O-Zr complexes after adsorption (Figure S13a),

337

which was consistent with the aforementioned EXAFS analysis. In addition, the

338

relatively high adsorption energy (-40.2 kJ mol-1) of the π-π interaction between p-ASA

339

and UiO-67 (face-centered distance, 3.8 Å, Figure S13b) indicated that the π-π stacking

340

also played an important role in the adsorption.

341

After amine modification, the optimal configurations between UiO-67-NH2(2) and

342

p-ASA are shown in Figure 5. The As-O-Zr bidentate binuclear complex, π-π stacking

343

and hydrogen bonds were all present in the UiO-67-NH2(2) samples after adsorption.

344

Moreover, it was obvious that the binding energies of As-O-Zr coordination and π-π

345

interaction in UiO-67-NH2(2) were both higher than those in the pristine UiO-67, while

346

the As-Zr atom distance and the benzene rings face-centered distance were shorter

347

(Figure 5, S13). These results indicated that As-O-Zr coordination and π-π stacking

348

were enhanced after -NH2 modification. Moreover, -NH2 groups in UiO-67-NH2(2) also

349

provided extra adsorption sites for p-ASA through hydrogen bonds with an optimal

350

configuration of NH…O and a bonding energy of -52.4 kJ mol-1 (Figure 5c), thus 17

ACS Paragon Plus Environment

Environmental Science & Technology

351

further increasing the adsorption affinity. These results were in agreement with the

352

EXAFS and XPS analyses. The interaction energies followed the order: UiO-67-NH2(2)

353

and p-ASA > water molecule and p-ASA > p-ASA and p-ASA according to DFT

354

calculation (Figure 5, S13), suggesting that p-ASA molecules were monolayerly

355

adsorbed on UiO-67-NH2(2) because free p-ASA molecules cannot be further adsorbed

356

by the bound p-ASA on UiO-67-NH2(2). Overall, p-ASA would form a monolayer

357

adsorption on the surface of UiO-67-NH2(2), under the combined interactions of As-O-

358

Zr coordination, π-π stacking and hydrogen bonding (Figure 6). It should be noted that

359

amine groups on UiO-67-NH2(2) played a significant role in enhancing the adsorption

360

affinity through forming new adsorption sites via hydrogen bonding, as well as

361

strengthening the As-O-Zr coordination and π-π stacking.

362 363

Adsorption behavior of amine modified UiO-67 in simulated natural and waste waters

364

To test if UiO-67-NH2s could efficiently remove p-ASA for practical application,

365

DOM containing water and swine manure lixivium were used to simulated natural and

366

waste water, respectively. Adsorption results indicated that at a p-ASA concentration of

367

0.5 mg L-1, the removal rate of UiO-67-NH2(2) was slightly decreased from 99.6% to

368

98.7% when the DOM concentration increased from 0 to 16 mg C L-1, which were

369

higher than those of UiO-67 (Figure 7a). When the DOM concentration was 16 mg C

370

L-1, the residual arsenic concentrations after adsorbed by UiO-67-NH2(1) and UiO-67-

371

NH2(2) were 4.15 and 2.26 μg L-1, respectively (Figure 7b), which were lower than the

372

arsenic standard in drinking water of WHO (10 μg L-1). The adsorption results of UiO18

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36

Environmental Science & Technology

373

67-NH2s in swine manure lixivium at different p-ASA concentrations are shown in

374

Figure 7c. It can be seen that when the initial concentration of p-ASA was 5 mg L-1, a

375

dosage of 0.15 g L-1 UiO-67-NH2(2) could decrease the arsenic concentration to 28.1

376

μg L-1, which satisfied the arsenic standard in surface water of China (50 μg L-1,

377

GB3838-2002). To reduce the arsenic concentration to the same level (below 50 μg L-

378

1

379

respectively, which were 26.7 and 44.8 times higher than that of UiO-67-NH2(2) (Table

380

S7). Moreover, since UiO-67 has already been studied as a water-treatment adsorbent

381

for a period of time,16 its modification is supposed to be more simple and commercially

382

available, compared with synthesizing a new MOF. These results suggested that the

383

amine modified UiO-67 was a promising candidate for removing p-ASA from natural

384

and waste waters.

