Tuning the K+ - ACS Publications - American Chemical Society

Jul 3, 2018 - concentration were measured by an ozone detector (model. 202, 2B Technologies). .... peaks value at 3430 and 1630 cm. −1. , attributed...
1 downloads 0 Views 4MB Size
Subscriber access provided by Kaohsiung Medical University

Remediation and Control Technologies +

2

Tuning the K concentration in the tunnels of #-MnO to increase the content of oxygen vacancy for ozone elimination Guoxiang Zhu, Jinguo Zhu, Wenlu Li, Wenqing Yao, Ruilong Zong, Yongfa Zhu, and Qianfan Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b01594 • Publication Date (Web): 03 Jul 2018 Downloaded from http://pubs.acs.org on July 3, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a 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 29

Environmental Science & Technology

1

Tuning the K+ concentration in the tunnels of α-MnO2 to increase the content of oxygen

2

vacancy for ozone elimination

3

Guoxiang Zhua, Jinguo Zhub, Wenlu Lia, Wenqing Yaoa, Ruilong Zonga, Yongfa Zhua* and Qianfan Zhangb*

4 5

a Department of Chemistry, Tsinghua University, Beijing 100084, China

6

b School of Materials Science and Engineering, Beihang University, Beijing 100191, China

7

1

ACS Paragon Plus Environment

Environmental Science & Technology

8

ABSTRACT

9

α-MnO2 is a promising material for ozone catalytic decomposition and the oxygen vacancy is

10

often regarded as the active site for ozone adsorption and decomposition. Here, α-MnO2 nanowire

11

with tunable K+ concentration was prepared through a hydrothermal process in KOH solution. High

12

concentration K+ in the tunnel can expand crystal cell and break the charge balance, leading to a

13

lower average oxidation state (AOS) of Mn, which means abundant oxygen vacancy. DFT

14

calculation has also proven that the samples with higher K+ concentration exhibit lower formation

15

energy for oxygen vacancy. Due to the enormous active oxygen vacancies existing in the α-MnO2

16

nanowire, the lifetime of the catalyst (corresponding to 100 % ozone removal rate, 25 ºC) is

17

increased from 3 h to 15 h. The FT-IR results confirmed that the accumulation of intermediate

18

oxygen species on the catalyst surface is the main reason why it is deactivated after long time

19

reaction. In this work, the performance of the catalyst has been improved because the abundant

20

active oxygen vacancies are fabricated by the electrostatic interaction between oxygen atoms inside

21

the tunnels and the introduced K+, which offers us a new perspective to design a high efficiency

22

catalyst and may promote manganese oxide for practical ozone elimination.

23

2

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

24

Environmental Science & Technology

TOC Art

25 26

3

ACS Paragon Plus Environment

Environmental Science & Technology

27

1. INTRODUCTION

28

In recent years, ozone is given increased attention as a common air pollutant due to the potential

29

impact on public health and our living environment. A long-term exposure to even low level of ozone

30

would cause serious health problems, including cardiopulmonary disease,1 respiratory disease2 and

31

cardiovascular disease.3, 4 Generally, ozone is formed near the ground in the photochemical reaction

32

referring to volatile organic compounds (VOCs), NOx and carbon monoxide.5 However, household

33

electrostatic equipments6 and ultraviolet disinfection equipments7 also cannot be ignored for indoor

34

ozone production. Besides, ozonation and plasma process also cause serious ozone pollution in

35

factory. Apart from the hazard of ozone, the indoor residual ozone also leads to the formation of

36

VOCs by the ozone-surface reaction with building materials and indoor furnishing,8-10 and the

37

produced compounds may form the secondary organic aerosol (SOA).11, 12 Thus, the Occupational

38

Safety and Health Administration (OSHA, US) stipulates the maximum allowable exposure to ozone

39

in the terms of eight-hour is 0.10 ppm.13 The regulations of Chinese indoor air quality standard’

40

(GB/T 18883 - 2002) also require that the indoor ozone concentration should not exceed 0.08 ppm.

41

Therefore, the research on the ozone elimination is significant for environmental protection and

42

public health.

43

Various methods have been applied for ozone removal, such as plants purification,14

44

adsorption,15 thermal decomposition and catalytic decomposition.16-19 Compared with other methods,

45

the catalytic decomposition has been recognized as a promising approach, due to its higher efficiency,

46

safety and lower cost. The catalyst for ozone elimination can be divided into noble metal,20, 21

47

transitional metal oxide16-19,

48

large-scale application, the research about ozone catalytic decomposition mainly focus on

22, 23

and noble-transitional compounds. Considering the cost and

4

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Environmental Science & Technology

49

transitional metal oxide, especially manganese oxide. Jia et al.24 compared the activity of

50

α, β and γ−MnO2 for ozone removal and observed the highest ozone removal rate with α−MnO2.

