Selective Nitrate Reduction to Dinitrogen by Electrocatalysis on

Dec 7, 2017 - Excessive nutrients (N and P) are among the most concerned pollutants in surface and ground waters. Herein, we report nanoscale zero-val...
0 downloads 8 Views 785KB Size
Subscriber access provided by READING UNIV

Article

Selective Nitrate Reduction to Dinitrogen by Electrocatalysis on Nanoscale Iron Encapsulated in Mesoporous Carbon Wei Teng, Nan Bai, Yang Liu, Yupu Liu, Jianwei Fan, and Wei-Xian Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b04775 • Publication Date (Web): 07 Dec 2017 Downloaded from http://pubs.acs.org on December 8, 2017

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

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

Page 1 of 21

Environmental Science & Technology

1

2

Selective Nitrate Reduction to Dinitrogen by Electrocatalysis on Nanoscale

3

Iron Encapsulated in Mesoporous Carbon

4

Wei Teng,† Nan Bai, † Yang Liu,‡ Yupu Liu, ‡ Jianwei Fan† and Wei-xian Zhang*,†

5 6 7 8 9



State Key Laboratory for Pollution Control, School of Environmental Science and

Engineering, Tongji University, Shanghai, China 200092 ‡

Department of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and

10

Innovative Materials, and Advanced Materials Laboratory, Fudan University, Shanghai,

11

China 200433

12

ACS Paragon Plus Environment

Environmental Science & Technology

13

ABSTRACT

14

Excessive nutrients (N and P) are among the most concerned pollutants in surface and

15

ground waters. Herein, we report nanoscale zero-valent iron supported on ordered

16

mesoporous carbon (nZVI@OMC) for electrocatalytic reduction of nitrate (NO3-) to nitrogen

17

gas (N2). This material has a maximum removal capacity of 315 mg N/g Fe and nitrogen

18

selectivity up to 74%. The Fe-C nanocomposite is prepared via a postsynthetic modification

19

including carbon surface oxidation, in-situ ammonia prehydrolysis of iron precursor and

20

hydrogen reduction. The synthesized materials have large surface areas (660 – 830 m2/g) and

21

small iron nanoparticles (3 – 9 nm) uniformly dispersed in the carbon mesochannels. The iron

22

loading can be adjusted in the range of 0 to 45%. Results demonstrate that the reaction

23

reactivity of electrocatalysis can be fine-tuned by manipulating iron nanoparticle size, degree

24

of crystallization, as well as porous structure. Meanwhile, the small, uniform, and stable iron

25

nanoparticle promotes fast hydrogen generation for rapid cleavage of the N-O bond.

26

Furthermore, this material can maintain its high performance over repetitive experimental

27

cycles. Results suggest a new approach for fast and eco-friendly nitrate reduction and a novel

28

nZVI application.

29 30

TOC

31 32 33

ACS Paragon Plus Environment

Page 2 of 21

Page 3 of 21

Environmental Science & Technology

34

INTRODUCTION

35

Nitrate is ubiquitous in nature waters. Excessive nitrate leads to eutrophication and algae

36

bloom.1 Together with phosphorus, nutrients are the top pollutants in surface waters in China

37

and to a large degree among the most significant water pollutants globally.2-5 A common

38

approach for remediating water contaminated with nitrate is via denitrification, i.e.,

39

conversion of nitrate to nitrogen gas (N2) by biological and chemical/catalytic reduction.5

40

Biological denitrification is effective for removing nitrate through a stepwise formation of N2.

41

However, this process is slow and sensitive to varying treatment conditions such as dissolved

42

oxygen, temperature and dissolved organic matters.6 Chemical reduction of nitrate can also

43

be achieved by using metallic powder (e.g., Fe,7,

44

excessive generation of ammonium rather than nitrogen, and additional pH adjustment.

45

Meanwhile, catalytic reduction garners increasing attentions, but this process requires

46

hydrogen gas as a reductant or electron donor, and transport and use of pressurized hydrogen

47

have safety issues, which has limited the application in large scale.10 Combined with the

48

advantages of no chemical input, high-efficiency catalysis, ambient operating conditions and

49

minimal sludge generation, electrocatalysis is a promising method for denitrification.11, 12

8

Al9). This approach often results in

50

Various anode electrodes including Cu, Ni, Sn, Ti and Pt have been investigated for the

51

nitrate electroreduction.11, 13, 14 Typical catalytic mechanism is known that a promoter metal,

52

such as Cu, Sn or In,15-17 transforms nitrate to nitrite, which is then reduced to nitrous oxides,

53

nitrogen and ammonium by use of the precious metal (Pd or Pt).6, 18-20 Understandably, the

54

use of precious metal as catalyst comes with high costs. Recently, nanoscale Zero-Valent Iron

55

(nZVI) has been extensively documented for effective transformation of a wide variety of

56

pollutants in water21-23 including various reports on nitrate reduction,7, 24 due to its excellent

57

electron-donating capability. Many studies suggest that iron plate can be considered as a

58

non-precious metal catalyst for electrochemical reduction of nitrate.25 Mechanism is

59

nonetheless unknown concerning electrocatalysis over the iron nanoparticles for nitrate

60

reduction. Virtually all work published so far has focused on Fe(0) as a reducing agent.

