In situ growth of Metal-Organic Frameworks in Three-Dimensional

Feb 6, 2019 - In situ growth of Metal-Organic Frameworks in Three-Dimensional Aligned Lumen Arrays of Wood for Rapid and Highly Efficient Organic ...
0 downloads 0 Views 3MB Size
Subscriber access provided by TULANE UNIVERSITY

Remediation and Control Technologies

In situ growth of Metal-Organic Frameworks in ThreeDimensional Aligned Lumen Arrays of Wood for Rapid and Highly Efficient Organic Pollutants Removal Ruixue Guo, Xiaohui Cai, Hanwen Liu, Zi Yang, Yajie Meng, Fengjuan Chen, Yiju Li, and Baodui Wang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06564 • Publication Date (Web): 06 Feb 2019 Downloaded from http://pubs.acs.org on February 7, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 49

Environmental Science & Technology

In situ growth of Metal-Organic Frameworks in 3-Dimensional Aligned Lumen Arrays of Wood for Rapid and Highly Efficient Organic Pollutants Removal Ruixue Guo1†, Xiaohui Cai1†, Hanwen Liu1, Zi Yang1, Yajie Meng 1, Fengjuan Chen1*, Yiju Li2*, Baodui Wang1*

1

State Key Laboratory of Applied Organic Chemistry and Key Laboratory of Nonferrous Metal Chemistry and Resources Utilization of Gansu Province, Lanzhou University, Lanzhou, 730000, P. R. China 2 Department of Materials Science and Engineering, College of Engineering, Peking University, Beijing 100871, P. R. China *Email: [email protected]; [email protected]; [email protected] †These authors contributed equally to this work.

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 49

Table of Content

ACS Paragon Plus Environment

2

Page 3 of 49

Environmental Science & Technology

1

Abstract

2

Organic contaminants in water have become one of the most serious environmental problems

3

worldwide. Adsorption is one of the most promising approaches to remove organic pollutants

4

from water. However, the existing adsorbents have relatively low removal efficiency, complex

5

preparation process and high cost, which limit their practical applications. Here,we developed a

6

three-dimensional (3D) zirconium metal-organic frameworks (MOFs) encapsulated in a natural

7

wood membrane (UiO-66/wood membrane) for highly efficient organic pollutants removal from

8

water. UiO-66 MOFs were in situ growth in the 3D low-tortuosity wood lumens by a facile

9

solvothermal strategy. The resulting UiO-66/wood membrane contains the highly mesoporous

10

UiO-66 MOFs structure as well as many elongated and open lumens along the direction of the

11

wood growth. Such unique structural feature improves the mass transfer of organic pollutants

12

and increases the contact probability of organic contaminants with UiO-66 MOFs as the water

13

flows through the membrane, thereby improving the removal efficiency. Furthermore, the

14

integrated multilayer filter consisting of three pieces of UiO-66/wood membranes exhibits a high

15

removal efficiency (96.0%) for organic pollutants such as rhodamine 6G, propranolol, and

16

bisphenol A at the flux of 1.0×103 L∙m-2∙h-1. The adsorbed capacity of UiO-66/wood for Rh6G

17

(based on the content of UiO-66 MOFs) is calculated to be 690 mg∙g-1. We believe that such

18

low-cost and scalable production of UiO-66/wood membrane has broad applications for

19

wastewater treatment and other related pollutants removal.

20

Keywords: Metal organic frameworks, water treatment, organic pollutants removal, in situ

21

growth, lumen arrays

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 49

22

Introduction

23

Organic pollutants in water resources, such as organic dyes, pesticides, and pharmaceuticals,

24

have attracted great attention because of their potential negative impact on ecological

25

sustainability and human health.

26

methods for organic pollutants removal. So far, a variety of sorbents have been developed for

27

organic pollutants removal, including activated carbon, modified polymer,3 chelating resins,4

28

natural materials5 and so on. Among these adsorbents, activated carbon has been widely used as

29

a commercial material because of its high adsorption capacity and removal efficiency for some

30

organic pollutants.6 However, this material still suffers from several deficiencies, such as slow

31

treatment rate and poor removal efficiency for hydrophilic pollutants. Therefore, it is still

32

necessary to develop new materials with high efficiency, low cost and recyclability to rapidly

33

remove organic pollutants from aqueous solution. 5,7

34

MOFs with an abundant microporous structure, a large surface area, and high thermal stability 8,9

35

have triggered tremendous attention due to their various potential applications in molecular

36

separation,10 gas capture,11 chemical sensors,12 catalysis,13 and drug delivery.14,15 Recently,

37

water-stable MOFs have been used for the removal of organic contaminants including dyes,16

38

benzenes,17 phenols,18 hydrochloric acid, and acetic acid from aqueous solutions19,20 and

39

displayed great adsorption capacities in a laboratory scale. However, there are still facing issues

40

used as pollutant absorbents on a practical scale. Generally, MOFs particles are usually filled in

41

pipes in the form of powders, which can block mass transfer pathways and make it difficult to

42

recycle. Therefore, it is urgent to prepare MOFs based freestanding membranes with open,

43

porous, and low-tortuosity microstructure.

1,2

Adsorption is thought to be one of the most promising

ACS Paragon Plus Environment

4

Page 5 of 49

Environmental Science & Technology

44

Wood is a widely used natural material in our daily lives. 21 Because of its large number of 3D

45

openings and low bending lumen (channels) along the growth direction,22-24 wood is an ideal

46

substrate for diverse applications, including solar steam generation,25 transparent building

47

materials,26,27 and green chemical reactors28,29 etc. We envisaged that a composite membrane

48

with high transport and adsorption properties can be obtained by growing MOFs materials in situ

49

in wood channels. However, so far, there have been no reports on the use of MOFs/wood

50

membrane to remove organic pollutants from aqueous solutions.

51

Herein, we developed a 3D Zr-MOFs/wood (UiO-66/wood) composite membrane for rapid and

52

highly efficient organic removal by in situ growth of water-stable UiO-66 MOFs in the 3D wood

53

channel arrays. Cellulose is the major constituent of wood cell walls, which contains a large

54

number of hydroxyl groups30 that can coordinate zirconium ions (IV, Zr4+). The coordinated Zr4+

55

ions in wood channels further reacted with organic ligands to form 3D UiO-66/wood membrane.

56

Due to the open and aligned structure of wood channels,31 the contact probability between

57

organic pollutants and the modified UiO-66 MOFs nanoparticles is increased, which leads to the

58

efficient adsorption of organic pollutants. Additionally, the high natural abundance of wood

59

helps to reduce the cost of organic pollutant removal and achieve large-scale treatment in

60

practical applications. As a proof of concept, we designed an all-in-one filter by assembling three

61

pieces of UiO-66/wood membranes for efficient removal of several organic pollutants in water.

62

The efficiency of the UiO-66/wood membrane was over 96.0% for all organic pollutants tested at

63

a flux of 1.0 ×103 L∙m-2∙h-1, showing broad application prospects for the practical water treatment.