51

They found that the catalytic performance of manganese oxide strongly depends on the density of

52

oxygen vacancies. In our previous study, surface oxygen vacancy was proved to be the active site for

53

ozone adsorption and further decomposition.19 The formation of oxygen vacancy would change the

54

original charge distribution, resulting in a lower ozone adsorption energy and better catalytic

55

activity.25 For example, Wang et al.26 reported that an cryptomelane-type manganese oxide (OMS-2)

56

catalyst synthesized with MnAc2 as Mn2+ precursor could form abundant Mn3+, leading to a higher

57

catalytic performance. Ma et al.18 also reported that a higher catalytic performance could be obtained

58

in transition mental doped OMS-2, resulting in the abundant oxygen vacancies. Therefore, the

59

enhancement of oxygen vacancy is regarded as an effective strategy to improve the catalytic

60

performance.

61

α−MnO2 is consisted of 2 × 2 edge-share MnO6 octahedral chains, which corner-connected to

62

form 0.46 × 0.46 nm tunnels. Abundant MnO6 edges exposes in the structure, leading to easy release

63

of lattice oxygen.27 Manganese is located in the framework and the guest cation, such as Li+, Na+, K+,

64

NH4+, is filled in the tunnels to stabilized the structure. Because of the unique tunnel structure,

65

various cations, especially potassium ion, can be introduced into the tunnels to tailor its chemical and

66

physical properties. For example, Yuan et al.28 reported that the presence of potassium ions inside the

67

tunnels of α−MnO2 would increase its electronic conductivity. Tseng et al.29 pointed out that

68

magnetic properties of α−MnO2 nanotube is closely connected with the K+ doping concentration.

69

Hou et al.30 also demonstrated that K+-doped OMS-2 nanowire has better catalytic activity for

70

benzene oxidation. Generally, these cations are located in the center of tunnels and the charge on the 5

ACS Paragon Plus Environment

Environmental Science & Technology

71

cations would be balanced by substitution of Mn3+ for some of the Mn4+.31 So, oxygen vacancies are

72

easy to form at the MnO6 edges and a higher ozone removal rate is expected when the K+ entered

73

into the tunnels. Herein, we developed a simply post-processing method to prepare K-rich α-MnO2

74

nanowire for ozone catalytic decomposition. The effect of K+ concentration in the tunnels of α-MnO2

75

on the performance for ozone removal has been studied theoretically and experimentally.

76

2. EXPERIMENTAL SECTION

77

Preparation of K-rich MnO2

78

α-MnO2 nanowires were synthesized by a hydrothermal process using KMnO4 and MnSO4.

79

Specifically, 3.042 g MnSO4·H2O and 1.896 g KMnO4 (Mn7+ : Mn2+ = 2 : 3) were added into 80 mL

80

deionized (DI) water with constant stir for 20 mins to form a suspension. Subsequently, the

81

suspension was transformed into 100 mL Teflon-lined stainless steel autoclave and heated at 150 °C

82

for 12 h. The products were collected by centrifugation, then washed with DI water and dried at

83

80 °C for 10 h. The obtained nanowire was identified as MnO2. The K+ concentration was tuned by a

84

hydrothermal process in KOH solution as illustrated in Figure 1. 1.0 g synthesized nanowire was

85

added in 80 mL KOH solution accompanied with ultrasonic dispersion for 20 mins, and then

86

transformed into 100 mL Teflon-lined stainless steel autoclave and heated at 150 °C for several hours.

87

Then the collected products were washed with DI water until the filtrate was neutral (PH ≈ 7).

88

Finally, the K-rich MnO2 nanowire was obtained after drying at 80 °C for 10 h.

6

ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29

Environmental Science & Technology

89 90

91

Figure 1. Schematic illustration of the preparation of K-rich α-MnO2.

Catalyst Characterization.

92

X-ray diffraction (XRD) patterns of the samples were obtained by a Rigaku D/max-2400 X-ray

93

diffractometer with Cu Kα (λ = 1. 5406 Å) radiation at 40 kV and 150 mA. Morphologies of the

94

samples were captured by a Field Emission Gun Scanning Electron Microscopy (FESEM, Hitachi

95

SU-8010) and a transmission electron microscopy (TEM, Hitachi 7700) with an accelerating voltage

96

of 100 kV. High resolution transmission electron microscopy (HRTEM) images were obtained by a

97

JEM 2100F field emission transmission electron microscope at an accelerating voltage of 200 kV.

98

The element composition and distribution were recorded by ICP-MS (iCAPQ, Thermo Fisher

99

SCientific) and an energy dispersive (EDS) detector in Hitachi SU-8010 and JEM 2100F. X-ray

100

photoelectron spectroscopy (XPS) was conducted in a PHI Quantera SXMTM system and the binding

101

energy was calibrated with the signal for adventitious carbon at 284.8 eV. Fourier transform infrared

102

(FT-IR) spectra and in situ diffused Fourier transform infrared spectroscopy (DFTIR) was recorded

103

by Bruker V70 spectrometer. Thermogravimetric analysis (TGA) was carried out in a TGA/DSC1

104

STARe system (METTLER TOLEDO). Temperature programmed desorption (TPD) was carried out

105

on a Cat-Lab (BEL Japen, Inc.) eqiupped with an online QIC-200 quadrupole mass (Inprocess

106

Instruments, GAM 200) as a detector.