), the needed dosages of graphene and activated carbon were 4 g L-1 and 6.7 g L-1,

385 386

ENVIRONMENTAL IMPLICATIONS

387

The defective UiO-67-NH2 designed based on the structural features of p-ASA

388

showed high adsorption affinity towards p-ASA. The study of adsorption mechanism

389

indicated that the contribution of different adsorption forces followed an order of As-

390

O-Zr coordination > π-π stacking > H-bonding. The -NH2 groups in UiO-67-NH2

391

played an important role for the efficient adsorption, including enhancing the As-O-Zr

392

coordination and π-π stacking, as well as providing new adsorption sites for p-ASA

393

adsorption through H-bonding. These synergistic interactions resulted in strong affinity

394

of UiO-67-NH2 towards p-ASA, which were beneficial for the removal of p-ASA in 19

ACS Paragon Plus Environment

Environmental Science & Technology

395

simulated natural and waste waters. This study revealed great potential of -NH2

396

modified UiO-67s for the removal of p-ASA in practical applications. Moreover, this

397

work also provided a new route for increasing the adsorption affinity of adsorbents

398

towards targeted pollutants through engineering different functional groups to achieve

399

synergistic effects.

400 401

ASSOCIATED CONTENT

402

Supporting Information

403

Thirteen figures, seven tables and experimental details are presented in Supporting

404

Information section. The material is available free of charge via the Internet at

405

http://pubs.acs.org.

406

AUTHOR INFORMATION

407

Corresponding Authors

408

* Email: [email protected] (Dr. Zhang Lin); phone: 86-20-39380503; fax: 86-20-

409

39380508

410

Notes

411

The authors declare no competing financial interest.

412 413

ACKNOWLEDGMENTS

414

This work was supported by the National Natural Science Foundation of China (grant

415

no. 21477129), the Guangdong Innovative and Entrepreneurial Research Team

416

Program (No. 2016ZT06N569), the China Postdoctoral Science Foundation (no. 20

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36

Environmental Science & Technology

417

2016M600654), and the Fundamental Research Funds for the Central Universities (no.

418

2017PY009 and 2017BQ054). The authors thank the beamline 4W1B (Beijing

419

Synchrotron Radiation Facility) for providing the beam time.

420 421

REFERENCES

422

(1) Liu, X.; Zhang, W.; Hu, Y.; Hu, E.; Xie, X.; Wang, L.; Cheng, H. Arsenic Pollution

423

of Agricultural Soils by Concentrated Animal Feeding Operations (CAFOS).

424

Chemosphere 2015, 119, 273-281.

425

(2) Liu, X.; Zhang, W.; Hu, Y.; Cheng, H. Extraction and Detection of Organoarsenic

426

Feed Additives and Common Arsenic Species in Environmental Matrices by HPLC-

427

ICP-MS. Microchem. J. 2013, 108, 38-45.

428

(3) Hu, Y.; Zhang, W.; Cheng, H.; Tao, S. Public Health Risk of Arsenic Species in

429

Chicken Tissues From Live Poultry Markets of Guangdong Province, China. Environ.

430

Sci. Technol. 2017, 51, (6), 3508-3517.

431

(4) Jones, F. T. A Broad View of Arsenic. Poultry Sci. 2007, 86, (1), 2-14.

432

(5) Depalma, S.; Cowen, S.; Hoang, T.; Al-Abadleh, H. A. Adsorption

433

Thermodynamics of p-Arsanilic Acid on Iron (Oxyhydr)Oxides: In-Situ ATR-FTIR

434

Studies. Environ. Sci. Technol. 2008, 42, (6), 1922-1927.

435

(6) Wang, L.; Cheng, H. Birnessite (-MnO2) Mediated Degradation of Organoarsenic

436

Feed Additive p-Arsanilic Acid. Environ. Sci. Technol. 2015, 49, (6), 3473-3481.

437

(7) Czaplicka, M.; Bratek, A.; Jaworek, K.; Bonarski, J.; Pawlak, S. Photo-Oxidation

438

of p-Arsanilic Acid in Acidic Solutions: Kinetics and the Identification of by-Products 21

ACS Paragon Plus Environment

Environmental Science & Technology

439

and Reaction Pathways. Chem. Eng. J. 2014, 243, 364-371.

440

(8) Liu, Y.; Hu, P.; Zheng, J.; Wu, M.; Jiang, B. Utilization of Spent Aluminum for p-

441

Arsanilic Acid Degradation and Arsenic Immobilization Mediated by Fe(II) Under

442

Aerobic Condition. Chem. Eng. J. 2016, 297, 45-54.