64

32, 33

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 49

66

Materials and Methods

67

Materials and chemicals. All reagents and solvents were used without further purification

68

unless otherwise mentioned. The basswood used in this study was obtained from the Chenlin

69

Wood Company. Terephthalic acid (TPA, 98.5%) and propranolol hydrochloride were purchased

70

from the J&K Scientific Ltd. 1-naphthyl amine (1-NA), Bisphenol A (BPA), Bisphenol S (BPS),

71

zirconium (IV) chloride (ZrCl4, 98%), acetic acid (HAc, 99%), rhodamine 6G (Rh6G), methanol,

72

and N, N-dimethyl formamide (DMF) were obtained from Tianjin Med.

73

Preparation of the UiO-66/wood membrane. A piece of natural basswood (diameter: 13 mm,

74

thickness: 5 mm) was immersed in 30 mL DMF containing 366 mg ZrCl4, 258 mg terephthalic

75

acid, and 2.68 mL acetic acid. Then, the mixture was transferred into a sealed autoclave and

76

placed in a muffle at 120℃ for 24 h. The UiO-66 MOFs nanoparticles were formed in-situ in the

77

wood matrix.

78

methanol (MeOH) to remove free UiO-66 and excessive precursors. The Zr content of the UiO-

79

66/wood membrane was tested by inductively coupled plasma mass spectrometry (ICP-MS,

80

Perkin-Elmer Elan DRC II). The content of UiO-66 MOFs in the prepared membrane was 2.22

81

wt%.

82

Water treatment measurements. Pollutants (propranolol, 1-NA, BPA, BPS) were dissolved

83

directly in deionized filtered water (DI water) to prepare the water samples polluted with the

84

organic molecule (0.1 mmol·L-1). Rh6G-polluted water (10 mg·L-1) was used to investigate the

85

absorption isotherm and the kinetic process. The pH of the mixture solution was adjusted using

86

HCl or NaOH solutions (0.1 mol·L-1). The Rh6G, propranolol, 1-NA, BPA, and BPS contents of

87

the water before and after treatment with the UiO-66/wood membrane were determined by UV-

34,35

Finally, the prepared UiO-66/wood membrane was washed with DMF and

ACS Paragon Plus Environment

6

Page 7 of 49

Environmental Science & Technology

88

visible adsorption spectroscopy (UV-vis). The removal efficiency was calculated using the

89

following equation: 6,36

90

Removal efficiency (%) = 100*(C0-C)/C0

(1)

91

where C0 is the initial concentration of pollutant and C is the pollutant concentration after

92

filtration. An all-in-one three-layer filter was designed by assembling three pieces of UiO-

93

66/wood membranes together. A syringe pump was used to control the flux. The removal

94

efficiency of the three-layer filter for various organic pollutants (Rh6G, propranolol, 1-NA, BPA

95

and BPS) was evaluated.

96

Characterization. X-ray diffraction (XRD) patterns were recorded using the Cu Ka radiation

97

(λ=1.5418 Å) on a Bruker AXS D8 advance diffractometer. The morphology of the samples was

98

investigated by field-emission scanning electron microscopy (FE-SEM, FEI, Sirion 200). The

99

UV-vis curves were obtained with a UV 1750 spectrometer.

100 101

Results and Discussion

102

The schematic of the in situ synthesis UiO-66 MOFs nanoparticles in 3D wood microchannels

103

for rapid and highly efficient organic pollutants removal is shown in Figure 1. A piece of natural

104

basswood was immersed in the precursor solution of UiO-66 MOFs until adsorption saturation

105

was reached (Figure 1a). The UiO-66 MOFs nanoparticles are in-situ formed in the wood matrix

106

after heating at 120 0C for 24 hours

107

enlarged schematic of the UiO-66/wood membrane exhibits the numerous elongated and open

108

lumens that are uniformly decorated with the UiO-66 MOFs nanoparticles. As shown in Figure

109

1c, an all-in-one filter device was designed by assembling three pieces of the UiO-66/wood

110

membranes together for scalable applications. Furthermore, the prepared UiO-66/wood

34,35

and the color of wood turns brown (Figure 1b). The

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 49

111

membrane can be readily regenerated by cleaning with methanol several times and can be scaled

112

up by adjusting the number of layers of the UiO-66/wood membrane for industrial applications.

113 114

Figure 1. Schematic showing the synthesis process of the UiO-66 MOFs in a 3D wood membrane for rapid

115

and highly efficient organic pollutants removal. (a) A piece of natural wood is absorbed with the precursor

116

solution of UiO-66 MOFs, which contains ZrCl4, TPA, DMF, and HAc. (b) UiO-66 MOFs nanoparticles in situ

117

formed in the 3D wood matrix through a solvothermal treatment. The zoomed-in image shows the microstructure of

118

the UiO-66/wood membrane uniformly decorated with the UiO-66 MOFs nanoparticles in the numerous aligned and

119

open microchannels. (c) All-in-one device for large-scale organic pollutants removal is built by assembling three

120

pieces of UiO-66/wood membranes into a filter.

121

A piece of basswood with a diameter of 13 mm and a thickness of 5 mm was cut along its growth

122

direction to prepare the UiO-66/wood membrane. Figure S1 shows the microstructure of natural

123

basswood. Figure 2a shows the photo images of the natural basswood (yellow) and UiO-66/wood

124

membrane (brown). Figure 2b shows the SEM image of the UiO-66/wood membrane. There are

125

numerous aligned and open microchannels in the wood framework, which are highlighted with

126

red dash lines. As shown in Figure S2, FT-IR spectra showed that a new absorption band at

127

around 660 cm−1 was observed, which is assigned to Zr-O stretching vibration.

ACS Paragon Plus Environment

37

The XPS

8

Page 9 of 49

Environmental Science & Technology

128

spectra shows that the bonding energy of O 1s shifts from 531.2 eV for natural wood to 531.6 eV

129

for Zr-wood (Figure S3), indicating that the bond of Zr-O was formed. 38,39 The coordinated Zr4+

130

ions in wood channels further reacted with organic ligands, thus to achieve the in situ synthesis

131

of UiO-66 MOFs nanoparticles within the aligned microchannels. Magnified SEM image of the

132

UiO-66/wood membrane shows that the UiO-66 MOFs nanoparticles are uniformly decorated on

133

the surface of the multiple wood channels (Figure 2c, Figure S4). The content of elemental Zr in

134

UiO-66/wood membrane is 0.73 wt% from the ICP-MS test. In addition, the loading amount of

135

UiO-66 MOFs in wood based on the Zr6O4(OH)4(bdc)6 formula, is 2.22 wt% (Table S1). The

136

size of the UiO-66 MOFs nanoparticles is approximately in the range of 130 ± 10 nm (Figure 2d,

137

Figure S5). The transmission electron microscopy (TEM) image shows that the morphology of

138

the UiO-66 MOFs nanoparticles is cubic (inset of Figure 2d). The elemental maps show the

139

coexistence of Zr, C, and O in the UiO-66/wood membrane (Figure 2e), which was further

140

confirmed by the Energy-Dispersive X-ray (EDX) spectra (Figure S6). In addition, the uniform

141

distribution of Zr in the wood microchannels further confirmed the uniform distribution of UiO-

142

66 nanoparticles in the wood matrix. The structure of the UiO-66/wood composite was

143

determined by the X-ray diffraction (XRD) analysis. The results demonstrate that the XRD

144

pattern of the UiO-66/wood composite is in good agreement with the standard patterns of UiO-

145

66 MOFs and wood (Figure 2f). 40,41 Furthermore, the stability of UiO-66/wood composite was

146

investigated in solution at different pH values (pH =1, 7 or 14) and characterized using XRD.

