Tuning Pb(II) Adsorption from Aqueous Solutions on Ultrathin Iron

Jan 30, 2019 - ... 828 West Peachtree Street, Atlanta , Georgia 30332 , United States .... and Organophosphate Esters in Office Air and Dust from Swed...
0 downloads 0 Views 1MB Size
Subscriber access provided by Iowa State University | Library

Remediation and Control Technologies

Tuning Pb(II) Adsorption from Aqueous Solutions on Ultrathin Iron Oxychloride (FeOCl) Nanosheets Jinming Luo, Meng Sun, Cody Ritt, Xia Liu, Yong Pei, John C. Crittenden, and Menachem Elimelech Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b07027 • Publication Date (Web): 30 Jan 2019 Downloaded from http://pubs.acs.org on February 4, 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 31

Environmental Science & Technology

1

Tuning Pb(II) Adsorption from Aqueous Solutions

2

on Ultrathin Iron Oxychloride (FeOCl) Nanosheets

3 4 5 6 7

Environmental Science & Technology

8 9

Revised: January 27, 2019

10 11 12

Jinming Luo,†,⊥ Meng Sun,*,‡,⊥ Cody L. Ritt,‡ Xia Liu,ǁ Yong Pei,ǁ John C. Crittenden,†

13

and Menachem Elimelech*,‡

14 15

† Brook Byers Institute for Sustainable Systems and School of Civil and Environmental

16

Engineering, Georgia Institute of Technology, 828 West Peachtree Street, Atlanta, GA

17

30332, United States

18

‡ Department of Chemical and Environmental Engineering, Yale University, New

19

Haven, Connecticut 06520-8286, United States

20

ǁ Department of Chemistry, Key Laboratory for Green Organic Synthesis and

21

Application of Hunan Province, Key Laboratory of Environmentally Friendly Chemistry

22

and Applications of Ministry of Education, Xiangtan University, Xiangtan, Hunan

23

Province 411105, China

24 25

* Corresponding authors:

26

Email: [email protected]

27

Email: [email protected]; Tel. +1 (203) 432-2789

ACS Paragon Plus Environment

Environmental Science & Technology

28

ABSTRACT

29

Structural tunability and surface functionality of layered two-dimensional (2-D) iron

30

oxychloride (FeOCl) nanosheets are critical for attaining exceptional adsorption properties. In

31

this study, we combine computational and experimental tools to elucidate the distinct

32

adsorption nature of Pb(II) on 2-D FeOCl nanosheets. After finding promising Pb(II)

33

adsorption characteristics by bulk FeOCl sheets (B-FeOCl), we applied computational

34

quantum mechanical modeling to mechanistically explore Pb(II) adsorption on representative

35

FeOCl facets. Results indicate that increasing the exposure of FeOCl oxygen and chlorine sites

36

significantly enhances Pb(II) adsorption. The (110) and (010) facets of FeOCl possess distinct

37

orientations of oxygen and chlorine, resulting in different Pb(II) adsorption energies.

38

Consequently, the (110) facet was found to be more selective toward Pb(II) adsorption than the

39

(010) facet. To exploit this insight, we exfoliated B-FeOCl to obtain ultrathin FeOCl

40

nanosheets (U-FeOCl) possessing unique chlorine- and oxygen-enriched surfaces. As we

41

surmised, U-FeOCl nanosheets achieved excellent Pb(II) adsorption capacity (709 mg g-1 or

42

3.24 mmol g-1). Moreover, U-FeOCl demonstrated rapid adsorption kinetics, shortening

43

adsorption equilibration time to one-third of the time for B-FeOCl. Extensive characterization

44

of FeOCl-Pb adsorption complexes corroborated the simulation results, illustrating that

45

increasing the number of Pb-O and Pb-Cl interaction sites led to the improved Pb(II) adsorption

46

capacity of U-FeOCl.

47 48

KEYWORDS: adsorption, iron oxychloride, heavy metal, coordination, EXAFS

49

1

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

50

Environmental Science & Technology

TOC Art

51

2

ACS Paragon Plus Environment

Environmental Science & Technology

52

INTRODUCTION

53

Extensive use of lead (Pb)-containing products such as power cells, mineral resources, and

54

cosmetics1, 2 have accelerated the global occurrence of Pb in the atmosphere, sediments, and

55

waters,3, 4 thereby escalating environmental risks.5-10 Recent increases in the bioaccumulation

56

of Pb in the food chain have given rise to the possibility of increased Pb ingestion by humans.