61

Catalytic reduction capacity and selectivity for nitrogen gas generation are vital factors for

62

nitrate reduction, which depend on the catalyst properties, e.g., metal loading, metal ACS Paragon Plus Environment

Environmental Science & Technology

63

nanoparticles size, crystallization and electronic properties of the catalyst support.6, 10 Many

64

studies reported that the nanoparticles can be dispersed onto various solid supports to

65

enhance the catalytic performance, including conventional activated carbon,26 alumina,27

66

silica28 and novel supports, such as TiO2,29 ZrO2,30 CeO2,30 graphene oxide32 and so on.

67

Support plays an important role in stabilizing nanoparticles and minimizing particle size to

68

create more active sites. For common support, it is difficult to control the size and uniform

69

dispersity of metal nanoparticles by reasons of irregular pore structures.33-35 The ordered

70

mesoporous carbon is a promising support material due to its unique advantages, involving

71

uniform pore size (2 – 50 nm), regular mesostructure, good conductivity and large surface

72

area and pore volume for metal loading36-39.

73

Herein, we designed and demonstrated an electrocatalytic denitrification using mesoporous

74

carbon supported nanoscale zero-valent iron as catalyst for efficient nitrate reduction

75

(Scheme 1). The novel nanocomposites (nZVI@OMC) are prepared via post-synthetic route.

76

Uniform and ultra-small iron nanoparticles can be well-dispersed into the carbon matrix, and

77

the amount of metal loading, iron particle size and constituents are regulated by varying the

78

amount of iron precursor and hydrogen reduction temperature. The nanocomposites were

79

used as electrode materials for electrocatalytic reduction of nitrate in sodium chloride

80

electrolyte. By additional voltage, the whole reaction carried out in a mild and controllable

81

condition, which is better than direct hydrogen sparging in terms of operational safety. We

82

evaluated the reduction ability and selectivity of nitrogen on a number of nanocomposites

83

with varied nanoparticle sizes, loadings and constituents. The long-term performance was

84

investigated through multiple electrocatalytic cycles. Furthermore, the reduction mechanisms

85

were proposed and verified.

86 87

EXPERIMENTAL SECTION

88

Preparation of nZVI@OMC Materials

89

Ordered bimodal mesoporous carbon as a matrix was prepared first according to the

90

method reported by Liu et al.40 The obtained mesoporous carbon was processed by a surface

ACS Paragon Plus Environment

Page 4 of 21

Page 5 of 21

Environmental Science & Technology

91

oxidation in a 1.0 M of acidic ammonia persulfate solution (in 2 M H2SO4) at 60°C for 3 h to

92

obtain a surface oxygen-containing mesoporous carbon.

93

A series of composite materials of nZVI@OMC were then prepared by impregnation of

94

iron precursor into the surface oxygen-containing mesoporous carbon and followed with

95

ammonia-atmosphere prehydrolysis and thermal reduction in a H2 atmosphere. Typically, 0.6

96

g of the oxygen-containing mesoporous carbon was dispersed in 8 mL of ethanol. Iron

97

precursor Fe(NO3)39H2O (0.9 g) was dissolved in ethanol (8 mL) and then added into the

98

carbon ethanolic solution for continuous stirring at room temperature. Until solvent

99

evaporation, the sample was dried at 60°C under vacuum, and then put into a glass tube,

100

which was in a teflon bottle filled with 10 mL of NH3H2O (14 wt. %), to avoid direct contact

101

between sample and NH3H2O. The bottle was sealed and placed at 60°C for 3h. After being

102

washed with deionized water and ethanol to remove the generated NH4NO3, the collected

103

sample was dried at 100°C overnight. Finally, the sample was heated and reduced at 300 -

104

500°C for 2 h under H2 atmosphere in a tube furnace and the composites of nZVI@OMC

105

were obtained (referred as nZVI@OMC-T). Moreover, the nanocomposites with different

106

iron contents were prepared by adjusting the amount of iron precursor (all heated in H2 at

107

400°C), labeled as nZVI@OMC-x%, where x is the iron content.

108 109

Reduction of Nitrate

110

A 200 mg/L of sodium nitrate stock solution was prepared. Different volumes of the stock

111

solution were added into a beaker to get 50 mL of reaction solution with various

112

concentrations (20-200 mg/L). NaCl was added (0.02 M) as the electrolyte. All the

113

electrochemical tests were performed on a CHI 660D electrochemical analyzer system

114

(Shanghai) at -1.3 V for 24 h. At a predetermined time, about 2 mL of solution was taken out

115

for analysis of the concentration of nitrate, nitrite, and ammonium ions.

116

Platinum and saturated calomel electrodes were used as the counter and reference

117

electrodes in a three-electrode cell, respectively. The material of working electrode was

118

prepared by a mixture of nZVI@OMC composites, acetylene black and polyvinylidene

119

fluoride (dissolved in N-methyl-2-pyrrolidone, 10 g/L) with the ratio of 8:1:1. Then, the

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 21

120

mixture was coated on a nickel foam (1 × 1 cm), followed by dried at ~ 50°C for 6 h and then

121

~ 120°C under vacuum for 12 h. The final working electrode sheet was obtained by pressing

122

the prepared nickel foam under 20 MPa.

123 124

Characterization

125

Transmission electron microscopy (TEM) images were taken by a JEOL JEM 2011 or a

126

JEM 2100F microscope at 200 kV. The samples for TEM analysis were first dispersed in

127

ethanol and then dropped onto carbon films supported on Cu or micro grids. X-ray diffraction

128

(XRD) patterns were carried out with a D8 X-ray diffractometer using Ni-filtered Cu Kα

129

radiation (40 kV, 40 mA), in a scanning range of 20 – 80°. Nitrogen sorption isotherms were

130

measured with a Micromeritics Tristar 3020 analyzer at 77 K. Before the measurements, the

131

samples were degassed under vacuum at 120°C for at least 8 h. The concentration of leaching

132

metal iron in aqueous solution was analyzed by Inductive coupled plasma-atomic emission

133

spectrometer (ICP-AES, Agilent 720ES). More details in characterization are listed in

134

supporting information.