147

Figure S7 shows that the samples maintain their crystalline structure after immersion in the

148

solution at different pH values for 24 hours.

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 49

149 150

Figure 2. Characterizations of the UiO-66/wood membrane. (a) Photos of the natural basswood and UiO-

151

66/wood membrane. The color of the basswood turns brown due to the formation of UiO-66 MOFs nanoparticles

152

within the wood framework. The inside of the UiO-66/wood membrane is brown, indicating that the UiO-66 MOFs

153

nanoparticles are evenly distributed throughout the entire wood matrix. (b) SEM image of the UiO-66/wood

154

membrane showing the numerous elongated and low-tortuosity microchannels. The microchannels are highlighted

155

with red dash lines. (c) SEM image showing the UiO-66 MOFs nanoparticles anchored in the wood microchannels.

156

(d) Magnified SEM image of the UiO-66 MOFs nanoparticles. The inset shows the TEM image of the UiO-66

157

MOFs nanoparticle. (e) SEM image and corresponding elemental maps of the UiO-66/wood. (f) XRD patterns of the

158

natural wood, UiO-66 MOFs, and UiO-66/wood membrane.

159

Rh6G (molecular structure shown in Figure S8) is a potential carcinogen. In our study, Rh6G

160

was selected as a pollutant to test the removal efficiency of the UiO-66/wood membrane. Firstly,

161

we studied the kinetics of the adsorption process. As shown in Figure S9a, the concentration of

162

Rh6G in the solution reduced from 10 mg∙L-1 to 0.0096 mg∙L-1 within 5 min. The removal

163

efficiency reached 98.8%. The distribution coefficient (Kd) is 2.6×106 mL∙g-1, which shows that

ACS Paragon Plus Environment

10

Page 11 of 49

Environmental Science & Technology

164

the UiO-66/wood membrane had excellent affinity for Rh6G. 42 Adsorption kinetics studies show

165

that the adsorption process fits well with a pseudo-second-order model, which confirms that the

166

interaction between the UiO-66/wood membrane and Rh6G is mainly controlled by a chemical

167

processes. 43,44 The adsorption isotherm analysis indicates the adsorption mechanism between the

168

Rh6G and UiO-66/wood composite follows the Langmuir model, which confirms that the

169

adsorption is a single-layer pattern process (Figure S9 and S10).44,45 The maximum adsorption

170

capacity of UiO-66/wood based on the Rh6G content of the UiO-66 MOFs is 690 mg∙g-1, which

171

is much higher than other UiO-66 based membranes.

172

water transport within a single vessel was also studied based on an advection-diffusion

173

phenomenon. 5

174

Taking advantage of the open and aligned microstructure of wood, the UiO-66/wood membrane

175

can be used as a filter for efficient pollutants removal from water by filtration. Figure 3a exhibits

176

the images of natural wood and UiO-66/wood membranes for water treatment. The orange

177

aqueous Rh6G solution becomes colorless after flowing through the UiO-66/wood membrane. In

178

contrast, when the aqueous Rh6G solution flows through the natural wood, there is no obvious

179

color change in aqueous solutions containing Rh6G. This result indicates that the UiO-66/wood

180

membrane has efficient removal ability for Rh6G. The structure of crystal unit cell of the UiO-66

181

MOFs is shown in Figure 3b. The blue, red, and white spheres represent the Zr, O, and C atoms,

182

respectively. H atoms on the ligands are omitted for clarity. The large yellow sphere in the

183

middle represents the void cage inside the framework. The UiO-66 MOFs with the

184

Zr6O4(OH)4(bdc)6 (bdc=1,4-benzenedicarboxylate) formula have an fcu topology (Figure S11). 48

185

The maximum diameter of the octahedral cavity of the UiO-66 MOFs is 12.3 Å, which is large

186

enough to adsorb various organic molecules (Table S2). Zeta potential was measured to

46,47

In our previous work, the polluted

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 49

187

determine the surface charge state of the UiO-66 MOFs. As shown in Figure S12, the UiO-66

188

MOFs have a positive potential from pH = 1 to 6 and a negative potential from pH = 7 to 10. The

189

polar surface of the UiO-66 MOFs helps capturing organic molecules and binding with the polar

190

functional groups through electrostatic attraction.49,37 The schematic adsorption process of

191

organic pollutants in the UiO-66/wood membrane is shown in Figure 3c. The uniformly

192

distributed UiO-66 MOFs nanoparticles on the surface of the aligned and low-tortuosity wood

193

microchannels increase the contact area with the organic pollutants, thereby enhancing the

194

extraction of the organic molecules. The unique open and elongated wood channels facilitate the

195

mass transfer of polluted water, which is essential for rapid and highly efficient organic

196

contaminant uptake from water. The adsorption ability of the UiO-66/wood membrane towards

197

Rh6G is further confirmed by the UV-vis spectra (Figure 3d). The characteristic absorbance peak

198

of Rh6G (525 nm) completely disappears after the aqueous Rh6G solutions are treated with the

199

UiO-66/wood membrane. However, there is only a rarely decrease after aqueous Rh6G solutions

200

are treated with natural wood. The UiO-66/wood membrane displays high removal efficiency

201

(98%) for the different concentrations of Rh6G. The removal efficiency of UiO-66/wood

202

membrane decreases only slightly when the concentration of Rh6G is higher than 12 mg L-1

203

(Figure 3e). We also investigated the adsorption properties of the UiO-66/wood membrane for

204

Rh6G solutions at different pH values. As shown in Figure S13, the removal efficiency is

205

dramatically increased when the pH is higher than 4, demonstrating that a weak acid or alkaline

206

environment facilitates the adsorption of Rh6G on the UiO-66/wood membrane. Rh6G is a kind

207

of cationic dye 50 which is easily adsorbed by UiO-66 MOFs in the wood microchannels through

208

electrostatic attraction in the higher pH range since the UiO-66 MOFs have a negative potential

209

in the high pH range.

ACS Paragon Plus Environment

12

Page 13 of 49

Environmental Science & Technology

210 211

Figure 3. Rh6G removal performance of the UiO-66/wood membrane. (a) Comparison of the Rh6G (10 mg∙L-1)

212

adsorption efficiency of the natural wood and the UiO-66/wood membrane. (b) Crystal unit cell of the UiO-66

213

MOFs. (blue: Zr; red: O; white: C; H atoms on ligands are omitted for clarity. The large yellow sphere represents the

214

void cage inside the framework. (c) Schematic showing the absorption process of the organic molecules as the

215

polluted water flows through the UiO-66/wood filter. (d) UV-vis spectra of the Rh6G-contaning solution before

216

(black) and after treatment with natural wood (red) and the UiO-66/wood membrane (blue). (e) Removal efficiency

217

of the UiO-66/wood membrane for different Rh6G concentrations.