443

(9) Xie, X.; Hu, Y.; Cheng, H. Rapid Degradation of p-Arsanilic Acid with

444

Simultaneous Arsenic Removal from Aqueous Solution Using Fenton Process. Water

445

Res. 2016, 89, 59-67.

446

(10) Joshi, T. P.; Zhang, G.; Jefferson, W. A.; Perfilev, A. V.; Liu, R.; Liu, H.; Qu, J.

447

Adsorption of Aromatic Organoarsenic Compounds by Ferric and Manganese Binary

448

Oxide and Description of the Associated Mechanism. Chem. Eng. J. 2017, 309, 577-

449

587.

450

(11) Peng, Y.; Wei, W.; Zhou, H.; Ge, S.; Li, S.; Wang, G.; Zhang, Y. Iron Humate as

451

a Novel Adsorbent for p-Arsanilic Acid Removal From Aqueous Solution. J. Disper.

452

Sci. Technol. 2016, 37, (11), 1590-1598.

453

(12) Jun, J. W.; Tong, M.; Jung, B. K.; Hasan, Z.; Zhong, C.; Jhung, S. H. Effect of

454

Central Metal Ions of Analogous Metal-Organic Frameworks On Adsorption of

455

Organoarsenic Compounds from Water: Plausible Mechanism of Adsorption and Water

456

Purification. Chem.-Eur. J. 2015, 21, (1), 347-354.

457

(13) Jung, B. K.; Jun, J. W.; Hasan, Z.; Jhung, S. H. Adsorptive Removal of p-Arsanilic

458

Acid From Water Using Mesoporous Zeolitic Imidazolate Framework-8. Chem. Eng.

459

J. 2015, 267, 9-15.

460

(14) Tian, C.; Zhao, J.; Zhang, J.; Chu, S.; Dang, Z.; Lin, Z.; Xing, B. Enhanced 22

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36

Environmental Science & Technology

461

Removal of Roxarsone by Fe3O4@3D Graphene Nanocomposites: Synergistic

462

Adsorption and Mechanism. Environ. Sci. Nano 2017, 4, (11), 2134-2143.

463

(15) Salomon, W.; Roch-Marchal, C.; Mialane, P.; Rouschmeyer, P.; Serre, C.; Haouas,

464

M.; Taulelle, F.; Yang, S.; Ruhlmann, L.; Dolbecq, A. Immobilization of

465

Polyoxometalates in the Zr-Based Metal Organic Framework Uio-67. Chem. Commun.

466

2015, 51, (14), 2972-2975.

467

(16) Dias, E. M.; Petit, C. Towards the Use of Metal-Organic Frameworks for Water

468

Reuse: A Review of the Recent Advances in the Field of Organic Pollutants Removal

469

and Degradation and the Next Steps in the Field. J. Mater. Chem. A 2015, 3, (45),

470

22484-22506.

471

(17) Katz, M. J.; Brown, Z. J.; Colon, Y. J.; Siu, P. W.; Scheidt, K. A.; Snurr, R. Q.;

472

Hupp, J. T.; Farha, O. K. A Facile Synthesis of UiO-66, UiO-67 and their Derivatives.

473

Chem. Commun. 2013, 49, (82), 9449-9451.

474

(18) Vermoortele, F.; Bueken, B.; Le Bars, G.; Van de Voorde, B.; Vandichel, M.;

475

Houthoofd, K.; Vimont, A.; Daturi, M.; Waroquier, M.; Van Speybroeck, V.;

476

Kirschhock, C.; De Vos, D. E. Synthesis Modulation as a Tool to Increase the Catalytic

477

Activity of Metal-Organic Frameworks: The Unique Case of UiO-66(Zr). J. Am. Chem.

478

Soc. 2013, 135, (31), 11465-11468.

479

(19) Hasan, Z.; Tong, M.; Jung, B. K.; Ahmed, I.; Zhong, C.; Jhung, S. H. Adsorption

480

of Pyridine Over Amino-Functionalized Metal-Organic Frameworks: Attraction Via

481

Hydrogen Bonding Versus Base-Base Repulsion. J. of Phy. Chem. C 2014, 118, (36),

482

21049-21056. 23

ACS Paragon Plus Environment

Environmental Science & Technology

483

(20) Ahmed, I.; Jhung, S. H. Effective Adsorptive Removal of Indole From Model Fuel

484

Using a Metal-Organic Framework Functionalized with Amino Groups. J. Hazard.