57

Inherently toxic to brain tissue, the nervous system, and the reproductive system,11 Pb poses

58

significant risks for human health. Therefore, mitigation of Pb contamination has become a

59

vital issue in separation science and environmental remediation.

60

Adsorption, considered a state of the art technology for heavy metal removal,12, 13 is already

61

well-developed and appears to be more economical and effective than other methods.14-18 To

62

date, a number of adsorbents have been developed for aqueous Pb(II) adsorption. Biochar,

63

produced from dehydrated banana peels via hydrothermal carbonization, exhibited an excellent

64

Pb(II) adsorption capacity of 359 mg g-1. The presence of Pb-O and Pb-O-C interactions

65

illustrated an ion exchange mechanism dominating the adsorption of Pb(II) on biochar

66

adsorbents.19 Double network hydrogels, produced from waste cotton fabrics, exhibited a Pb(II)

67

adsorption capacity of 382.8 mg g-1, owing to an abundance of oxygen-containing groups (i.e.,

68

C-O-H, C-O-C, and C=O) across the framework.20 In another work, a carbomethoxy-

69

functionalized metal-organic framework (MOF) exhibited a modest Pb(II) adsorption capacity

70

with relatively high adsorption kinetics (0.162 g mg-1 min-1), while also validating the vital role

71

of O atoms of the carbomethoxy group in capturing Pb(II).21 Additionally, observations of

72

newly developed transition metal (e.g., Mn and Fe) oxide adsorbents indicate that the

73

constituent O atoms could also interact with Pb(II).22, 23

74

Previous studies have presented data exemplifying the effect of Pb(II) interactions with

75

chemical groups or constituent anions (e.g., oxygen in WO3).24 However, fundamental

76

understanding of adsorption mechanisms, which would provide insight into enhancing

77

adsorptive properties, is still lacking. Furthermore, adsorbents enabling multicomponent Pb(II)

78

interactions are rarely reported. In particular, the interaction of constituent anions (e.g., O, Cl) 3

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Environmental Science & Technology

79

during the adsorption process is still unclear; therefore, investigation of the impact of

80

multicomponent interactions on adsorption performance (i.e., adsorption capacity and kinetics)

81

is of great value.

82

Iron oxychloride (FeOCl), possessing a two-dimensional (2-D) nanosheet structure, has

83

drawn increasing attention in recent years due to its unique structure-dependent catalytic

84

activity.25,

85

unique atomic coordination27 are features which may offer exciting opportunities for FeOCl in

86

adsorption processes. The co-existence of Fe(II)/Fe(III) in the architecture of FeOCl is an

87

impetus for research efforts to optimize the bicomponent functionality, possibly leading to the

88

discovery of novel material properties.28, 29 Moreover, chlorine and oxygen atoms diversify the

89

internal coordination environment of FeOCl while also varying coordination modes in

90

accordance with external environmental changes.30 These interesting properties have allowed

91

researchers to introduce various metal ions into the FeOCl nanosheet framework through sheet

92

intercalation and delamination.31,

93

investigation of metal ion interaction with FeOCl may provide insight into novel adsorption

94

mechanisms.

95

26

Attributes such as scalability, compatibility with multiple iron valences, and

32

Inspired by these pioneering works, we believe the

In this work, we first present the Pb(II) adsorption efficiency of bulk FeOCl sheets (B-FeOCl)

96

followed by the exploration of Pb(II) adsorption mechanisms on different FeOCl facets using

97

density functional theory (DFT). Insights gained from DFT informed the exfoliation of B-

98

FeOCl sheets to produce ultrathin FeOCl nanosheets (U-FeOCl). Adsorption performance of

99

both U-FeOCl nanosheets and B-FeOCl sheets was investigated and compared to demonstrate

100

the enhanced Pb(II)-binding interactions offered by U-FeOCl, as was predicted by DFT. This

101

work highlights the novel application of FeOCl for heavy metal adsorption and provides the

102

scientific base for enhancing adsorptive properties of 2-D metal oxychlorides.