218

To demonstrate a practical application, an all-in-one filter was designed and assembled by

219

integrating 3 pieces of the UiO-66/wood membranes in a plastic kettle (Figure 4a). The set-up for

220

organic pollutant removal is shown in Figure 4b. A syringe pump is used to control the water

221

flux. The flux of the treatment is calculated with equation (2):

222

W = V/S

(2)

223

Where V is the solution flux of the syringe pump (L∙h-1), S is the effective area of the UiO-

224

66/wood membrane (m2), and W is the treatment flux (L∙m-2 ∙h-1).

225

Propranolol (molecular structure shown in Figure S14) is a common medicine against

226

hypertension and is hardly removed by common wastewater treatment systems.

227

evaluated the adsorption ability of the three-layer filter for a propranolol solution (0.1 mmol∙L-

ACS Paragon Plus Environment

32

We firstly

13

Environmental Science & Technology

Page 14 of 49

228

1).51

229

with the all-in-one three-layer filter. The characteristic peak of the propranolol solution at 290

230

nm is absent after treatment, indicating efficient removal. The influence of the flux on the

231

removal efficiency of the three-layer filter was further investigated. As shown in Figure 4d, a

232

high removal efficiency over 97% was maintained even when the flux reached 1.0×103 L∙m-2∙h-1,

233

which is significantly higher than previous reports.33 The rapid and highly efficient removal of

234

propranolol from water using the proposed three-layer filter originates from three elements. First,

235

the fine UiO-66 MOFs nanoparticles with a high surface area are uniformly distributed

236

throughout the aligned wood lumens, which enhances the full utilization of UiO-66 MOFs.

237

Secondly, the elongated and irregular multi-channels in the wood matrix contribute to creating

238

sufficient contact between the UiO-66 MOFs nanoparticles and the organic contaminants in

239

water. Finally, the low-tortuosity and aligned microchannels with a favorable hydrophilic

240

property (Figure S15) can facilitate the fast transport of polluted water without significantly

241

sacrificing the removal efficiency. The reusability of water treatment materials is a crucial factor

242

for practical applications. The designed three-layer filter can be readily regenerated by washing

243

with methanol (3×5 mL) and DI water (10 mL). The removal efficiency of the three-layer filter

244

for propranolol was over 95.0% even after 6 cycles of regeneration (Figure 4e, Figure S16). The

245

results demonstrate the excellent reproducibility and reusability of the three-layer filter based on

246

the UiO-66/wood membranes. The stability of the three-layer filter was investigated by SEM. As

247

shown in Figure S17, the morphology and average size of the UiO-66 MOFs nanoparticles is

248

well maintained even after 6 cycles. The XRD pattern of the UiO-66 MOFs is also in good

249

agreement with the standard patterns of UiO-66 MOFs after 6 cycles, demonstrating its excellent

250

structural stability (Figure S18). In addition, the weight loss of Zr in the UiO-66/wood membrane

Figure 4c shows the UV-vis spectra of the propranolol solution before and after treatment

ACS Paragon Plus Environment

14

Page 15 of 49

Environmental Science & Technology

251

after 6 regeneration cycles was measured by ICP-MS and was less than 0.1 wt%, confirming the

252

superior stability of our three-layer filter. Figure 4f compares the removal efficiency for

253

propranolol of activated carbon,32 UV-treated carbon (UV-C),33 natural wood, and our three-

254

layer filter based on the UiO-66/wood membranes. The removal efficiency of our three-layer

255

filter is significantly higher than the other materials.

256 257

Figure 4. Removal performance for propranolol of the all-in-one three-layer filter based on the UiO-66/wood

258

membrane. (a) Photographs of the three-layer filter using three pieces of UiO-66/wood membranes. (b)

259

Experimental set-up for organic contaminants removal using the all-in-one three-layer filter. (c) UV-vis spectra of

260

the propranolol solution before (red) and after (black) treatment using the three-layer filter (flux: 1.0×103 L∙m-2∙h-1).

261

(d) Removal efficiency for propranolol of the three-layer filter at different flux. (e) Removal efficiency for

262

successive regeneration cycles. (f) Comparison of the removal efficiency of activated carbon, UV-C, natural wood,

263

and our three-layer filter.

264

Besides Rh6G and propranolol, we also investigated the adsorption performance of the three-

265

layer filter for other common organic pollutants (Figure 5a). The selected organic contaminants

266

were: 1-NA, a known carcinogen and an azo-dye precursor,51 BPA, an endocrine disrupting

267

chemical present in plastic products,53,54 and BPS, a substitute of BPA in many polycarbonates

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 49

268

suspected to be an endocrine disruptor with greater environmental persistence.55,56 The physical

269

and chemical properties of the selected organic pollutants are shown in Table S3. The removal

270

performance of the three-layer filter for these organic contaminants at different flux was

271

evaluated (Figure 5b, Figure S19-21). The proposed all-in-one three-layer filter displays a high

272

efficiency ( ≥ 96%) for each organic contaminant at a flow rate of 1.0×103 L∙m-2∙h-1, which

273

shows highly outperforms natural wood and pure UiO-66 MOFs particles in powder form

274

(Figure 5c). In addition, we have compared the organic pollutants removal efficiency of UiO-

275

66/wood membrane with other systems such as B-ZnO nanoparticles, zeolitic imidazolate

276

framework-9

277

Fe3O4@SiO2-β-cyclodextrin (Fe3O4@SiO2-PGMACD), and porous β-CD-containing polymer

278

(P-CDP)42, 57-60. As shown in Table S4, the organic pollutants treatment performance of the UiO-

279

66/wood membrane is higher than for other systems. In this work, we proposed a UiO-66/wood

280

membrane for highly efficient organic pollutants removal from water. A three-layer filter was

281

designed, which shows a high removal efficiency over 96.0% for various organic contaminants

282

in water at a treatment flow rate of 1.0×103 L∙m-2∙h-1. The use of the three-layer filter based on

283

the UiO-66/wood membrane offers a scalable, renewable, and cost-effective strategy for rapid

284

and efficient water treatment.

(ZIF-9),

porous

β-cyclodextrin

polymer@cotton

ACS Paragon Plus Environment

(CD-TFP@COTTON),

16

Page 17 of 49

Environmental Science & Technology

285 286

Figure 5. Removal performance of the all-in-one three-layer filter for other organic pollutants. (a)

287

Constitutional formulas of the organic pollutants tested. (b) Removal efficiency of the designed three-layer filter for

288

each organic pollutant at different flow rates. (c) Comparison of the removal efficiency of the three-layer filter,

289

natural wood, and pure UiO-66 MOFs particles in a powder form for several organic pollutants at a flow rate of 1.0

290

×103 L∙m-2∙h-1.

291 292 293

Acknowledgement

294

X. Guo and X. Cai contributed equally to this work. The work was supported by the National

295

Natural Science Foundation of China (21876072). We would like to thank Dr. Mingwei Zhu for

296

the SEM measurement, Dr. Zonglei Zhang for the ICP-MS test. We acknowledge Dr. Wei Dou,

297

Dr. Aijiang Lu and Dr. Weihua Han contributed technical knowledge and understanding of the

298

subject. We thank the Electron Microscopy Centre of Lanzhou University for the microscopy

299

and micro-analysis of our samples.