135 136

Analytical Methods and Data Analysis

137

During the nitrite reduction, different products are formed including nitrite, nitrous oxide,

138

nitrogen and ammonium. The gas products such as nitrous oxide and nitrogen are sparged

139

from solution during reaction. Previous literature reported that nitrous oxide eventually

140

reduced to dinitrogen exclusively with excess hydrogen in a sealed batch reactor.6,

141

Therefore, in our experiments, the generated gases were regarded as nitrogen. The

142

concentration of nitrate, nitrite and ammonium was detected by UV-Vis spectrophotometer.

143

The detailed procedures based on previous literature.42 Nitrate removal capacities (Q) of

144

different materials was calculated by the following equation: Q=(C0-Ct)V/mFe. Where C0

145

(mg/L) is the initial concentration, Ct (mg/L) is the concentration at time t, V (L) is the

146

volume of nitrate solution, and mFe (g) stands for the mass of iron in nanocomposites coated

147

on the nickel foam.

148

The selectivity was evaluated by the equation:

ACS Paragon Plus Environment

41

Page 7 of 21

Environmental Science & Technology

S (%) = 149

∆C ( NO3− ) − ∆C ( NO2− ) − ∆C ( NH 4+ ) ×100 ∆C ( NO3− )

150

Where ∆C ( NO3− ) , ∆C ( NO2− ) and ∆C ( NH 4+ ) are the absolute difference of nitrate, nitrite

151

and ammonium concentration before and after reaction, respectively. All reduction

152

experiments were performed at least three times.

153 154

RESULTS AND DISCUSSION

155

Properties of the Nanocomposites

156

Ultra-small zero-valent iron nanoparticles in the carbon mesochannels with high iron

157

content are obtained by following steps (Figure 1A): (i) A bimodal mesoporous carbon matrix

158

is selected and surface-oxidized for an easy loading of metal precursor (Fe(NO3)39H2O) into

159

the mesopores. (ii) The precursor is first converted into hydroxides by the in-situ hydrolysis

160

under ammonia atmosphere to localize the upcoming metal nanoparticles without aggregation.

161

(iii) The sample is subsequently treated at 300 - 500°C in hydrogen leading to nucleation,

162

growth and reduction of metal nanoparticles. Transmission electron microscope (TEM)

163

images (Figure 1B-D) show ordered stripe-like carbon frameworks and very small

164

nanoparticles are uniformly confined in the mesochannels. With the hydrogen reduction

165

temperatures increasing from 300 to 500°C, the sizes of the nanoparticles gradually become

166

larger (~ 3.8, 6.2 and 8.9 nm), but still uniformly distributed across the whole carbon matrix.

167

The size of nanoparticles is too small to be observed in the composite nZVI@OMC-8% with

168

low iron content (Figure S1a), whereas the nanoparticles grow up to larger ones (> 20 nm) in

169

nZVI@OMC-45% with high content by twice impregnation of iron precursor (not shown in

170

article). High-resolution transmission electron microscopy (HRTEM) image clearly shows

171

separated nanoparticles with ultra-small particle sizes (5 - 6 nm) uniformly distributed across

172

the whole mesostructured matrix (Figure 1E). The nanoparticle is crystallized to some extent

173

with the lattice fringe of ~ 2.02 Å, matched to the d110 of Fe phase (Figure 1E inset). Energy

174

dispersive spectroscopy (EDS) spectrum displays the obvious signal of Fe, further

175

conforming the existing of iron (Figure S1b). Field-emission scanning electron microscopy

176

(FESEM) images (Figure S2) exhibit ordered mesostructures in large domains with fully

ACS Paragon Plus Environment

Environmental Science & Technology

177

open pore channels on the surface of carbon matrix without any large particles outside of the

178

mesopores. The dispersed nZVI nanoparticles are elucidated by elemental mapping, further

179

displaying uniformly distributed iron across the matrix (Figure S3). Additionally, surface

180

sensitive X-ray photoelectron spectra (XPS) of the nZIV@OMC exhibits weak peak of iron,

181

demonstrating that most of the metal nanoparticles are located in the carbon frameworks

182

(Figure S4).

183

Wide-angle XRD patterns of the nZVI@OMC nanocomposites at different temperatures in

184

hydrogen reduction show an evolution of the nanoparticles lattice (Figure 1F). At 300°C, no

185

obvious diffraction peaks suggest that the generation of iron nanoparticle is small and

186

well-dispersed. When the nanocomposite is treated at 400°C, the characteristic diffraction

187

peaks at 36 and 44° appear, indexed to the maghemite phase (JCPDS 19-0629) and

188

body-centered cubic α-Fe (JCPDS 06-0696), respectively. After the treatment of hydrogen

189

reduction at 500°C, the intensity of the diffraction peak enhances, indicating a better

190

crystallization. On the basis of the Debye-Scherrer formula, the crystalline sizes of the nZVI

191

nanoparticles are ~ 5.8 and 10.1 nm for nZVI@OMC-400 and 500, in accordance with the

192

TEM results. Furthermore, the XRD pattern (Figure S5) shows that the peak at around 44° is

193

still present in a sample exposed in air for 8 months, suggesting that the nZVI dispersed in the

194

mesopores is more stable than that bare nZVI.