103

EXPERIMENTAL SECTION

104

Chemicals and materials. Ferric chloride hexahydrate (FeCl3·6H2O, >98.0%), iron oxide

105

(Fe2O3, ≥99%), magnetic iron oxide (Fe3O4, 97%), lead nitrate (Pb(NO3)2, ≥99.0%), nitric acid 4

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 31

106

(HNO3, 70%), hydrochloric acid (HCl, 37%), sodium chloride (NaCl, ≥99.5%), potassium

107

chloride (KCl, ≥99.0%), calcium chloride (CaCl2, ≥97.0%), magnesium chloride (MgCl2,

108

≥98.0%), sodium hydroxide (NaOH, ≥98.0%), were purchased from Sigma-Aldrich.

109

chemicals are analytical grade and used as received. Water was treated by a Milli-Q ultrapure

110

water purification system (Millipore, Billerica, MA) to produce deionized (DI) water. Stock

111

solutions of Pb(II) were prepared by dissolving Pb(NO3)2 in DI water.

All

112

Synthesis of FeOCl Sheets. B-FeOCl was prepared by a facile annealing method,

113

where a precursor, FeCl3‧6H2O, was placed in a sealed crucible and heated in a microwave

114

oven for 5 min. After heating, the prepared B-FeOCl sheets were allowed to cool to room

115

temperature. U-FeOCl nanosheets were prepared via sonication of B-FeOCl sheets dispersed

116

in DI water for 30 minutes. Dispersed nanosheets were run through several centrifugation-

117

rinsing cycles with DI water. The supernatant was then decanted and the remaining

118

concentrated U-FeOCl solution was vacuum-filtered and dried to remove remnant moisture.

119

Material Characterization. Crystalline phases of the samples were investigated via

120

Rigaku SmartLab X-ray diffractometer (XRD) using Cu Kα radiation (λ = 1.5418 Å). XRD

121

was operated at 45 kV and 200 mA with a scanning speed of 2o min-1 and 2θ range of 5o to 80o.

122

PDXL 2 Rigaku data analysis software was used to obtain lattice constants and crystallite sizes.

123

The sample morphologies were examined using a Hitachi SU8230 UHR Cold Field Emission

124

Scanning Electron S3 Microscope (FESEM) and a FEI Tecnai Osiris 200kV transmission

125

electron microscopy (TEM). High spatial resolution X-ray energy dispersive spectroscopy

126

(EDS) was performed with a Bruker XFlash 5060FQ Annular EDS detector. X-ray

127

photoelectron spectroscopy (XPS) spectra were obtained using monochromatic 1486.7 eV Al

128

Kα X-ray source on a PHI VersaProbe II X-ray photoelectron spectrometer with a 0.47 eV

129

system resolution. The energy scale was calibrated using the Cu 2p3/2 (932.67 eV) peak of a

130

clean copper plate and the Au 4f7/2 (84.00 eV) peak of a clean gold foil. For XPS, a 23.5 eV

131

pass energy, 20 ms integration time, and 100 meV step size were chosen; additionally, the C

132

1s peak was calibrated to 284.8 eV for all spectra obtained. XPS data analysis was performed

133

using the Multipak software provided by Physical Electronics, INC. Fe K-edge X-ray 5

ACS Paragon Plus Environment

Page 7 of 31

Environmental Science & Technology

134

absorption fine structure spectroscopy (XAFS) spectra were collected on beamline 1W1B at

135

Beijing Synchrotron Radiation Facility. The ring storage energy of the synchrotron radiation

136

accelerator during data collection was 2.5 GeV with a current intensity of 50 mA. Fe K-edge

137

spectrum of the sample was collected using transmission mode. Fe foil, Fe3O4, and Fe2O3

138

powder were used as standard compounds. The tested FeOCl-based adsorbents were

139

homogenized and compressed into 2-3 mm slices, bearing a diameter of 1 cm. Athena software

140

was used to process the normalization and liner combination fitting (LCF) of the X-ray

141

absorption near edge structure (XANES) spectra. Inductively coupled plasma mass

142

spectrometry (ICP-MS, ELAN DRC-e, Perkin Elmer, Waltham, MA) was used to quantify the

143

Pb(II) ion concentration. Nitric acid solutions (1% HNO3) consisting of Pb(II) concentrations

144

ranging from 0 to 200 μg L-1 were used to calibrate the system. Indium nitrate (40 μg L-1) was

145

used as an internal standard for ICP-MS.