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 49

301

References

302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345

(1) Schwarzenbach, R. P.; Escher, B. I.; Fenner, K.; et al. The Challenge of Micropollutants in Aquatic Systems. Science, 2006, 313 (5790), 1072-1077. (2) Murray, K. E.; Thomas, S. M.; Bodour, A. A. Prioritizing Research for Trace Pollutants and Emerging Contaminants in The Freshwater Environment. Environ. Pollut. 2010, 158 (12), 3462-3471. (3) Arkas, M.; Allabashi, R.; Tsiourvas, D.; et al. Organic/Inorganic Hybrid Filters Based on Dendritic and Cyclodextrin “Nanosponges” for the Removal of Organic Pollutants from Water. Environ. Sci. Technol. 2006, 40 (8), 2771-2777. (4) Morin-Crini, N.; Crini, G. Environmental Applications of Water-insoluble β-cyclodextrinepichlorohydrin Polymers. Prog. Polym. Sci. 2013, 38 (2), 344-368. (5) Chen, F. J.; Gong, A. S.; Zhu, M.; et al. Mesoporous, Three-Dimensional Wood Membrane Decorated with Nanoparticles for Highly Efficient Water Treatment. ACS Nano. 2017, 11 (4), 4275-4282. (6) Zhang, W. J.; Yang, X. Y.; Wang, D. S. Complete Removal of Organic Contaminants from Hypersaline Wastewater by the Integrated Process of Powdered Activated Carbon Adsorption and Thermal Fenton Oxidation. Ind. Eng. Chem. Res. 2013, 52 (16), 5765-5771. (7) Chen, Y. F.; Chen, F.; Zhang, S. H.; et al. Facile Fabrication of Multifunctional Metal– Organic Framework Hollow Tubes to Trap Pollutants. J. Am. Chem. Soc. 2017, 139 (46), 16482-16485. (8) Farha, O. K.; Eryazici, I.; Jeong, N. C.; et al. Metal-Organic Framework Materials with Ultrahigh Surface Areas: is the Sky the Limit? J. Am. Chem. Soc. 2012, 134 (36), 1501615021. (9) Alaerts, L.; Maes, M.; Giebeler, L.; et al. Selective Adsorption and Separation of OrthoSubstituted Alkylaromatics with the Microporous Aluminum Terephthalate MIL-53. J. Am. Chem. Soc. 2008, 130 (43), 14170-14178. (10) Zhu, G. W.; Ng Cheng, W.; Lin, W. Y.; Koh, S. N.; Wang, C. H. Effective Recovery of Vanadium from Oil Refinery Waste into Vanadium-Based Metal-organic Frameworks. Environ. Sci. Technol. 2018, 52 (5), 3008-3015. (11) Creamer, A. E.; Gao, B. Carbon-based Adsorbents for Postcombustion CO2 Capture: A Critical Review. Environ. Sci. Technol. 2016, 50 (14), 7276-7289. (12) Mondloch, J. E.; Katz, M. J.; Isley III, W. C.; et al. Destruction of Chemical Warfare Agents Using Metal-organic Frameworks. Nat. Mater. 2015, 14 (5), 512. (13) Liu, K.; Gao, Y. X.; Liu, J.; Wen, Y. F.; Zhao, Y. C.; Zhang, K. Y.; Yu, G. Photoreactivity of Metal-organic Frameworks in Aqueous Solutions: Metal Dependence of Reactive Oxygen Species Production. Environ. Sci. Technol. 2016, 50 (7), 3634-3640. (14) Horcajada, P.; Chalati, T.; Serre, C.; et al. Porous Metal-organic Framework Nanoscale Carriers as a Potential Platform for Drug Delivery and Imaging. Nat. Mater. 2010, 9 (2), 172. (15) Tong, M.; Liu, D.; Yang, Q.; et al. Influence of Framework Metal Ions on the Dye Capture Behavior of MIL-100 (Fe, Cr) MOF Type Solids. J. Mater. Chem. A, 2013, 1 (30), 85348537. (16) Jhung, S. H.; Lee, J. H.; Yoon, J. W.; et al. Microwave Synthesis of Chromium Terephthalate MIL-101 and Its Benzene Sorption Ability. Adv. Mater. 2007, 19 (1), 121124.

ACS Paragon Plus Environment

18

Page 19 of 49

346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391

Environmental Science & Technology

(17) Liu, B. J.; Yang, F.; Zou, Y. X.; et al. Adsorption of Phenol and P-nitrophenol from Aqueous Solutions on Metal-organic Frameworks: Effect of Hydrogen Bonding. J. Chem. Eng. Data, 2014, 59 (5), 1476-1482. (18) Han, S.; Lah, M. S. Simple and Efficient Regeneration of MOF-5 and HKUST-1 via Acidbase Treatment. Cryst. Growth. Des. 2015, 15 (11), 5568-5572. (19) Bhunia, M. K.; Hughes, J. T.; Fettinger, J. C.; et al. Thermochemistry of Paddle Wheel MOFs: Cu-HKUST-1 and Zn-HKUST-1. Langmuir, 2013, 29 (25), 8140-8145. (20) Russell, A. P.; Richard, L. G. Formation and Structure of Wood. Am. Chem. Soc. Washington, DC, 1984. (21) Zhu, H. L.; Luo, W.; Ciesielski, P. N.; et al. Wood-derived Materials for Green Electronics, Biological Devices, and Energy Applications. Chem. Rev. 2016, 116 (16), 9305-9374. (22) Zhu, H. L.; Fang, Z. Q.; Wang, Z.; et al. Extreme Light Management in Mesoporous Wood Cellulose Paper for Optoelectronics. ACS Nano. 2015, 10 (1), 1369-1377. (23) Chen, C.; Zhang, Y.; Li, Y.; et al. Highly Conductive, Lightweight, Low-tortuosity Carbon Frameworks as Ultrathick 3D Current Collectors. Adv. Energy Mater. 2017, 7 (17), 1700595. (24) Chen, C. J.; Li, Y. J.; Song, J. W.; et al. Highly Flexible and Efficient Solar Steam Generation Device. Adv. Mater. 2017, 29 (30), 1701756. (25) Yu, Z.; Yao, Y.; Yao, J.; et al. Transparent Wood Containing CsxWO3 Nanoparticles for Heat-shielding Window Applications. J. Mater. Chem. A, 2017, 5 (13), 6019-6024. (26) Zhu, M. W.; Song, J. W.; Li, T.; et al. Highly Anisotropic, Highly Transparent Wood Composites. Adv. Mater. 2016, 28 (26), 5181-5187. (27) Zhu, M.; Wang, Y.; Zhu S, et al. Anisotropic, Transparent Films with Aligned Cellulose Nanofibers. Adv. Mater. 2017, 29 (21), 1606284. (28) Wang, Y.; Sun, G.; Dai, J.; et al. A High-performance, Low-tortuosity Wood-carbon Monolith Reactor. Adv. Mater. 2017, 29 (2), 1604257. (29) Zhu, M. W.; Li, T.; Davis, C. S.; et al. Transparent and Haze Wood Composites for Highly Efficient Broadband Light Management in Solar Cells. Nano Energy, 2016, 26, 332-339. (30) Kong, W. Q; Wang, C. G; Jia, C; et al. Muscle-Inspired Highly Anisotropic, Strong, IonConductive Hydrogels. Adv. Mater. 2018, 30 (39), 1801934. (31) Sjostrom, E. Wood Chemistry: Fundamentals and Applications. Gulf professional publishing. 1993. (32) Valenzano, L.; Civalleri, B.; Chavan, S.; et al. Disclosing the Complex Structure of UiO-66 Metal Organic Framework: A Synergic Combination of Experiment and Theory. Chem. Mater. 2011, 23 (7), 1700-1718. (33) Zhu, H.; Yang, X.; Cranston, E. D.; et al. Flexible and Porous Nanocellulose Aerogels with High Loadings of Metal-organic Framework Particles for Separations Applications. Adv. Mater. 2016, 28 (35), 7652-7657. (34) Aguilera-Sigalat, J.; Bradshaw, D. A Colloidal Water-stable MOF as A Broad-range Fluorescent pH Sensor via Post-synthetic Modification. Chem. Commun. 2014, 50 (36), 4711-4713. (35) Kandiah, M.; Nilsen, M. H.; Usseglio, S.; et al. Synthesis and Stability of Tagged UiO-66 Zr-MOFs. Chem. Mater. 2010, 22 (24), 6632-6640. (36) Sirés, I.; Oturan, N.; Oturan, M. A. Electrochemical Degradation of β-blockers. Studies on Single and Multicomponent Synthetic Aqueous Solutions. Water research, 2010, 44 (10), 3109-3120.