195

Thermogravimetric (TG) curves (Figure S6) in air flows show the change in weight with

196

the increased temperature. Notably, a rapid rise occurs at 200-400°C, implying that

197

zero-valent iron reacts with oxygen to generate iron oxides. The rising percentage is

198

correlated to the iron content. The following stage at 400-550°C shows a fast weight loss,

199

mainly attributed to the carbon frameworks combustion. The platform after 550°C with no

200

obvious change in the weight indicates the carbon is completely removed and only iron

201

oxides left. Based on the results, the iron contents in the nanocomposites can be inferred

202

about ~8, 25 and 45% for nZVI@OMC-8%, -400 and -45%, respectively.

203

The SAXS patterns (Figure S7) of the nanocomposites nZVI@OMC present three

204

scattering peaks assigned to the 10, 11 and 21 reflections, indicating a highly ordered

205

hexagonal mesostructure (space group p6m). N2 sorption-desorption isotherms of all the

206

mesoporous composites nZVI@OMC show typical type-IV curves with distinct hysteresis ACS Paragon Plus Environment

Page 8 of 21

Page 9 of 21

Environmental Science & Technology

207

loops, indicating similar pore structures and features to the pristine carbon matrix (Figure

208

S8A). The decrease of the surface area from ~ 1970 to 660, 770, 830, 1410 and 210 m2/g and

209

pore volume from ~ 1.5 to 0.41, 0.52, 0.57, 1.16 and 0.1 cm3/g is observed for the

210

nanocomposites nZVI@OMC-300, 400, 500, -8% and -45% compared with those of the

211

carbon matrix (Table S1), respectively. Nevertheless, the synthesized materials still possess

212

high surface areas and large mesoporosities. The pore size distribution curves show the

213

primary mesopores decrease from 6.0 to 5.2 - 4.6 nm due to the occupation of the iron

214

nanoparticles and no changes in the secondary mesopores (1.8 nm) after the iron loading

215

(Figure S8B). As a result, the proper iron content can maintain superb pore features for

216

application.

217 218

Nitrate Reduction

219

To identify the electrocatalytic efficiency and selectivity of nitrogen, a series of

220

nanocomposites are synthesized and evaluated as cathode materials for nitrate reduction

221

(Figure 2). Compared with the composites nZVI@OMC pyrolyzed and reduced at different

222

temperatures (300-500°C), the sample nZVI@OMC-400 shows the maximum nitrate removal

223

about 315 mg N/g Fe and the highest nitrogen selectivity of 74%. This phenomenon can be

224

explained that when the nZVI@OMC treated at 300°C, the iron precursors is only partially

225

converted to nZVI, and thus the removal capacity is only 155 mg N/g Fe. Whereas the

226

temperature increased to 500°C, although the degree of crystallization increases a lot, the iron

227

particles grow up to larger ones (~8.9 nm) resulting in a decline in surface active sites, and as

228

a result the removal capacity reduces to 230 mg N/g Fe. Meanwhile, the performances of the

229

nanocomposites

230

nZVI@OMC-400-8% with low iron content shows a relative lower removal of 65 mg N/g Fe.

231

Although this material has a large surface area and pore volume, the iron content is too low to

232

offer enough active sites, resulting in low overall removal efficiency. When the iron content is

233

too much (~ 45%), the removal ability of nZVI@OMC-400-45% reaches ~ 200 mg N/g Fe,

234

but still lower than that of nZVI@OMC-400 with 25% of iron, mainly because too much

235

filling of iron to decrease the surface area (210 m2/g) and pore volume (0.1 cm3/g), leading to

236

obstruction and afterwards that a large number of active iron sites are unavailable in the

with

different

iron

contents

are

also

ACS Paragon Plus Environment

compared.

The

sample

Environmental Science & Technology

237

internal mesochannels. The results suggest that the suitable nanoparticle size and

238

crystallization of iron, and proper pore structural property guarantee sufficient catalytic sites

239

for excellent reduction of nitrate. In addition, the mesoporous carbon loaded with iron oxide

240

nanoparticles is used for comparison. With the same iron content, the sample iron

241

oxide@OMC-400 shows relatively lower removal capacity of ~ 120 mg N/g Fe. That iron

242

species being reduced to zero valence rather than oxide is a key factor for the nitrate

243

reduction by electrochemical catalysis. Finally, the mesoporous carbon matrix without any

244

metal loading (OMC) is employed as electrode material. The removal capacity of nitrate is

245

very low, only 20 mg N/g C. Meanwhile, the nickel foam without any coating shows

246

negligible effect on the removal of nitrate (from initial 50 mg/L to final 48.7 mg/L in 24

247

hours). Results further suggest that the nitrate reduction is mainly contributed to the

248

zero-valent iron but not the carbon matrix or nickel foam. The bare nZVI (about 60 nm) is

249

prepared by sodium borohydride of Fe(III) salts for comparison. Removal capacity of about

250

376 mg N/g Fe and lower nitrogen selectivity (~ 43%), indicating that the mesoporous carbon

251

as the carrier plays a very important role to disperse the nZVI nanoparticles from growth and

252

aggregation, thereby increasing the surface area and active sites for enhanced performance.

253

Additionally, the same system without electrolysis is tested. It is found that the nitrate is not

254

removed, indicating that the adoption of the electrolysis is needed.