107

Catalytic Activity. 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 29

108

The performance of catalytic ozone decomposition was evaluated in a continuous fixed bed

109

reactor (11 mm inner diameter) at room temperature (25 ºC). 0.1 g catalyst was used for each test

110

with the space velocity of 5.4 × 105 mL·g-1·h-1. The total gas flow through the reactor was kept at

111

900 mL/min and the inlet ozone concentration maintained at 50 ± 1 ppm. The relative humidity was

112

measured by a hygronom (605-H1 Testo) and controlled by tuning the gas flow through a humidifier

113

without changing the total gas flowrate. An ozonator (model 1000BT-12, Shanghai Enaly

114

Mechanical and Electrical Technology Company) was applied to generate ozone by arc discharge in

115

O2 stream, where the O2 flow and the discharge voltage can be tuned. The generated ozone mixed

116

adequately with air in a mixing drum and then transported to the reactor. The inlet and outlet ozone

117

concentration were measured by an ozone detector (model 202, 2B Technologies). The ozone

118

removal rate can be calculated from the following equation:

119

Ozone removal rate = 100 % × (Cin - Cout)/Cin.

120

where Cin and Cout are inlet and outlet ozone concentration respectively.

121

3. RESULTS AND DISCUSSION

122

3.1 The Characterization of Structure and Morphology.

123 8

ACS Paragon Plus Environment

(1)

Page 9 of 29

Environmental Science & Technology

124

Figure 2. XRD patterns of α-MnO2 before and after treated in 1M KOH solution (a); the enlarged

125

view of the α-MnO2 (200) peak (b).

126

The structure of α-MnO2 nanowire before and after the treatment in KOH solution were

127

analyzed by XRD as shown in Figure 2a. All the obtained samples have the same peaks,

128

corresponding to a pure tetragonal cryptomelane type MnO2 (JCPDS No. 29-1020), which means the

129

crystalline phase is well maintained during the treatment.32 In order to confirm the influence of the

130

post-processing on the composition of the samples, the bulk contents of K and Mn were detected by

131

ICP-MS (Table 1). After hydrothermal treatment in 1 M KOH solution for 4 h, the K/Mn molar ratio

132

is increased from 13.44 % to 20.75 %. The K/Mn molar ratio on the catalyst surface has also been

133

measured by XPS and the treatment can increase the ratio from 13.46 % to 22.89 % compared with

134

pure MnO2, as shown in Figure S3. The same results can also be obtained by the EDS equipped in

135

FESEM and HRTEM, indicating that the K+ content has been indeed enhanced.

136

In the α-MnO2 crystal structure, the crystal radii of K+ (1.65 Å) is significantly larger than that

137

of Mn3+ (high spin, 0.785 Å; low spin, 0.72 Å) and Mn4+ (0.67 Å), suggesting that K+ entered into

138

the 2 × 2 tunnels of α-MnO2 rather than the octahedral framework.18 Further observation of XRD

139

spectra shows that all the peaks gradually shift to lower angles with the extension of hydrothermal

140

time and the trend is more clearly for (200) peak (Figure 2b), which means the tunnel is expanded.28,

141

29, 33

142

maintained a pure cryptomelane structure and possess of high crystallinity. However, the lattice

143

spacing of (200) plane has increased from 0.476 nm to 0.489 nm after treated, corresponding to the

144

expansion of the tunnel structure. In addition, the mapping images (Figure S4) suggest that the K and

145

Mn is well distributed in the nanowire before and after hydrothermal treatment. Therefore, it can be

High-resolution TEM (HRTEM) images (Figure 3c, f) indicate that MnO2 and KOH-4h

9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 29

146

concluded that abundant K+ were entered into the tunnel structure and well dispersed in the

147

hydrothermal process.

148

Table 1. Atomic ratio of K/Mn, BET surface areas, and AOS of Mn of the manganese oxide samples. Atomic ratio of K/Mn Sample

BET surface areas AOS of Mn

ICP-MS

FESEM-EDS

[m2/g]

MnO2

13.44

11.45

61.00

3.76

KOH-2h

19.24

16.19

52.50

3.49

KOH-4h

20.75

16.93

52.43

3.43

KOH-6h

23.18

17.16

51.94

3.37

149

The morphologies of as-prepared samples were investigated by FESEM and TEM. As shown in

150

Figure 3a, b, the as-synthesized MnO2 presents a typical nanowire morphology with a diameter in the

151

range of 20-40 nm and a length for about 2-5 µm. However, after hydrothermal treatment in 1 M

152

KOH solution for 4 h, the original isolated nanowire arranges along the diameter direction and forms

153

a bunchy structure. TEM images (Figure 3d, e) also confirmed that the length of KOH-4h bundle has

154

no obvious change but the width is beyond 60 nm. As shown in FT-IR spectra (Figure 4a), the peaks

155

value at 3430 cm-1 and 1630 cm-1, attributed to the surface absorbed water and hydroxyl groups

156

respectively,34 become larger with longer treatment time, suggesting abundant surface hydroxyl

157

groups is formed during the post-processing. Hu et al.35 pointed out that abundant surface hydroxyl

158

groups are produced from the hydroxylation of the surface oxides to balance the overall charge after

159

K+ entered into the tunnels of α-MnO2 nanowire. Therefore, the existing of abundant surface