146

Adsorption Performance Experiments. Pb(II) solution (40 mL) with initial

147

concentrations ranging from 10 to 800 mg L-1, was prepared by dilution of concentrated stock

148

solution with DI water. Adsorption isotherms of Pb(II) at 15, 25, and 35 °C were determined

149

by batch experiments, with an adsorbent dosage of 0.5 g L-1 for all experiments. The adsorbed

150

amount (qe) for Pb(II) was calculated using

151

qe =

V ( C0 − Ce ) m

(1)

152

where V is the volume of the Pb(II) solution, m is the mass of the FeOCl adsorbent, and C0 and

153

Ce are the initial and equilibrium solution concentrations of Pb(II), respectively.

154

Langmuir, Freundlich, and Dubinin–Radushkevich (D-R) isotherm models were used for

155

fitting the temperature-dependent adsorption data. Thermodynamic parameters, such as

156

standard Gibbs free energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°) were calculated

157

according to the van’t Hoff equation (Section 1, Supporting Information). Pb(II) adsorption

158

kinetics were examined with an initial Pb(II) concentration of 25 mg L-1 and placed under

159

constant stirring to reduce mass transfer resistance. To analyze the effect of solution pH on

160

Pb(II) adsorption, isotherms were determined for pH values varying from 1.0 to 5.0. Desired 6

ACS Paragon Plus Environment

Environmental Science & Technology

161

pH was obtained via dropwise addition of 0.1 M HCl or 0.1 M NaOH to the solution. Batch

162

adsorption experiments were run in triplicates under stirring at 240 rpm for 24 h. After

163

equilibration, all samples were filtered through 0.45 µm cellulose acetate membranes to

164

separate the FeOCl nanosheets for analysis of Pb(II) in solution. The impact of competing

165

cations on the adsorption of Pb(II) was evaluated by adding NaCl, KCl, CaCl2, and MgCl2 at

166

concentrations of 0.01 and 0.1 M. Residual concentrations of Pb(II) in solution were

167

determined with the use of ICP-MS.

168

Density Functional Theory Calculations. Structure optimization calculations were

169

performed within the framework of density functional theory (DFT) as implemented in the

170

Vienna Ab initio Simulation Package (VASP).33, 34 Exchange-correlation interactions were

171

treated by generalized gradient approximation (GGA), parameterized by Perdew, Burke, and

172

Ernzerhof (PBE).35 We described the interaction between ions and electrons using the projected

173

augmented wave (PAW).36 The periodic unit was optimized by the conjugate gradient

174

algorithm with an energy cutoff of 500 eV until the force on each atom was less than 0.01 eV

175

Å-1. A vacuum space exceeding 20 Å was employed to minimize interactions between the two

176

periodic units. The Brillouin-zone was meshed by the γ-centered Monkhorst-Pack method with

177

5×5×1 k-points for geometry optimizations. As a typical layered 2-D nanomaterial, FeOCl is

178

constructed by stacking individual layers through the van der Waals force.37 The (010) facet of

179

FeOCl was positioned perpendicular to the Y-axis and terminated via weak interactions with

180

adjacent layers of Cl atoms (Figure S1). The (010) facet of FeOCl was represented as a

181

monolayer slab containing four atomic layers to simplify structural optimization and calculate

182

adsorption energy at the facet.38 In contrast, the in-plane (110) facet possessed strong

183

intermolecular chemical bonds between adjacent atomic layers. Therefore, minimizing the

184

influence of adjacent atoms required an eight-layer slab model to accurately represent the (110)

185

facet structure located at the edge of FeOCl. The three bottom slab layers in the (110) facet

186

were restricted during calculations. Adsorption energy (Ead) on the slab surfaces is defined as

187

Ead = Etot − Esurf − Eatom, where Etot is the total energy of the adsorption model, Esurf is the

188

optimized energy on the slab surface in the absence of Pb atom, and Eatom is the energy of a Pb 7

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Environmental Science & Technology

189

atom under vacuum. Therefore, a more negative Ead is indicative of better Pb(II) adsorption

190

and can provide insight into the mechanisms of adsorption at the different FeOCl facets.39

191

Furthermore, the differential charge density (Δρ), used to calculate the charge distribution

192

across the adsorption complexes, was defined as Δρ = ρ[FeOCl-Pb] −ρ[FeOCl] −ρ(Pb).