255

Among these materials, small differences of nitrogen selectivity (65-75%) are displayed

256

when the nZVI@OMC nanocomposites treated at or above 400°C in hydrogen, whereas the

257

sample nZVI@OMC-300 shows only 40% (Figure 2b). This demonstrates that the treated

258

temperature is important for nitrogen selectivity, mainly due to the better reduction and

259

crystallization of iron nanoparticles at higher temperature. Moreover, the selectivity of iron

260

oxide@OMC-400 is only ~ 35%, further certifying the importance of zero valent iron

261

nanoparticle for the nitrate reduction.

262

The evolution of nitrate, nitrite, ammonia, nitrogen and its selectivity with reaction time is

263

demonstrated during the electrochemical catalysis on the nZVI@OMC-400 (Figure 3). The

264

concentration of nitrate gradually decreases with final removal efficiency of 65%. Reduction

265

products including nitrite, ammonium and nitrogen generate accordingly and their amounts

266

increase gradually. The contents of NH4+ increase at the beginning, reach the maximum about ACS Paragon Plus Environment

Page 10 of 21

Page 11 of 21

Environmental Science & Technology

267

12 h and then decrease to ~10%. More than half products are nitrogen gas and its selectivity

268

from initial 45% to the final ~ 74%. The tendency suggests that nitrate is mainly reduced to

269

NH4+ at first, and gradually transformed to nitrogen. During the process, the concentration of

270

the iron ion in the solution is detected to determine the dissolution of nZVI nanoparticles in

271

the nanocomposite. The dissolvable iron can be negligible (< 1%, initial pH 6.8), implying

272

that the obtained electrons in the process of nitrate reduction is not directly derived from the

273

nZVI, which is very different from the traditional chemical redox through iron consuming,

274

but as a catalyst to speed up the conversion of nitrate into nitrogen.

275

The effect of the nitrate concentration on the removal efficiency is evaluated by using the

276

nanocomposite nZVI@OMC-400. Results (Figure 4a) show that the removal capacity

277

enhances from ~ 130 to 920 mg N/g Fe as the initial concentration of nitrate increases from

278

20 to 200 mg/L. In order to assess the long-term performance, the sample nZVI@OMC-400

279

is employed for reduction of nitrate several times. Results exhibit that the nitrate removal and

280

the nitrogen selectivity decrease somewhat after 5 times of recycling (Figure 4b). But the

281

overall removal capacity remains at over 270 mg N/g Fe (< 20% drop), and nitrogen

282

selectivity is still around 60%. The nanocomposites used repeatedly can be considered as a

283

catalyst and the iron nanoparticles are not consumed in the reduction process, consistent with

284

the result that the soluble iron is low.

285

After electrocatalysis, the original structure of the nZVI@OMC-400 remains from TEM

286

imaging observation (Figure S9). The XPS spectra of Fe(2p) exhibit three peaks at 707, 710.6

287

and 724 eV, corresponding to Fe(0), Fe(II) and Fe(III), respectively (Figure S10). The Fe(II)

288

increases from 49 to 61%, indicating that Fe(0) is partially oxidized during the

289

electrocatalysis. Because XPS is surface sensitive, the majority of iron nanoparticles are

290

protected in the mesochannels and few exposed to bulk solution, which may have

291

over-estimated the percentages of Fe(II) and Fe(III). The XRD patterns were carried out by

292

the electrode slice after use because of very little amount of materials (2-3 mg) are coated

293

(Figure S11). Despite the presence of Ni, the main peaks of the material are almost

294

maintained, indicating little compositional change of iron. Results suggest that generated H2

295

can protect the nZVI from oxidation and reduce the Fe(II) to regenerate Fe(0).

296

As aforementioned the electrochemical catalysis, by additional voltage on the ACS Paragon Plus Environment

Environmental Science & Technology

297

nZVI@OMC nanocomposites, nitrate can be easily and effectively removed and converted to

298

harmless nitrogen without iron consumption. The likely mechanisms are as follows (Scheme

299

1): When the electrolysis begins, a large amount of hydrogen (H2 or H·) and chlorine (Cl2)

300

generate on the surface of working and counter electrode, respectively, in the presence of

301

NaCl as electrolyte

302

44

303

protects the zero-valent iron from oxidation. First, NO3- is reduced at cathode due to the

304

presence of high valence N (+5). If there is only electrolysis in the absence of nZVI (i.e., on

305

OMC or nickel foam), nitrate reduction is negligible with the generated H2 quickly escaped

306

from the solution. When zero-valent irons are on cathode, the nitrate reduction is accelerated

307

with 45% of nitrogen selectivity at the beginning of the reaction. Fe is a relatively reactive

308

metal. Its 3d orbit is unstable and easy to pair with extranuclear electron of oxygen in nitrayte

309

(N-O) first to form Fe(II)O [or Fe(III)O]. The Fe(II)O is likely the intermediate rather than

310

iron oxide, because the iron oxide@OMC sample is observed to exhibit low efficiency for

311

nitrate removal (120 mg N/g Fe). Using iron as catalyst also promotes other reductants, for

312

example the generated hydrogen to break the N-O bond and yields NO2- (NO3- + Fe(0) →

313

Fe(II)O + NO2-). Following a similar mechanism, the O in NO2- pairs sequentially with iron

314

and the N-O bond is broken again by the H2 attack, thus achieving the nitrate reduction to

315

generate NO2-, N2 and NH4+ step by step (NO2- + H+ + 0.5H2 → NO + H2O, 2NO + H2 →

316

N2O + H2O, N2O +H2 → N2 + H2O and NO2- + 5H2O + 6e- → NH3 + 7OH-).45 The optimal

317

particles size (5-6 nm), crystallization and stable dispersion of the nZVI are crucial for the