160

hydroxyl groups means a higher K+ content in the tunnels. From Figure S5, it can also be found that

161

the bundled trend become serious with increased hydrothermal time. Table 1 also has shown that the

162

treatment in KOH solution leads to the decrease of surface area from 61 m2/g to 51.91 m2/g, 10

ACS Paragon Plus Environment

Page 11 of 29

Environmental Science & Technology

163

corresponding to the change in morphology. Besides, the bundled extent of treated samples reduces

164

partly after calcination at 300 ºC for 2 h, while the crystal structure has no obvious difference (Figure

165

S7, 8). Hence, it can be concluded that the bunchy structure is formed with the help of the abundant

166

surface hydroxyl groups, which induce hydrogen bond interaction among isolated nanowires. The

167

bunchy structure also reconfirms the enhancement of K+ concentration in the treated samples.

168 169

Figure 3. FESEM images of α-MnO2 (a) and KOH-4h (d); TEM images of α-MnO2 (b) and

170

KOH-4h (e); HRTEM images of α-MnO2 (c) and KOH-4h (f).

171

The evolution of oxygen from the samples was investigated by temperature programmed

172

desorption (TPD-MS) as shown in Figure 4b. The desorption process can be divided into four parts.

173

The peaks lower than 200 ºC corresponds to the physical absorbed water molecule.36 The peak

174

between 200 ºC and 400 ºC is assigned as the desorption of surface hydroxyl radical. For KOH-4h,

175

the peak at 233 ºC shift to 278 ºC, suggesting the existence of a stronger bond between surface 11

ACS Paragon Plus Environment

Environmental Science & Technology

176

hydroxyl radical and KOH-4h. The peaks between 400 ºC to 650 ºC is assigned as the release of the

177

lattice oxygen close to the surface19 and the crystalline phase will completely transform to Mn2O3

178

while the temperature rises to 800 ºC. Compared with MnO2, the peaks of KOH-4h at 475 ºC and

179

586 ºC shift to 531 ºC and 617 ºC respectively and the peak area has an obvious decrease, indicating

180

that a large amount of lattice oxygen is more stable after treated in KOH solution.33 As shown in

181

Figure 4c, the TGA curve under N2 atmosphere has been recorded to characterize the thermostability.

182

After the treatment in KOH solution, the weight loss from 45 ºC to 800 ºC is reduced from 11.65 %

183

to 9.97 %, as a result of the higher K+ content in KOH-4h. Interestingly, the weight loss between 400

184

ºC to 650 ºC decreases form 5.72 % to 3.26 % while the ratio between 650 ºC to 800 ºC increases

185

from 3.44 % to 4.53 %, corresponding to the TPD-MS results, suggesting a more stable structure is

186

indeed induced by increasing K+ concentration in tunnels. Based on the elaborate analysis of the

187

crystal structure of cryptomelane type MnO2 (Figure S9), it can be found that the increase of K+

188

concentration in the tunnels can induce the electrostatic interaction between K+ and the lattice

189

oxygen with a sp3 hybridization [O (sp3)] to form a balanced force distribution, which leads to a

190

higher temperature for the evolution of oxygen.

191

Considering the vital impact of Mn-O bond on the catalytic activity, the average oxidation state

192

(AOS) of surface Mn atom was analyzed via XPS. As shown in Figure S11, the signals of Mn, O and

193

K can be found in all the samples. After treated in KOH solution, the charge on the tunnel cation (K+)

194

is balanced by substitution of Mn3+ for some of the Mn4+ in the framework.31 Further observation

195

(Figure 4d) has shown that the binding energy difference (∆‫ܧ‬௦ ) between the two peaks of Mn 3s

196

varies with the increase of K+ content and the AOS of surface Mn has been estimated via the

197

following formula37,

38

: AOS=8.956-1.126∆‫ܧ‬௦ . After treated in KOH solution, the AOS of Mn 12

ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29

Environmental Science & Technology

198

gradually decreases from 3.76 to 3.37 with the extension of treatment time, meaning a lower average

199

coordination number and higher oxygen vacancy content. In addition, the decrease of AOS means

200

that Jahn-Teller (JT) distortion can occur to lower the system’s ground state37, which has the same

201

conclusion with XRD results.

202 203

Figure 4. The FTIR spectra (a), the TPD-MS profiles (b), TGA curve (c) and Mn 3s XPS spectra (d)

204

of the samples.

205

3.2 Performance of Ozone Removal

206 13

ACS Paragon Plus Environment

Environmental Science & Technology

207 208

Figure 5. The ozone removal rate of the samples treated with different concentration KOH solution

209

(a) and the samples treated for different time in 1M KOH solution (b); The ozone removal rate of the

210

samples treated with different alkali solution for 4h (c); The ozone removal rate of the samples

211

treated with different potassium salt solution for 4h (d); The ozone removal rate on KOH-4h at

212

different RH (e); The ozone removal rate on KOH-4h at alternate humidity condition (f); All the

213

reaction was carried out at 25 ºC.