ACS Paragon Plus Environment

19

Environmental Science & Technology

392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437

Page 20 of 49

(37) Han, Y; Liu, M.; Li, K.; et al. In situ synthesis of titanium doped hybrid metal–organic framework UiO-66 with enhanced adsorption capacity for organic dyes. Inorg.Chem. Front. 2017, 4 (11), 1870-1880. (38) BE Lookup Table for Signals from Elements and Common Chemical Species. XPS international, Inc. 1999. (39) Wagner, C. D.; Riggs, W. M.; Davis, L. E.; et al. Handbook of X-ray Photoelectron Spectroscopy. Perkin Elmer: Eden Prairie, MN, 1979. (40) http://muchong.com/html/201709/9767241.html. (41) Chowdhury, P.; Mekala, S.; Dreisbach, F.; et al. Adsorption of CO, CO2 and CH4 on CuBTC and MIL-101 metal organic frameworks: Effect of Open Metal Sites and Adsorbate Polarity. Microporous and Mesoporous Materials, 2012, 152, 246-252. (42) Shin, Y. S; Fryxell, G. E.; Um, W.; et al. Sulfur-Functionalized Mesoporous Carbon. Adv.Funct.Mater. 2007, 17 (15), 2897-2901. (43) Zou, Y. D.; Wang, X. X.; Ai, Y. J.; et al. β-Cyclodextrin modified graphitic carbon nitride for the removal of pollutants from aqueous solution: experimental and theoretical calculation study. J. Mater. Chem. A. 2016, 4 (37), 14170-14179. (44) Yu, D. Y.; Wang, L. L.; Wu, M. H.; Simultaneous removal of dye and heavy metal by banana peels derived hierarchically porous carbons. J. Taiwan Inst. Chem. Eng. 2018, 93, 543-553. (45) Wang, B.; Jiang, Y. S; Li, F. Y; et al. Preparation of biochar by simultaneous carbonization, magnetization and activation for norfloxacin removal in water. Bioresour Technol. 2017, 233, 159-165. (46) Tian, C; Xiang, X; Wu, J; et al. Facile Synthesis of MoS2/CuS Nanosheet Composites as an Efficient and Ultrafast Adsorbent for Water-Soluble Dyes. Journal of Chemical & Engineering Data 2018, 63 (10), 3966-3974. (47) Liu, F. Z; Leung, Y. H.; Djurišić, A. B.;et al. Native Defects in ZnO: Effect on Dye Adsorption and Photocatalytic Degradation. J. Phys.Chem. C 2013, 117 (23), 12218-12228. (48) Bajpai, A; Mukhopadhyay, A.; Krishna, M. S.; et al. A Fluorescent Paramagnetic Mn Metal–organic Framework Based on Semi-rigid Pyrene Tetracarboxylic Acid: Sensing of Solvent Polarity and Explosive Nitroaromatics. IUCrJ, 2015, 2 (5), 552-562. (49) Yang, Y.; Hu, G.; Chen, F.; et al. An Atom-scale Interfacial Coordination Strategy to Prepare Hierarchically Porous Fe3O4-graphene Frameworks and Their Application in Charge and Size Selective Dye Removal. Chem. Commun. 2015, 51 (76), 14405-14408. (50) Yang, Y. M; Hu, G. W; Chen, F. J; et al. An atom-scale interfacial coordination strategy to prepare hierarchically porous Fe3O4-graphene frameworks and their application in charge and size selective dye removal. Chem Commun 2015, 51 (76), 14405-14408. (51) Dantas, R. F.; Rossiter, O.; Teixeira, A. K. R.; et al. Direct UV Photolysis of Propranolol and Metronidazole in Aqueous Solution. Chem. Eng. J. 2010, 158 (2), 143-147. (52) Kandiah, M.; Nilsen, M. H.; Usseglio, S.; et al. Synthesis and Stability of Tagged UiO-66 Zr-MOFs. Chem. Mater. 2010, 22 (24), 6632-6640. (53) Occupational Safety and Health Administration (OSHA) Standard, USA. Toxic and Hazardous Substances: 13 Carcinogens (4-Nitrobiphenyl, etc.). Standard number 1910.1003. http://www.osha.gov/pls/oshaweb/owadisp.show_document?p_table=STANDARDS&p_id =10007, 2012. (54) Vandenberg, L. N.; Hauser, R.; Marcus; M.; et al. Human Exposure to Bisphenol A (BPA). Reprod. Toxicol. 2007, 24 (2), 139-177.

ACS Paragon Plus Environment

20

Page 21 of 49

438 439 440 441 442 443 444 445 446 447 448 449 450 451 452

Environmental Science & Technology

(55) Liang, L.; Zhang, J.; Feng, P.; et al. Occurrence of Bisphenol A in Ssurface and Drinking Waters and Its Physicochemical Removal Technologies. Front. Environ. Sci. Eng. 2015, 9 (1), 16-38. (56) Ke, M.; Chen, M. Y.; Danzl, E.; et al. Biodegradation of A Variety of Bisphenols under Aerobic and Anaerobic Conditions. Water Sci. Technol. 2006, 53 (6), 153-159. (57) Liu, F.; Leung, Y. H.; Djurišić, A. B.; et al. Native Defects in ZnO: Effect on Dye Adsorption and Photocatalytic Degradation. J. Phys.Chem. C. 2013, 117 (5), 12218-12228. (58) Alzate-Sánchez, D. M.; Smith, B. J.; Alsbaiee, A.; et al. Cotton Fabric Functionalized with a β-Cyclodextrin Polymer Captures Organic Pollutants from Contaminated Air and Water. Chem. Mater. 2016, 28 (22), 8340-8346. (59) Kang, Y.; Zhou, L.; Li, X.; Yuan, J., β-Cyclodextrin-modified hybrid magnetic nanoparticles for catalysis and adsorption. J. Mater. Chem. 2011, 21 (11), 3704. (60) Alsbaiee, A.; Smith, B. J.; Xiao, L. L.; et al. Rapid removal of organic micropollutants from water by a porous beta-cyclodextrin polymer. Nature, 2016, 529 (7585), 190-194.