318

nitrate reduction by offering abundant, effective and stable catalytic sites. Furthermore, the

319

Fe(II)O can be rejuvenated to zero-valent iron by the active hydrogen as the dissolution of

320

iron is not detected. During the whole reaction, nZVI mainly acts as a catalyst rather than a

321

reducing agent. Due to the sustained consumption of nitrate, the remainder constantly

322

migrates to the cathode for reduction due to concentration gradient difference, resulting in the

323

eventually high removal capacity. On the other hand, the generated product NH4+ diffuses to

324

the solution during the reaction. Based on break point chlorination, a part of NH4+ is further

325

oxidized to N2 by generated HClO (Cl2 + H2O → HClO + Cl- + H+ and NH4+ + HClO → N2

326

+ H2O + Cl- + H+).13 After reaction for 24 h, the percentage of ammonia reverses from

(cathode surface: 2H+ → H2 - 2e and anode surface: 2Cl- → Cl2 + 2e).43,

The generated H2 not only provides the reducing agent for nitrate reduction, but also

ACS Paragon Plus Environment

Page 12 of 21

Page 13 of 21

Environmental Science & Technology

327

increase to a modest decline. This is accompanied with continuous nitrogen generation to a

328

maximal yield of 74%. Additionally, the ordered mesoporous carbon cooperated with nZVI

329

enhances the nitrate reduction, because it is not only served as a matrix to confine and

330

disperse nanoparticles, but also offers interconnected space for fast molecular diffusion and

331

transportation to the iron active sites.

332

Compared with conventional electrochemical reduction with precious metals or chemical

333

reduction of nitrate by nZVI, electrocatalysis with nZVI@OMC exhibits significant

334

advantages (Table S2).11, 44 A sample of polluted lake water is taken for electrocatalytic

335

denitrification experiments. The nitrate concentration decreases within 24 hours from 22.5 to

336

9.6 mg/L (268 mg N/g Fe) with the nitrogen selectivity of 91%. The results demonstrate this

337

material has potential for nitrate removal.

338 339

ASSOCIATED CONTENT

340

Supporting Information

341

Supporting information includes chemicals, characterization of the nanocomposites, and

342

consists of 18 pages, 11 figures and 2 tables. This material is available free of charge via the

343

Internet at http://pubs.acs.org.

344 345

AUTHOR INFORMATION

346

Corresponding Author

347

*E-mail: [email protected]. Tel: +86-21-65985885.

348

Notes

349

The authors declare no competing financial interest.

350 351

ACKNOWLEDGEMENTS

352

This work was supported by National Natural Science Foundation of China (NSFC Grants

353

21707104 and 51578398), Fundamental Research Funds for the Central Universities, and

354

State Key Laboratory of Pollution Control and Resource Reuse Foundation (NO. PCRRK

ACS Paragon Plus Environment

Environmental Science & Technology

355

16009).

356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397

REFERENCES (1) McIsaac, G. F.; David, M. B.; Gertner, G. Z.; Goolsby, D. A. Eutrophication - Nitrate flux in the Mississippi river. Nature 2001, 414(6860), 166-167. (2) Burow, K. R.; Nolan, B. T.; Rupert, M. G.; Dubrovsky, N. M. Nitrate in Groundwater of the United States, 1991-2003. Environ. Sci. Technol. 2010, 44(13), 4988-4997. (3) Nolan, B. T.; Ruddy, B. C.; Hitt, K. J.; Helsel, D. R. Risk of nitrate in groundwaters of the United States A national perspective. Environ. Sci. Technol. 1997, 31(8), 2229-2236. (4) Lundberg, J. O.; Weitzberg, E.; Cole, J. A.; Benjamin, N. Nitrate, bacteria and human health. Nat. Rev. Microbiol. 2004, 2(8), 681-681. (5) Kapoor, A.; Viraraghavan, T. Nitrate removal from drinking water - Review. Journal of Environmental Engineering-ASCE 1997, 123(4), 371-380. (6) Shuai, D.; Choe, J. K.; Shapley, J. R.; Werth, C. J. Enhanced Activity and Selectivity of Carbon Nanofiber Supported Pd Catalysts for Nitrite Reduction. Environ. Sci. Technol. 2012, 46(5), 2847-2855. (7) Sohn, K.; Kang, S. W.; Ahn, S.; Woo, M.; Yang, S. Fe(0) nanoparticles for nitrate reduction: Stability, reactivity, and transformation. Environ. Sci. Technol. 2006, 40(17), 5514-5519. (8) Lee, C.; Gong, J.; Chi, V. H.; Oh, D.; Chang, Y. Macroporous alginate substrate-bound growth of Fe-0 nanoparticles with high redox activities for nitrate removal from aqueous solutions. Chem. Eng. J. 2016, 298, 206-213. (9) MURPHY, A. P. Chemical removal of nitrate from water. Nature 1991, 350(6315), 223-225. (10) Prusse, U.; Hahnlein, M.; Daum, J.; Vorlop, K. D. Improving the catalytic nitrate reduction. Catal. Today 2000, 55(1-2), 79-90. (11) Fan, J.; Xu, H.; Lv, M.; Wang, J.; Teng, W.; Ran, X.; Gou, X.; Wang, X.; Sun, Y.; Yang, J. Mesoporous carbon confined palladium-copper alloy composites for high performance nitrogen selective nitrate reduction electrocatalysis. New J. Chem. 2017, 41(6), 2349-2357. (12) Dash, B. P.; Chaudhari, S. Electrochemical denitrificaton of simulated ground water. Water Res. 2005, 39(17), 4065-4072. (13) Li, M.; Feng, C.; Zhang, Z.; Shen, Z.; Sugiura, N. Electrochemical reduction of nitrate using various anodes and a Cu/Zn cathode. Electrochem. Commun. 2009, 11(10), 1853-1856. (14) Virkutyte, J.; Jegatheesan, V. Electro-Fenton, hydrogenotrophic and Fe2+ ions mediated TOC and nitrate removal from aquaculture system: Different experimental strategies. Bioresource Technol. 2009, 100(7), 2189-2197. (15) Deganello, F.; Liotta, L. F.; Macaluso, A.; Venezia, A. M.; Deganello, G. Catalytic reduction of nitrates and nitrites in water solution on pumice-supported Pd-Cu catalysts. Appl. Catal. B-Environ. 2000, 24(3-4), 265-273. (16) Alves Rocha, E. P.; Passos, F. B.; Peixoto, F. C. Modeling of Hydrogenation of Nitrate in Water on Pd-Sn/Al2O3 Catalyst: Estimation of Microkinetic Parameters and Transport Phenomena Properties. Ind. Eng. Chem. Res. 2014, 53(21), 8726-8734. (17) Witonska, I.; Karski, S.; Rogowski, J.; Krawczyk, N. The influence of interaction between palladium and indium on the activity of Pd-In/Al2O3 catalysts in reduction of nitrates and nitrites. J. Mol. Catal. A-Chem. 2008, 287(1-2), 87-94. (18) Jung, S.; Bae, S.; Lee, W. Development of Pd-Cu/Hematite Catalyst for Selective Nitrate Reduction.