214

Based on the characterized analysis, K+ entered into the tunnels of α-MnO2 with the treatment,

215

which can generate abundant oxygen vacancy for ozone decomposition. Therefore, a higher ozone

216

removal rate is expected for α-MnO2 with higher K+ content. As shown in Figure 5a, the ozone

217

removal rate of the samples treated in different concentration KOH solution have been evaluated. For

218

α-MnO2 without treatment, the ozone removal rate begins to decrease while the reaction runs for 3 h

219

in the gas flow with 50 ppm ozone and it decreases to 25 % when the reaction maintained for 12 h.

220

The bunchy structure in the treated samples, with a smaller surface area, provides a poor activity if 14

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

Environmental Science & Technology

221

the density of active site has no obvious increase. However, when the as-synthesized nanowire was

222

treated in 0.25 M KOH solution for 4 h, the time possessed of 100 % ozone removal rate increased

223

from 3 h to 10 h and it could extend to 15 h if the concentration of KOH solution keeps at 1.00 M. It

224

indicates that the treatment in KOH solution has indeed improved the catalytic performance for

225

ozone removal. The atomic K/Mn ratio detected by EDS detector equipped in FESEM also shows

226

there is a positive correlation between the catalytic performance and the K+ content, suggesting that

227

the increase of K+ content in the tunnels is beneficial for ozone decomposition. The performance of

228

the catalyst is also affected by hydrothermal time. The optimal catalytic activity can be obtained

229

when α-MnO2 nanowire is treated in 1.00 M KOH solution for 4 h (Figure 5b). However, if the

230

hydrothermal time or the concentration of KOH solution continuously increased, the increase of K+

231

content will be slow (Figure S12a, b) and the effect of bunchy structure on the surface area will be

232

obvious, leading to an almost unchanged catalytic activity.

233

In order to further confirm the reason why the ozone removal rate has been improved, the

234

α-MnO2 nanowire was treated in different alkali and kali salt solution respectively. As can be seen in

235

Figure 5c, the samples treated in LiOH, NaOH or NH3·H2O solution are all possessed of a lower

236

performance compared with untreated MnO2, indicating that it is K+ content in the tunnels rather

237

than surface hydroxyl group that improves the ozone removal rate. When the α-MnO2 nanowire was

238

treated in KNO3, K2SO4 and CH3COOK solution respectively (the K+ concentration keeps at 1M),

239

the activity for ozone removal has different levels of increase except for the sample treated in KNO3

240

solution (seeing Figure 5d), suggesting ozone decomposition is also effected by the anion absorbed

241

on the surface. High-resolution XPS spectra of KNO3-4h, K2SO4-4h, CH3COOK-4h also has

242

confirmed the existence of absorbed anions on the surface of the samples (Figure S14). Besides, the 15

ACS Paragon Plus Environment

Environmental Science & Technology

243

atomic K/Mn ratio has been further analyzed by the EDS detector, as shown in Figure S12c. The

244

samples treated in different kali salt solution are possessed of various K+ content (Figure S12d),

245

resulting in different catalytic performance for ozone removal. It can be concluded that the anion in

246

the solution will affect the ability of K+ entering the α-MnO2 tunnel structure, leading to various

247

content of oxygen vacancy. In general, the ozone catalytic decomposition can be facilitated by

248

increasing the K+ content in the tunnels and it is also influenced by the corresponding anion in the

249

solution.

250

The catalyst is easy to be deactivated in humid environment.17, 39 As can be seen in Figure 5e,

251

when the relative humidity (RH, 25 °C) increases from 22 % to 30 %, the ozone removal rate at 15 h

252

decreases from 99 % to 78 % and the removal rate continues to decreasing with further increase of

253

RH, suggesting that the performance of the catalysts is also sensitive to the humidity of the gas flow.

254

In the high humidity gas flow, a large amount of surface hydroxyl group would form under the

255

combined action of water vapor and ozone molecule. The hydrogen bond interaction between surface

256

hydroxyl group and water molecule would adverse to the ozone uptake on the catalyst.40 For

257

KOH-4h, due to the abundant hydroxyl group on the surface, a strong hydrogen bond interaction has

258

already exist in the initial stage of the reaction, resulting in a dramatically decrease of ozone removal

259

rate at the beginning (Figure S16). For untreated MnO2, the inherent surface hydroxyl group is finite,

260

so the ozone removal rate decreased gradually. However, KOH-4h is possessed of higher activity at

261

12 h than MnO2, which suggests more active sites existing in KOH-4h. As can be seen in Figure 5f,

262

the ozone removal rate has been evaluated at alternate humidity conditions. When the catalyst was

263

placed into the wet gas flow (50 % RH), the ozone removal rate has a sharp decrease due to the

264

competitive adsorption of water molecule. However, once the relative humidity was decreased, the 16

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29

Environmental Science & Technology

265

ozone removal rate recover to 100 % quickly, indicating that the deactivation resulting from water

266

molecule is temporary and the catalytic activity in the dry gas flow has not been affected by surface

267

hydroxyl group. When the relative humidity of the gas flow was increased to 50 % again, the ozone

268

removal rate decreases to the level of the latest cycle quickly and then continues at a slower rate,

269

suggesting the increase of the surface hydroxyl group can exacerbate the competitive adsorption

270

which leads to a lower performance for ozone removal.