453

454

455 456 457 458 459 460 461 462 463 464 465 466 467 468 469

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 49

SUPPORTING INFORMATION

470 471 472

In situ growth of Metal-Organic Frameworks in Three-Dimensional Aligned Lumen

473

Arrays of Wood for Rapid and Highly Efficient Organic Pollutants Removal

474 475

Ruixue Guo1†, Xiaohui Cai1†, Hanwen Liu1, Zi Yang1, Yajie Meng 1, Fengjuan Chen1*, Yiju Li2*,

476

Baodui Wang1*

477 478 479 480 481 482 483 484 485 486

1

State Key Laboratory of Applied Organic Chemistry and Key Laboratory of Nonferrous Metal Chemistry and Resources Utilization of Gansu Province, Lanzhou University Gansu, Lanzhou, 730000, P. R. China 2 Department of Materials Science and Engineering, College of Engineering, Peking University, Beijing 100871, P. R. China *Email: [email protected]; [email protected]; [email protected]

487 488 489 490

†These authors contribute equally to this work.

ACS Paragon Plus Environment

22

Page 23 of 49

Environmental Science & Technology

491 492 493

Figure S1. (a) SEM image of natural wood. There are numerous long and irregular

494

microchannels in wood. (b) High resolution SEM image of wood channels. (c, d) SEM images

495

showing the top view of the natural wood.

496 497 498 499 500 501 502 503

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 49

504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529

Figure S2. Infrared spectrum of the natural wood and Zr-wood.

ACS Paragon Plus Environment

24

Page 25 of 49

Environmental Science & Technology

530 531 532

Figure S3. XPS spectra of O 1s in (a) natural wood and (b) Zr-wood.

533 534 535 536 537 538 539 540 541 542 543 544 545 546

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 49

547 548 549

Figure S4. SEM image of UiO-66 MOFs nanoparticles in the microchannels of UiO-66/wood

550

membrane.

ACS Paragon Plus Environment

26

Page 27 of 49

Environmental Science & Technology

552

553 554 555

Figure S5. (a) SEM image of the UiO-66 MOFs nanoparticles in wood channels. (b) Size

556

distribution of the UiO-66 MOFs nanoparticles. The average size of the UiO-66 MOFs

557

nanoparticles is about 130 ± 10 nm.

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 49

558

559 560 561

Figure S6. Energy dispersive X-ray spectroscopy (EDX) spectrum of the UiO-66/wood

562

membrane.

ACS Paragon Plus Environment

28

Page 29 of 49

Environmental Science & Technology

563 564

Figure S7. XRD patterns of UiO-66/wood composites showing the stability after immersed in

565

solution with different pH values for 24 hours.

566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582

ACS Paragon Plus Environment

29

Environmental Science & Technology

583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614

Page 30 of 49

Figure S8. Molecular structure of Rh6G.

ACS Paragon Plus Environment

30

Page 31 of 49

Environmental Science & Technology

615

Contaminant Adsorption experiments

616

Rh6G sorption kinetics. The adsorption kinetic was performed at 25 °C. A piece of UiO-

617

66/wood membrane (20 mg, UiO-66 MOFs: 2.22 wt %) was cut to chips and then added to a 100

618

mL beaker, which included 50 mL aqueous solution of Rh6G (10 mg∙L-1). The mixture was

619

shaken for 2 hours. The supernatant was taken out at different specific time periods and filtered.

620

The concentration of Rh6G in filtrate was analyzed by UV-vis spectra (λ = 525 nm).

621

The distribution coefficient Kd is an important parameter to measure the performance of the

622

adsorbents. The value of Kd can be determined by 39

623

Kd = (Ci-Cf)/Cf × (V/m)

(1)

624

where Ci indicates the initial concentration of the contaminant (mg∙L-1), Cf indicates the final

625

equilibrium concentration of the contaminant (mg∙L-1), m is the mass of the adsorbent (g), and V

626

is the volume of the tested liquid (mL).

627

the adsorbed amount for Rh6G (qt) was determined by the below equation.

628 629

qt= (Co-Ct)/m × V

(2)

630 631

where Co and Ct are the concentration of the Rh6G (mg∙L-1) initially and at time t, respectively, V

632

is the volume solution used (L), and m is the mass of UiO-66 MOFs (g).

633

Mathematical Models.The adsorption kinetics of pseudo-first-order rate equation and pseudo-

634

second-order rate equation was used to investigate the adsorption process. The adsorption kinetic

635

equations are given as Eq 3 and Eq 4, respectively.1,2

636 637

Log (qe - qt) = logqe –k1t

(3)

638

t/qt =1/(k2qe2) + t/qe

(4)

639 640

where qe and qt are the amount of Rh6G adsorbed at equilibrium and time t (min), and k1 (min-1)

641

and k2 (g∙mg-1∙min-1) are the rate constant of the pseudo-first-order adsorption and the pseudo-

642

second-order rate constant, respectively.

643

Rh6G adsorption isotherm. The adsorption isotherm of Langmuir and Freundlich models was

644

used to further understand the adsorption mechanism. The adsorption isotherm equations are

645

given as Eq 5 and Eq 6, respectively.3,4

ACS Paragon Plus Environment

31

Environmental Science & Technology

Page 32 of 49

646 647

Ce/qe = 1/(kLqm) + Ce/qm

(5)

648

logqe = logKf + (1/n)logCe

(6)

649 650

where qe and qm represent the equilibrium and the maximum adsorption capacity (mg∙g-1),

651

respectively. Here, Ce is the Rh6G concentration (mg∙L-1) at equilibrium, and KL (L∙mg-1) is the

652

Langmuir constant. Also, Kf (L∙mg-1) and n are the Freundlich parameters related to adsorption

653

capacity and adsorption intensity, respectively.

654

According to the above equations, conclusion can be drawn that:

655

(1) The Kd values of the UiO-66/wood membrane for Rh6G can reach 2.6 × 106 mL∙g-1.

656

(2) The pseudo-second-order rate model and the Langmuir model were selected as the adsorption

657

kinetic and the adsorption isotherm between Rh6G and UiO-66/wood composite.

658

(3) The adsorbed capacity of UiO-66/wood based on the content of UiO-66 MOFs for Rh6G is

659

up to 690 mg∙g-1.