ACS Paragon Plus Environment

Page 14 of 21

Page 15 of 21

398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441

Environmental Science & Technology

Environ. Sci. Technol. 2014, 48(16), 9651-9658. (19) Epron, F.; Gauthard, F.; Pineda, C.; Barbier, J. Catalytic reduction of nitrate and nitrite on Pt-Cu/Al2O3 catalysts in aqueous solution: Role of the interaction between copper and platinum in the reaction. J. Catal. 2001, 198(2), 309-318. (20) Jung, S.; Bae, S.; Lee, W. Development of Pd–Cu/Hematite Catalyst for Selective Nitrate Reduction. Environ. Sci. Technol. 2014, 48(16), 9651-9658. (21) Li, X.; Elliott, D. W.; Zhang, W. Zero-valent iron nanoparticles for abatement of environmental pollutants: Materials and engineering aspects. Crit. Rev. Solid State 2006, 31(4), 111-122. (22) Zhang, W. X. Nanoscale iron particles for environmental remediation: An overview. J. Nanopart. Res. 2003, 5(3-4), 323-332. (23) Wang, C. B.; Zhang, W. X. Synthesizing nanoscale iron particles for rapid and complete dechlorination of TCE and PCBs. Environ. Sci. Technol. 1997, 31(7), 2154-2156. (24) Su, C. M.; Puls, R. W. Nitrate reduction by zerovalent iron: Effects of formate, oxalate, citrate, chloride, sulfate, borate, and phosphate. Environ. Sci. Technol. 2004, 38(9), 2715-2720. (25) Li, M.; Feng, C.; Zhang, Z.; Yang, S.; Sugiura, N. Treatment of nitrate contaminated water using an electrochemical method. Bioresource Technol. 2010, 101(16), 6553-6557. (26) Soares, O. S. G. P.; Orfao, J. J. M.; Pereira, M. F. R. Bimetallic catalysts supported on activated carbon for the nitrate reduction in water: Optimization of catalysts composition. Appl. Catal. B-Environ. 2009, 91(1-2), 441-448. (27) Chaplin, B. P.; Roundy, E.; Guy, K. A.; Shapley, J. R.; Werth, C. J. Effects of natural water ions and humic acid on catalytic nitrate reduction kinetics using an alumina supported Pd-Cu catalyst. Environ. Sci. Technol. 2006, 40(9), 3075-3081. (28) Marchesini, F. A.; Irusta, S.; Querini, C.; Miro, E. Nitrate hydrogenation over Pt,In/Al2O3 and Pt,In/SiO2. Effect of aqueous media and catalyst surface properties upon the catalytic activity. Catal. Commun. 2008, 9(6), 1021-1026. (29) Yamauchi, M.; Abe, R.; Tsukuda, T.; Kato, K.; Takata, M. Highly Selective Ammonia Synthesis from Nitrate with Photocatalytically Generated Hydrogen on CuPd/TiO2. J. Am. Chem. Soc. 2011, 133(5), 1150-1152. (30) Xu, Z.; Chen, L.; Shao, Y.; Yin, D.; Zheng, S. Catalytic Hydrogenation of Aqueous Nitrate over Pd-Cu/ZrO2 Catalysts. Ind. Eng. Chem. Res. 2009, 48(18), 8356-8363. (31) Epron, F.; Gauthard, F.; Barbier, J. Catalytic reduction of nitrate in water on a monometallic Pd/CeO2 catalyst. J. Catal. 2002, 206(2), 363-367. (32) Motamedi, E.; Atouei, M. T.; Kassaee, M. Z. Comparison of nitrate removal from water via graphene oxide coated Fe, Ni and Co nanoparticles. Mater. Res. Bull. 2014, 54, 34-40. (33) Baikousi, M.; Georgiou, Y.; Daikopoulos, C.; Bourlinos, A. B.; Filip, J.; Zboril, R.; Deligiannakis, Y.; Karakassides, M. A. Synthesis and characterization of robust zero valent iron/mesoporous carbon composites and their applications in arsenic removal. Carbon 2015, 93, 636-647. (34) Jabeen, H.; Chandra, V.; Jung, S.; Lee, J. W.; Kim, K. S.; Bin Kim, S. Enhanced Cr(VI) removal using iron nanoparticle decorated graphene. Nanoscale 2011, 3(9), 3583-3585. (35) Phenrat, T.; Saleh, N.; Sirk, K.; Tilton, R. D.; Lowry, G. V. Aggregation and sedimentation of aqueous nanoscale zerovalent iron dispersions. Environ. Sci. Technol. 2007, 41(1), 284-290. (36) Wan, Y.; Shi, Y.; Zhao, D. Supramolecular aggregates as templates: Ordered mesoporous polymers and carbons. Chem. Mater. 2008, 20(3), 932-945. (37) Wu, Z.; Li, W.; Webley, P. A.; Zhao, D. General and Controllable Synthesis of Novel Mesoporous