271

3.3 The Formation Energy of Oxygen Vacancy

272 273

Figure 6. The calculation results of oxygen vacancy formation energies at specific sites with varied

274

K content.

275

The process of oxygen vacancy formation often plays a key role in catalytic activity.41, 42 In

276

order to deeply understand the effect of K+ in the tunnels on the formation of oxygen vacancy, the

277

first-principles calculations are performed based on density functional theory as implemented in the

278

VASP code. The elaborated calculation method can be found in the Supporting Information. As

279

shown in Figure S9, in the framework of α-MnO2, each oxygen atom is coordinate with three 17

ACS Paragon Plus Environment

Environmental Science & Technology

280

manganese atoms.

281

three nearby Mn atoms to form the sp2 hybridization [O (sp2)], while other half locate out of plane to

282

form the sp3 hybridization [O (sp3)] with nearest manganese atoms. The oxygen vacancy formation

283

energy (Eform) at two specified sites (inset of Figure 6), with sp2 and sp3 hybridization respectively,

284

has been simulated in the absence or presence of different K+ concentration in the tunnels to present

285

us a tendency of oxygen vacancy formation energies. As shown in Figure 6, the oxygen vacancy

286

formation energy at the sp2 site dramatically decreases from 3.001 eV to 2.018 eV after 1 K+ is

287

constructed in the tunnel structure, and the formation energy keeps decreasing with the increase of

288

K+ concentration, corresponding to Hou’s results.30 The results indicate that the filled K+ can lead to

289

the electrostatic interaction between K+ and the lattice oxygen which is beneficial to the formation of

290

active oxygen vacancy. Besides, the effect of K+ is closely related to the distance between K+ and the

291

oxygen site. However, different results were obtained at the specific sp3 oxygen site. Without K+ in

292

the tunnels, the sp3 oxygen atom is faced with an asymmetrical force distribution which exhibits

293

lower Eform. However, when two K+ is constructed in the neighborhood of the specified oxygen atom,

294

a symmetrical force distribution forms, leading to a significant increase of oxygen vacancy formation

295

energies at sp3 oxygen atoms. So, the Eform first decreases and then increases with the growth of K+

296

content. Overall, the formation energy of oxygen vacancy is smaller at sp3 oxygen site, indicating

297

that the oxygen vacancy preferentially forms at sp3 oxygen site, corresponding to our former works.19

298

Therefore, tuning the K+ concentration in the tunnels appropriately will be an effective method to

299

enhance the content of oxygen vacancy.

300

3.4 The Schematic of Ozone Decomposition and Catalyst Regeneration.

Among them, half of the oxygen atom locate roughly in the same plane with the

18

ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29

Environmental Science & Technology

301 302

Figure 7. The schematic of ozone catalytic decomposition on the K-rich MnO2 (a); FT-IR spectra of

303

KOH-4h after reaction for different time (b); FT-IR spectra of MnO2 after reaction in different gas

304

flow (c); The TPD-MS profiles of KOH-4h before and after reaction (d); In situ DFTIR spectra of the

305

sample after reaction (e); High-resolution XPS spectra of KOH-4h before and after reaction (f).

306

The oxygen vacancy mechanism24, 43, 44 is generally accepted and our previous works also have

307

elaborated the effect of oxygen vacancy on the ozone decomposition.19 Firstly, one of the oxygen

308

atoms in the ozone molecule inserts into the surface oxygen vacancy and electrons transform from

309

Mn atom to ozone molecule, leading to the decomposition of ozone molecule into oxygen molecule

310

and oxygen species (O2-). Then, another ozone molecule reacts with the oxygen species and forms an

311

oxygen molecule and peroxide species (O22-). Finally, the peroxide species transforms into oxygen

312

molecule and the oxygen vacancy recovers. As shown in Figure 7a, the hydrothermal treatment in

313

KOH solution will induce the formation of oxygen vacancy. The more oxygen vacancy form, the

314

more active sites for ozone decomposition exist, resulting in a higher performance for ozone 19

ACS Paragon Plus Environment

Environmental Science & Technology

315

removal.

316 317

Figure 8. The FT-IR spectra of KOH-4h before reaction, KOH-4h after reaction and regenerated

318

KOH-4h (a); The ozone removal rate of the KOH-4h and regenerated KOH-4h (b).

319

However, the catalytic performance decreases with the reaction on, so the reason why the

320

catalyst deactivates is important for designing more effective catalyst. As can be seen in Figure 7b,

321

only the peaks around 3430 cm-1 and 1630 cm-1 can be found in the FT-IR spectra of fresh KOH-4h.