ACS Paragon Plus Environment

32

Page 33 of 49

Environmental Science & Technology

660 661

Figure S9. (a) The adsorption kinetic of UiO-66/wood composite for Rh6G at the initial

662

concentration of 10 mg∙L-1. (b) Adsorption curve of Rh6G versus contact time using UiO-

663

66/wood composite. Inset: Pseudo-second-order kinetic plot for Rh6G adsorption. (c) Adsorption

664

isotherm of UiO-66/wood composite for Rh6G. (d) Linear regression by fitting the data with

665

Langmuir adsorption model.

666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681

ACS Paragon Plus Environment

33

Environmental Science & Technology

Page 34 of 49

682 683 684

Figure S10. (a) Pseudo-first-order kinetic plot and (b) Freundlich isotherms for the adsorption

685

mechanism between Rh6G and UiO-66/wood composite (Rh6G concentration: 10 mg∙L-1).

686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709

ACS Paragon Plus Environment

34

Page 35 of 49

710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736

Environmental Science & Technology

Figure S11. The structure of UiO-66 constructed with Zr6 cluster and BDC ligand.

ACS Paragon Plus Environment

35

Environmental Science & Technology

Page 36 of 49

737 738 739

Figure S12. Zeta potential of UiO-66/wood composite in aqueous solution with different pH

740

values.

741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760

ACS Paragon Plus Environment

36

Page 37 of 49

Environmental Science & Technology

761 762

Figure S13. The removal efficiency of the three-layer filter (Flux: 1.0×103 L∙m-2∙h-1) towards

763

Rh6G (10 mg∙L-1) at different pH values.

764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786

ACS Paragon Plus Environment

37

Environmental Science & Technology

Page 38 of 49

787 788

789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810

Figure S14. Molecular structure of the propranolol.

ACS Paragon Plus Environment

38

Page 39 of 49

811 812 813 814 815 816 817

Environmental Science & Technology

Figure S15. Water contact angle of UiO-66/wood membrane.

ACS Paragon Plus Environment

39

Environmental Science & Technology

Page 40 of 49

818

819 820 821

Figure S16. The UV-vis spectra of the propranolol solution before and after treated with the

822

UiO-66/wood membrane for different cycles.

ACS Paragon Plus Environment

40

Page 41 of 49

Environmental Science & Technology

823 824

Figure S17. Different magnified SEM images of the UiO-66/wood membrane after 6 recycle.

825

Inset shows the size distribution of UiO-66 MOFs after 6 recycle.

ACS Paragon Plus Environment

41

Environmental Science & Technology

826 827 828 829 830 831 832 833 834 835 836 837 838

Page 42 of 49

Figure S18. XRD patterns of simulated UiO-66 and UiO-66/wood membrane after 6 cycles.

ACS Paragon Plus Environment

42

Page 43 of 49

Environmental Science & Technology

839

840 841 842

Figure S19. UV-vis spectra of the BPS solution (0.1 mmol∙L-1) before (in black) and after (in

843

colors) the treatment using the UiO-66/wood membrane based three-layer filter at different flux

844

rate.

ACS Paragon Plus Environment

43

Environmental Science & Technology

Page 44 of 49

845

846 847 848

Figure S20. UV-vis spectra of the 1-NA solution (0.1 mmol∙L-1) before (in black) and after (in

849

colors) treated using the UiO-66/wood membrane based three-layer filter at different flux rate.

ACS Paragon Plus Environment

44

Page 45 of 49

Environmental Science & Technology

850

851 852 853

Figure S21. UV-vis spectra of the BPA solution (0.1 mmol∙L-1) before (in black) and after (in

854

colors) the treatment using the UiO-66/wood membrane based three-layer filter at different flux

855

rates.

856 857 858 859 860 861 862 863 864 865

ACS Paragon Plus Environment

45

Environmental Science & Technology

866

Page 46 of 49

Table S1. ICP analysis for the content of Zr and the calculated UiO-66 loading amount. Molar mass of Zr (g∙mol-1)

Molar mass of UiO-66 (g∙mol-1)

Content of Zr (wt%)

Loading amount of UiO-66 (wt%)

91.2

1662.0

0.73

2.22

867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904

ACS Paragon Plus Environment

46

Page 47 of 49

905

906

Environmental Science & Technology

Table S2. The molecular size of various pollutants. 5 Pollutant

molecular size (Å)

1-NA

16.3×9.5×5.8

BPA

12.5×7.9×6.7

BPS

12.5×7.2×6.7

Propranolol Rhodamine 6G

16.2×9.4×6.0 16.7×12.5×11.4

The present molecular size was calculated with Gaussian 09w.

907

908

909

910

911

912

913

914

915

916

917 918

ACS Paragon Plus Environment

47

Environmental Science & Technology

919

Page 48 of 49

Table S3. Physical and chemical properties of the selected organic pollutants. 6 Selected target molecular weight solubility

Rh6G 479.01 Soluble in water

logKow catogery

7.22550 Dye

toxicity in terms of LD50 (mg∙Kg-1) 400

BPA 228 Slightly soluble in water 3.4 Endocrine disruptor

BPS 250.27 Soluble in water

Propranolol 259.01 Soluble in water

3.01140 Endocrine disruptor

2.96840 Beta-blocker

1-NA 143.19 Slightly soluble in water 3.00320 Carcinogen

2000-5000

2830

9.334-9.450

779

920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952

ACS Paragon Plus Environment

48

Page 49 of 49

Environmental Science & Technology

953

Table S4. The organic pollutants removal efficiency of UiO-66/wood membrane compared to

954

other materials reported in literature.

955 956 957 958 959 960 961 962 963 964 965 966 967 968

Materials

Pollutants

Recycling

Reference

Rh6G BPA BPS 1-NA Propranolol

Removal efficiency (%) 98 98 98 98 98

UiO-66/wood membrane

6 6 6 6 6

Our work

B-ZnO nanoparticles

Rh6G

80.7

Not mentioned

Ref 57

Z9-600

Rh6G

91

4

Ref 54

CD-TFP@cotton

BPA

60

4

Ref 55

Fe3O4@SiO2PGMACD

BPA

89

5

Ref 56

P-CDP

BPA BPS 1-NA Propranolol

95 85 92 96

5 5 5 5

Ref 58

Reference (1) Ho, Y. S. Citation review of Lagergren kinetic rate equation on adsorption reactions. Scientometrics. 2004, 59 (1), 171-177. (2) Ho, Y. S.; Ng, J. C. Y.; McKay, G. Kinetics of pollutant sorption by biosorbents: review. Sep. Purif. Rev. 2000, 29 (2), 189-232. (3) Langmuir, B. I. The Constitution and fundamental properties of solids and liquids. J. Am. Chem. Soc. 1916, 38 (11), 2221-2295. (4) Andjelkovic, I.; Tran, D. N. H.; Kabiri, S.; et al. Graphene aerogels decorated with α-FeOOH nanoparticles for efficient adsorption of arsenic from contaminated waters. ACS Appl. Mater. Interfaces. 2015, 7 (18), 9758-9766. (5) Gaussian 09, Revision B.01: Frisch, M. J. et al. Gaussian, Inc.: Wallingford, CT, 2009. (6) https://www.wikipedia.org/?tdsourcetag=s_pctim_aiomsg

ACS Paragon Plus Environment

49