485

Mater. 2015, 283, 544-550.

486

(21) Zhu, L.; Sheng, D.; Xu, C.; Dai, X.; Silver, M. A.; Li, J.; Li, P.; Wang, Y.; Wang,

487

Y.; Chen, L.; Xiao, C.; Chen, J.; Zhou, R.; Zhang, C.; Farha, O. K.; Chai, Z.; Albrecht-

488

Schmitt, T. E.; Wang, S. Identifying the Recognition Site for Selective Trapping of

489

99

490

Framework. J. Am. Chem. Soc. 2017, 139, (42), 14873-14876.

491

(22) Audu, C. O.; Nguyen, H. G. T.; Chang, C.; Katz, M. J.; Mao, L.; Farha, O. K.;

492

Hupp, J. T.; Nguyen, S. T. The Dual Capture of As(V) and As(III)by UiO-66 and

493

Analogues. Chem. Sci. 2016, 7, (10), 6492-6498.

494

(23) Li, B.; Zhu, X.; Hu, K.; Li, Y.; Feng, J.; Shi, J.; Gu, J. Defect Creation in Metal-

495

Organic Frameworks for Rapid and Controllable Decontamination of Roxarsone From

496

Aqueous Solution. J. Hazard. Mater. 2016, 302, 57-64.

497

(24) Seo, Y. S.; Khan, N. A.; Jhung, S. H. Adsorptive Removal of

498

Methylchlorophenoxypropionic Acid From Water with a Metal-Organic Framework.

499

Chem. Eng. J. 2015, 270, 22-27.

500

(25) Hasan, Z.; Khan, N. A.; Jhung, S. H. Adsorptive Removal of Diclofenac Sodium

501

from Water with Zr-Based Metal-Organic Frameworks. Chem. Eng. J. 2016, 284, 1406-

502

1413.

503

(26) Yang, J.; Dai, Y.; Zhu, X.; Wang, Z.; Li, Y.; Zhuang, Q.; Shi, J.; Gu, J. Metal-

504

Organic Frameworks with Inherent Recognition Sites for Selective Phosphate Sensing

TcO4- in a Hydrolytically Stable and Radiation Resistant Cationic Metal-Organic

24

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36

Environmental Science & Technology

505

through their Coordination-Induced Fluorescence Enhancement Effect. J. Mater. Chem.

506

A 2015, 3, (14), 7445-7452.

507

(27) Wu, C.; Tu, J.; Tian, C.; Geng, J.; Lin, Z.; Dang, Z. Defective Magnesium Ferrite

508

Nano-Platelets for the Adsorption of As(V): The Role of Surface Hydroxyl Groups.

509

Environ. Pollut. 2018, 235, 11-19.

510

(28) Ravel, B.; Newville, M. Athena, Artemis, Hephaestus: Data Analysis for X-Ray

511

Absorption Spectroscopy Using IFEFFIT. J. Synchrotron Radiat. 2005, 12, (4), 537-

512

541.

513

(29) Yang, L.; Ganz, E.; Svelle, S.; Tilset, M. Computational Exploration of Newly

514

Synthesized Zirconium Metal-Organic Frameworks UiO-66, -67, -68 and Analogues.

515

J. Mater. Chem. C 2014, 2, (34), 7111-7125.

516

(30) Ko, N.; Hong, J.; Sung, S.; Cordova, K. E.; Park, H. J.; Yang, J. K.; Kim, J. A

517

Significant Enhancement of Water Vapour Uptake at Low Pressure by Amine-

518

Functionalization of UiO-67. Dalton T. 2015, 44, (5), 2047-2051.

519

(31) Nguyen, H. G. T.; Schweitzer, N. M.; Chang, C.; Drake, T. L.; So, M. C.; Stair, P.

520

C.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. Vanadium-Node-Functionalized UiO-66:

521

A Thermally Stable MOF-Supported Catalyst for the Gas-Phase Oxidative

522

Dehydrogenation of Cyclohexene. Acs Catal. 2014, 4, (8), 2496-2500.