322

However, a new peak appears at 1380 cm-1 after the catalytic reaction, suggesting the surface

323

composition has changed in the process.45, 46 And the intensity of that peak increases with the extend

324

of the reaction time, indicating the peak at 1380 cm-1 results from the accumulation of intermediate

325

species. To further confirm the intermediate species, pure O2 was taken as the source gas to evaluate

326

the activity of KOH-4h and the FT-IR spectra of treated KOH-4h was shown in Figure 7c. The peak

327

appears whether the source gas is O2 or air, indicating that the intermediate species belongs to

328

surface oxygen species. The results of TPD-MS (Figure 7d) show that an oxygen desorption peaks

329

appear at 364 ºC, suggesting the intermediate species indeed belongs to oxygen species. To further

330

understand the thermostability of the intermediate oxygen species, in situ DFTIR has been carried

331

out. As shown in Figure 7e, the peak at 1380 cm-1 always exist even the temperature increased to 200 20

ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29

Environmental Science & Technology

332

ºC, suggesting that the intermediate oxygen species is relatively stable. However, Jia et al.24 reported

333

that the peroxide significantly decomposes at temperature less than 100 ºC from Raman spectrum.24

334

So, we speculate that the peroxide species (O22-)43, 47, 48 would evolve into a relatively stable oxygen

335

species, corresponding to the peak at 1380 cm-1, leading to the deactivation of the catalyst. The same

336

results can be obtained through the increase of the content of surface adsorbed oxygen (Figure S18)

337

and the AOS of Mn atom (Figure 7f). Therefore, the higher performance of KOH-4h results from the

338

abundant oxygen vacancy induced by K+, but the accumulation of intermediate oxygen species

339

causes the inactivation of the catalyst.

340

To reuse the catalyst, the deactivated KOH-4h has been calcinated in Ar atmosphere at 400 ºC

341

for 2 h to remove the accumulation oxygen species. As shown in figure 8a, the peak at 1380 cm-1

342

disappear after the deactivated KOH-4h has been calcinated. At the same times, the performance also

343

recovers partly (figure 8b), indicating that calcination in Ar atmosphere is an effective method to

344

regenerate the catalyst.

345 346

3.5 Implications and Future Direction of the Research.

347

In this work, the K+ content in the tunnel structure of α-MnO2 nanowire is increased dramatically

348

via a simply post-processing with KOH solution, generating abundant oxygen vacancy which leads

349

to a higher performance for ozone removal. Theoretical results show that the increase of K+ content

350

in the tunnels can decrease the formation energy of oxygen vacancy, which is beneficial for the

351

formation of oxygen vacancy in the crystal lattice. Experimental data also shows the AOS of Mn

352

decreases as the K+ content in the tunnels increases, meaning the formation of abundant oxygen

353

vacancy. The existence of active oxygen vacancy accelerates the decomposition of ozone. The K+ 21

ACS Paragon Plus Environment

Environmental Science & Technology

354

content can be controlled by adjusting KOH concentration and hydrothermal conditions. When the

355

as-synthesized α-MnO2 nanowire is treated in 1 M KOH solution for 4 h, the time possessed of 100 %

356

ozone removal rate at high space velocity (5.4 × 105 mL·g-1·h-1) can be increased from 3 h to 15 h.

357

In addition, the deactivation in high humidity is temporary and the activity would recover once the

358

humidity decrease. So, a high performance for ozone decomposition can be obtained by adding a

359

dehumidifier before the reactor. More importantly, the deactivated catalyst can be regenerated by

360

calcination in Ar atmosphere, which is significant for practical application.

361

Here, abundant active surface oxygen vacancy was fabricated by the electrostatic interaction

362

between oxygen atoms inside of the tunnels and introduced K+, which offer us a new perspective to

363

improve the catalytic performance for ozone removal. However, the results of FT-IR spectra show

364

peroxide species (O22-) can accumulate on the surface of the catalyst and evolve into a relatively

365

stable oxygen species, causing a decrease of the ozone removal rate. Therefore, the accumulation of

366

intermediate oxygen species is a key factor limiting ozone decomposition efficiently and sustainably

367

and need to be further researched.

22

ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29

Environmental Science & Technology

368

4. ASSOCIATED CONTENT

369

Supporting Information Available. Materials characterization methods, DFT calculation method

370

and other supplementary date are provided in the Supporting Information. The detail information

371

about the calculation in this study can be provided by contacting the corresponding author Qianfan

372

Zhang ([email protected]). These materials are available free of charge via the Internet at

373

http://pubs.acs.org.

374 375

AUTHOR INFORMATION

376

Corresponding Author

377

a* Email: [email protected]

378

b* Email: [email protected]

379

Note

380

The authors declare no competing financial interest.

381 382

ACKNOWLEDGEMENTS

383

This work was partly supported by Chinese National Science Foundation (21437003, 21673126,

384

21621003, 21761142017) and Collaborative Innovation Center for Regional Environmental Quality.

385

23

ACS Paragon Plus Environment

Environmental Science & Technology

386

REFERENCE

387

1.

388

Short-term effects of various ozone metrics on cardiopulmonary function in chronic obstructive pulmonary disease

389

patients: Results from a panel study in Beijing, China. Environ. Pollut. 2018, 232, 358-366.

390

2.

391

low-pollution region. J. Occup. Environ. Med. 2017, 59, (7), 624-630.

392

3.

393

White, R. H.; Curriero, F. C., Health benefits from large-scale ozone reduction in the United States. Environ. Health

394

Perspect. 2012, 120, (10), 1404-1410.

395

4.

396

A.; Schwartz, J.; Mittleman, M.; Stone, P.; Horton, E.; Gold, D. R., Opposing effects of particle pollution, ozone,

397

and ambient temperature on arterial blood pressure. Environ. Health Perspect. 2012, 120, (2), 241-246.