193

RESULTS AND DISCUSSION

194

Adsorption of Lead on Bulk FeOCl Slice (B-FeOCl). We applied B-FeOCl, for the first

195

time, toward the application of Pb(II) adsorption. As shown in Figure S2a, adsorption of Pb(II)

196

increases concurrently with solution temperature. The Langmuir isotherm model was a better

197

fit for the experimental data than the Freundlich model (Table S1), suggesting that adsorption

198

of Pb(II) on B-FeOCl occurs in a monolayer of equally available adsorption sites.17 Predictions

199

based on the Langmuir model indicate that the maximum Pb(II) adsorption capacity for B-

200

FeOCl is 409.71 mg g-1 at 35 oC (Table S1). The Dubinin-Radushkevich (D-R) model (Figure

201

S2b) predicts a free energy (E) of 9.92 J mol-1 (Table S2), indicating a chemisorption process.40

202

Thermodynamic parameters, including standard Gibbs free energy (ΔG°), enthalpy (ΔH°),

203

and entropy (ΔS°) of the adsorption process were calculated according to the van't Hoff

204

equation (Figure S2c). For experiments conducted at temperatures of 15, 25, and 35 °C, ΔG°

205

values were calculated as -3.38, -4.82, and -5.71 kJ mol-1, respectively (Table S3).

206

Theoretically, negative ΔG° values indicate that the adsorption of Pb(II) onto B-FeOCl is

207

thermodynamically favorable. ΔH° and ΔS° values were calculated as 30.29 kJ mol-1 and 0.12

208

kJ mol-1 K-1, respectively. The coupling of negative ΔG° and positive ΔH° values suggests that

209

Pb(II) adsorption on B-FeOCl occurs as a spontaneous, endothermic reaction.41 Furthermore,

210

positive ΔS° values reveal a thermodynamically favorable increase in system randomness

211

occurring at the interface between the adsorbent (B-FeOCl) and adsorbate (Pb(II)) as

212

temperature increases.42 These thermodynamic insights predict an essential improvement in

213

Pb(II) adsorption at elevated temperatures, corroborating the temperature dependency of Pb(II)

214

adsorption on B-FeOCl observed initially.

215

TABLE 1 8

ACS Paragon Plus Environment

Environmental Science & Technology

216

Insight into Pb(II)-FeOCl Interactions. Our reported Pb(II) adsorption capacities for B-

217

FeOCl, although higher than materials from many other studies,16, 23, 43-52 were still not the

218

highest (Tables 1 and S4). Therefore, further investigation of FeOCl and its adsorption

219

mechanisms was essential for optimizing the adsorption performance. To gain further insight,

220

a DFT study was conducted to evaluate the surface geometries and atomic configurations of

221

the FeOCl-Pb adsorption complexes. Due to the predominance of (010) and (110) facets in

222

FeOCl, these facets were systematically investigated to assess their influence on Pb(II)

223

adsorption. In brief, one equivalent and two non-equivalent53 adsorption complexes are formed

224

at the (010) and (110) facets, respectively. Pb(II) adsorption at the (010) facet is equivalently

225

bonded with the nearest four Cl atoms (bond length of 2.76 Å) and O atom (bond length of

226

2.88 Å) (Figure 1A), whereas non-equivalent adsorption complexes, found on the (110) facet,

227

result in two forms of Pb(II) complexation. For non-equivalent complexation, Pb(II) may bind

228

with two O atoms and one Cl atom (Figure 1B) or two Cl atoms and one O atom (Figure 1C),

229

denoted as (110)-a and (110)-b, respectively. In (110)-a bonding, Pb(II) is strongly bound to O

230

(bond length of 2.30 Å), but weakly bound to Cl (bond length of 2.86 Å), whereas the

231

interactions of Pb(II) with O (bond length of 2.20 Å) and Cl (bond length of 2.76 Å ) are both

232

enhanced in (110)-b bonding due to a reduction in their respective atomic distances. This

233

enhancement predicts that (110)-b adsorption should be more favorable than both (110)-a and

234

(010) adsorption. Adsorption energies (Ead) between Pb(II) and (010), (110)-a, and (110)-b

235

facets were calculated as -3.69, -3.68, and -3.96 eV, respectively. The more negative adsorption

236

energy between Pb(II) and (110)-b indicates a more stable atomic configuration, thus

237

corroborating our previous prediction that (110)-b is the most favorable site for Pb(II)

238

complexation.

239

To further explore the interactions between Pb(II) and constituent non-metal atoms found in

240

FeOCl, we evaluated the binding energy of Pb-O and Pb-Cl at the (010) and (110)-b facets.