523

(32) Liu, K.; Zhang, S.; Hu, X.; Zhang, K.; Roy, A.; Yu, G. Understanding the

524

Adsorption of PFOA On MIL-101(Cr)-Based Anionic-Exchange Metal-Organic

525

Frameworks: Comparing DFT Calculations with Aqueous Sorption Experiments.

526

Environ. Sci. Technol. 2015, 49, (14), 8657-8665. 25

ACS Paragon Plus Environment

Environmental Science & Technology

527

(33) Alotaibi, K. M.; Shiels, L.; Lacaze, L.; Peshkur, T. A.; Anderson, P.; Machala, L.;

528

Critchley, K.; Patwardhan, S. V.; Gibson, L. T. Iron Supported On Bioinspired Green

529

Silica for Water Remediation. Chem. Sci. 2017, 8, (1), 567-576.

530

(34) Sheng, D.; Zhu, L.; Xu, C.; Xiao, C.; Wang, Y.; Wang, Y.; Chen, L.; Diwu, J.;

531

Chen, J.; Chai, Z.; Albrecht-Schmitt, T. E.; Wang, S. Efficient and Selective Uptake of

532

TcO4- by a Cationic Metal – Organic Framework Material with Open Ag+ Sites.

533

Environ. Sci. Technol. 2017, 51, (6), 3471-3479.

534

(35) Cao, Q.; Huang, F.; Zhuang, Z.; Lin, Z. A Study of the Potential Application of

535

Nano-Mg(OH)2 in Adsorbing Low Concentrations of Uranyl Tricarbonate from Water.

536

Nanoscale 2012, 4, (7), 2423-2430.

537

(36) Wang, Z.; Zhao, J.; Song, L.; Mashayekhi, H.; Chefetz, B.; Xing, B. Adsorption

538

and Desorption of Phenanthrene on Carbon Nanotubes in Simulated Gastrointestinal

539

Fluids. Environ. Sci. Technol. 2011, 45, (14), 6018-6024.

540

(37) Li, Y.; Yang, Z.; Wang, Y.; Bai, Z.; Zheng, T.; Dai, X.; Liu, S.; Gui, D.; Liu, W.;

541

Chen, M.; Chen, L.; Diwu, J.; Zhu, L.; Zhou, R.; Chai, Z.; Albrecht-Schmitt, T. E.;

542

Wang, S. A Mesoporous Cationic Thorium-Organic Framework that Rapidly Traps

543

Anionic Persistent Organic Pollutants. Nat. Commun. 2017, 8, (1), 1354.

544

(38) Nualláin, C. Ó.; Cinnéide, S. Ó. The Thermodynamic Ionization Constants of

545

Aromatic Arsonic Acids. J. Inorg. Nucl. Chem. 1973, 35, (8), 2871-2881.

546

(39) Schmidt, G. T.; Vlasova, N.; Zuzaan, D.; Kersten, M.; Daus, B. Adsorption

547

Mechanism of Arsenate by Zirconyl-Functionalized Activated Carbon. J. Colloid Interf.

548

Sci. 2008, 317, (1), 228-234. 26

ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36

Environmental Science & Technology

549

(40) Zhu, X.; Li, B.; Yang, J.; Li, Y.; Zhao, W.; Shi, J.; Gu, J. Effective Adsorption and

550

Enhanced Removal of Organophosphorus Pesticides from Aqueous Solution by Zr-

551

Based MOFs of UiO-67. Acs Appl. Mater. Inter. 2015, 7, (1), 223-231.

552

(41) Howarth, A. J.; Katz, M. J.; Wang, T. C.; Platero-Prats, A. E.; Chapman, K. W.;

553

Hupp, J. T.; Farha, O. K. High Efficiency Adsorption and Removal of Selenate and

554

Selenite from Water Using Metal-Organic Frameworks. J. Am. Chem. Soc. 2015, 23,

555

(123), 7488-7494.

556

(42) Liang, Q.; Zhang, M.; Zhang, Z.; Liu, C.; Xu, S.; Li, Z. Zinc Phthalocyanine

557

Coupled with Uio-66 (NH2) via a Facile Condensation Process for Enhanced Visible-

558

Light-Driven Photocatalysis. J. Alloy. Compd. 2017, 690, 123-130.