ACS Paragon Plus Environment

Environmental Science & Technology

442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463

Magnetic Iron Oxide@Carbon Encapsulates for Efficient Arsenic Removal. Adv. Mater. 2012, 24(420), 485-491. (38) Teng, W.; Fan, J.; Wang, W.; Bai, N.; Liu, R.; Liu, Y.; Deng, Y.; Kong, B.; Yang, J.; Zhao, D.; Zhang, W. Nanoscale zero-valent iron in mesoporous carbon (nZVI@C): stable nanoparticles for metal extraction and catalysis. J. Mater. Chem. A 2017, 5(9), 4478-4485. (39) Teng, W.; Wu, Z.; Fan, J.; Chen, H.; Feng, D.; Lv, Y.; Wang, J.; Asiri, A. M.; Zhao, D. Ordered mesoporous carbons and their corresponding column for highly efficient removal of microcystin-LR. Energ. Environ. Sci. 2013, 6(9), 2765-2776. (40) Liu, R.; Shi, Y.; Wan, Y.; Meng, Y.; Zhang, F.; Gu, D.; Chen, Z.; Tu, B.; Zhao, D. Triconstituent Co-assembly to ordered mesostructured polymer-silica and carbon-silica nanocomposites and large-pore mesoporous carbons with high surface areas. J. Am. Chem. Soc. 2006, 128(35), 11652-11662. (41) Chaplin, B. P.; Shapley, J. R.; Werth, C. J. The Selectivity and Sustainability of a Pd-In/gamma-Al2O3 Catalyst in a Packed-Bed Reactor: The Effect of Solution Composition. Catal. Lett. 2009, 130(1-2), 56-62. (42) Third, K. A.; Newland, M.; Cord-Ruwisch, R. The effect of dissolved oxygen on PHB accumulation in activated sludge cultures. Biotechnol. Bioeng. 2003, 82(2), 238-250. (43) Mook, W. T.; Chakrabarti, M. H.; Aroua, M. K.; Khan, G. M. A.; Ali, B. S.; Islam, M. S.; Abu Hassan, M. A. Removal of total ammonia nitrogen (TAN), nitrate and total organic carbon (TOC) from aquaculture wastewater using electrochemical technology: A review. Desalination 2012, 285(Supplement C), 1-13. (44) Li, M.; Feng, C.; Zhang, Z.; Lei, X.; Chen, R.; Yang, Y.; Sugiura, N. Simultaneous reduction of nitrate and oxidation of by-products using electrochemical method. J. Hazard. Mater. 2009, 171(1), 724-730. (45) Ghafari, S.; Hasan, M.; Aroua, M. K. Bio-electrochemical removal of nitrate from water and wastewater— A review. Bioresource Technol. 2008, 99(10), 3965-3974.

464 465

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21

Environmental Science & Technology

466 467

Scheme 1. Mechanism of nitrate reduction by electrocatalysis on nanoscale zero-valent iron

468

supported on ordered mesoporous carbon (nZVI@OMC).

469

ACS Paragon Plus Environment

Environmental Science & Technology

470 471

Figure 1. Illustration of synthetic route for the preparation of nZVI@OMC nanocomposites

472

(A), TEM images of nZVI@OMC after pyrolysis and reduction at (B) 300°C, (C) 400°C, and

473

(D) 500°C, HRTEM image (E) of nZVI@OMC-400 and wide-angel X-ray diffraction (XRD)

474

patterns of nZVI@OMC (F) at (a) 300, (b) 400, and (c) 500°C, respectively.

475

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21

Environmental Science & Technology

476 477

Figure 2. Removal performance (a) and selectivity for N2 (b) of different materials as

478

cathodes for electrocatalytic reduction of nitrate (50 mg/L of nitrate, 0.02 M NaCl and 24 h).

479

ACS Paragon Plus Environment

Environmental Science & Technology

480 481 482

Figure 3. Product distributions of nitrate reduction (a) and selectivity for N2 (b) with the

483

reaction time (50 mg/L of nitrate, 0.02 M NaCl).

ACS Paragon Plus Environment

Page 20 of 21

Page 21 of 21

Environmental Science & Technology

484 485

Figure 4. The nitrate concentrations before and after reactions and removal capacity at

486

different initial concentration (a, 0.02 M NaCl and 24 h), and capacity and selectivity for N2

487

of consecutive cycles on nanocomposite nZVI@OMC-400 (b, 50 mg/L of nitrate, 0.02 M

488

NaCl and 24 h).

489 490

ACS Paragon Plus Environment