241

Due to the large bond length between Pb and O at the (010) facet, the binding energy of Pb-O

242

was negligible; therefore, the binding energy of Pb-Cl at the (010) facet was calculated as -0.92

243

eV by averaging the overall binding energy of the complex. Due to the identical bond lengths 9

ACS Paragon Plus Environment

Page 10 of 31

Page 11 of 31

Environmental Science & Technology

244

of Pb-Cl on the (010) and (110)-b facets, the binding energy of Pb-O on the (110)-b facet, with

245

a total binding energy of -3.96 eV, can be estimated as -2.11 eV. For the (110)-b facet, the total

246

Pb-O binding energy was slightly more negative than the total Pb-Cl binding, indicating that O

247

sites play a slightly greater role in Pb(II) adsorption at this facet.

248

Adsorption charge densities of complexes on different facets were investigated to elucidate

249

interatomic interactions (Figure 1D-F). It was found that exposed Cl atoms are able to donate

250

electrons as Pb(II) approaches the (010) facet, thus changing the electron cloud density around

251

the Pb(II) atom (Figure 1D). Alternatively, the distance between Pb(II) and O at the (010) facet

252

(2.88 Å) is too substantial to allow for redistribution of the electron cloud density around O.

253

As a result, the FeOCl-Pb adsorption complex on the (010) facet eventually configures into a

254

typical μ4-coordination. In contrast, both O and Cl atoms at the (110)-a and (110)-b facets

255

contribute to the redistribution of the electron cloud around Pb(II) (Figure 1E-F), hence

256

forming a μ3-coordination. Although the two adsorption complexes at the (110) facet show the

257

same coordination configuration, the binding energies for these two complexes are expected to

258

be quite different. We attribute this phenomenon to the self-tuning of atomic distances among

259

O, Cl, and Pb atoms. For example, the atomic distances between Pb(II) and anionic constituents

260

at the (110)-b facet are shorter than the atomic distances at the (110)-a facet, causing more

261

electron accumulation around the Pb(II) atom.

262

DFT calculations demonstrated that the adsorption of Pb(II) on FeOCl is predominantly due

263

to the interactions between Pb and Cl atoms; however, increasing the exposure of O atoms in

264

FeOCl will also largely improve adsorption performance as the most favorable facet for Pb(II)

265

adsorption, (110)-b, is significantly influenced by O. Therefore, increasing both O and Cl sites

266

is requisite for optimizing the Pb(II) adsorption performance of FeOCl.

267

FIGURE 1

268

Preparation and Characterization of FeOCl Nanosheets. Inspired by insights

269

gained from the DFT study, we attempted exposing more Cl- and O-rich surface sites located

270

throughout FeOCl to improve Pb(II) adsorption. B-FeOCl can be regarded as a stacked 10

ACS Paragon Plus Environment

Environmental Science & Technology

271

composite of lamellar FeOCl nanosheets, where the nanosheets are tightly bound through van

272

der Waals forces and the exterior of each nanosheet is decorated by Cl atoms (Figure S3).54 To

273

expose more binding sites, an ultrasonic probe sonicator was employed to exfoliate B-FeOCl

274

sheets into ultrathin FeOCl (U-FeOCl) nanosheets, thus unveiling numerous exterior Cl sites

275

(Figure 2A). Without exfoliation, B-FeOCl were platelet-shaped sheets (Figure 2B, SEM top-

276

view image) possessing a distinct hierarchical structure less than 100 nm in thickness (Figure

277

2B inset, SEM cross-section image). The crystalline structure of B-FeOCl was evidenced by

278

HRTEM images, indicating lattice fringes of 3.44, 2.54, and 1.90 Å, which correspond to the

279

(110), (021), and (200) facets, respectively (Figure 2C). The selected area electron diffraction

280

(SAED) pattern of B-FeOCl presents a highly ordered diffraction spot matrix (Figure 2C, inset).

281

Diffraction spots in a typical parallelogram can be assigned to the orthogonal planes of {200}

282

and {002} as well as their vector sum, the {021} plane, along the {0-12} zone in orthorhombic

283

FeOCl,29 therefore verifying our observations from TEM imaging. In contrast, the morphology

284

and crystal structure of U-FeOCl differed from B-FeOCl, as U-FeOCl consisted of sheets

285

smaller in size (