559

(43) Su, Y.; Zhang, Z.; Liu, H.; Wang, Y. Cd0.2Zn0.8S@UiO-66-NH2 Nanocomposites

560

as Efficient and Stable Visible-Light-Driven Photocatalyst for H2 Evolution and CO2

561

Reduction. Applied Catalysis B: Environmental 2017, 200, 448-457.

562

(44) Puziy, A. M.; Poddubnaya, O. I.; Socha, R. P.; Gurgul, J.; Wisniewski, M. XPS

563

and NMR Studies of Phosphoric Acid Activated Carbons. Carbon 2008, 46, (15), 2113-

564

2123.

565

(45) Zhang, F.; Srinivasan, M. P. Self-Assembled Molecular Films of Aminosilanes

566

and their Immobilization Capacities. Langmuir 2004, 20, (6), 2309-2314.

567

27

ACS Paragon Plus Environment

Environmental Science & Technology

568 569

FIGURES

570

571 572

Figure 1. Perspective views of (a) UiO-67, (b) UiO-67-NH2(1), and (c) UiO-67-NH2(2);

573

(d) Structure of the defective Zr6 node; (e) The BPDC linker for UiO-67 and its -NH2

574

modified forms for UiO-67-NH2(1) and UiO-67-NH2(2). In (a), (b), and (c), Zr6 nodes,

575

framework carbon and oxygen atoms are shown in cyan, white, and gray, respectively.

576

Modified -NH2 moieties are shown in red.

577

28

ACS Paragon Plus Environment

Page 28 of 36

Page 29 of 36

Environmental Science & Technology

578

579

580 581

Figure 2. (a) Adsorption kinetics, (b) adsorption isotherms, and (c) surface area-

582

normalized isotherms of p-ASA adsorbed by UiO-67, UiO-67-NH2(1) and UiO-67-

583

NH2(2).

29

ACS Paragon Plus Environment

Environmental Science & Technology

584

585

586

587 588

Figure 3. Arsenic K-edge data for p-ASA adsorbed UiO-67 and UiO-67-NH2 samples:

589

(a) filtered k3-weighted EXAFS data, (b) magnitude part of Fourier transformed R-

590

space, and (c) real part of Fourier transformed R-space. Experimental and calculated

591

curves are displayed as black solid lines and red open circles, respectively. Results of

592

the k3-weighted fitting are listed in Table S6. 30

ACS Paragon Plus Environment

Page 30 of 36

Page 31 of 36

Environmental Science & Technology

593

594

595

31

ACS Paragon Plus Environment

Environmental Science & Technology

596

597 598

Figure 4. The uptake of p-ASA by UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2), and

599

related XPS analysis after p-ASA adsorption. (a) The uptake of p-ASA (adsorbed p-

600

ASA / Zr6 node) by UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2); XPS C 1s and N 1s

601

spectra of UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2) before (b, d) and after (c, e) p-

602

ASA adsorption.

603

32

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36

Environmental Science & Technology

604

605

606 607

Figure 5. The optimal configurations obtained via DFT calculations of (a) As-O-Zr

608

coordination, (b) π-π stacking, (c) and H-bonding between UiO-67-NH2(2) and p- ASA

609

(Magenta for As, cyan for Zr, red for O, blue for N, gray for C, and white for H). The

610

interaction energy between UiO-67-NH2(2) and p-ASA was calculated to be -297.4 kJ

611

mol-1 (As-O-Zr coordination π-π stacking, and H-bonding).

612 33

ACS Paragon Plus Environment

Environmental Science & Technology

613 614 615 616

Figure 6. The adsorption mechanism of UiO-67-NH2 towards p-ASA

34

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36

Environmental Science & Technology

617

618

619 620

Figure 7. p-ASA adsorption by UiO-67, UiO-67-NH2(1) and UiO-67-NH2(2) in the

621

presence of DOM or in swine manure lixiviums. (a) Effect of DOM on the removal

622

rates of p-ASA; (b) the residual arsenic concentrations (C0 = 0.5 mg L-1) in the presence

623

of DOM; (c) the residual arsenic concentrations after adsorption in swine manure 35

ACS Paragon Plus Environment

Environmental Science & Technology

624

lixiviums with different initial concentrations of p-ASA. The adsorbent concentrations

625

were kept at 0.15 g L-1 in 30 mL solution. The mixtures were shaken at pH 4.0 ± 0.1

626

and 25 C with a speed of 200 rpm for 12 h.

36

ACS Paragon Plus Environment

Page 36 of 36