398

5.

399

Johnson, B. J.; Tarasick, D.; Leblanc, T.; McDermid, I. S.; Jaffe, D.; Gao, R.; Stith, J.; Ryerson, T.; Aikin, K.;

400

Campos, T.; Weinheimer, A.; Avery, M. A., Increasing springtime ozone mixing ratios in the free troposphere over

401

western North America. Nature 2010, 463, 344.

402

6.

403

office equipment: A review of reported data and information needs. Atmos. Environ. 2008, 42, (7), 1371-1388.

404

7.

405

129-166.

406

8.

407

from carpets. Atmos. Environ. 2017, 148, 42-48.

Li, H.; Wu, S.; Pan, L.; Xu, J.; Shan, J.; Yang, X.; Dong, W.; Deng, F.; Chen, Y.; Shima, M.; Guo, X.,

Magzamen, S.; Moore, B. F.; Yost, M. G.; Fenske, R. A.; Karr, C. J., Ozone-related respiratory morbidity in a

Berman, J. D.; Fann, N.; Hollingsworth, J. W.; Pinkerton, K. E.; Rom, W. N.; Szema, A. M.; Breysse, P. N.;

Hoffmann, B.; Luttmann-Gibson, H.; Cohen, A.; Zanobetti, A.; de Souza, C.; Foley, C.; Suh, H. H.; Coull, B.

Cooper, O. R.; Parrish, D. D.; Stohl, A.; Trainer, M.; Nédélec, P.; Thouret, V.; Cammas, J. P.; Oltmans, S. J.;

Destaillats, H.; Maddalena, R. L.; Singer, B. C.; Hodgson, A. T.; McKone, T. E., Indoor pollutants emitted by

Dhandapani, B.; Oyama, S. T., Gas phase ozone decomposition catalysts. Appl. Catal., B 1997, 11, (2),

Abbass, O. A.; Sailor, D. J.; Gall, E. T., Effect of fiber material on ozone removal and carbonyl production

24

ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

Environmental Science & Technology

408

9.

409

to the chemical erosion of squalene interfaces during ozonolysis. Environ. Sci. Technol. 2017, 51, (23),

410

13740-13748.

411

10. Weschler, C. J.; Nazaroff, W. W., Growth of organic films on indoor surfaces. Indoor Air 2017, 27, (6),

412

1101-1112.

413

11. Fadeyi, M. O.; Weschler, C. J.; Tham, K. W.; Wu, W. Y.; Sultan, Z. M., Impact of human presence on

414

secondary organic aerosols derived from ozone-initiated chemistry in a simulated office environment. Environ. Sci.

415

Technol. 2013, 47, (8), 3933-3941.

416

12. Waring, M. S.; Siegel, J. A., Indoor secondary organic aerosol formation initiated from reactions between

417

ozone and surface-sorbed d-limonene. Environ. Sci. Technol. 2013, 47, (12), 6341-6348.

418

13. Lian, Z.; Ma, J.; He, H., Decomposition of high-level ozone under high humidity over Mn–Fe catalyst: The

419

influence of iron precursors. Catal. Commun. 2015, 59, 156-160.

420

14. Abbass, O. A.; Sailor, D. J.; Gall, E. T., Effectiveness of indoor plants for passive removal of indoor ozone.

421

Build. Environ. 2017, 119, 62-70.

422

15. Yang, S.; Zhu, Z.; Wei, F.; Yang, X., Carbon nanotubes/activated carbon fiber based air filter media for

423

simultaneous removal of particulate matter and ozone. Build. Environ. 2017, 125, 60-66.

424

16. Gopi, T.; Swetha, G.; Chandra Shekar, S.; Ramakrishna, C.; Saini, B.; Krishna, R.; Rao, P. V. L., Catalytic

425

decomposition of ozone on nanostructured potassium and proton containing δ-MnO2 catalysts. Catal. Commun.

426

2017, 92, 51-55.

427

17. Gong, S.; Chen, J.; Wu, X.; Han, N.; Chen, Y., In-situ synthesis of Cu2O/reduced graphene oxide composite as

428

effective catalyst for ozone decomposition. Catal. Commun. 2018, 106, 25-29.

429

18. Ma, J.; Wang, C.; He, H., Transition metal doped cryptomelane-type manganese oxide catalysts for ozone

Heine, N.; Houle, F. A.; Wilson, K. R., Connecting the elementary reaction pathways of criegee intermediates

25

ACS Paragon Plus Environment

Environmental Science & Technology

430

decomposition. Appl. Catal., B 2017, 201, 503-510.

431

19. Zhu, G.; Zhu, J.; Jiang, W.; Zhang, Z.; Wang, J.; Zhu, Y.; Zhang, Q., Surface oxygen vacancy induced α-MnO2

432

nanofiber for highly efficient ozone elimination. Appl. Catal., B 2017, 209, 729-737.

433

20. Nikolov, P.; Genov, K.; Konova, P.; Milenova, K.; Batakliev, T.; Georgiev, V.; Kumar, N.; Sarker, D. K.;

434

Pishev, D.; Rakovsky, S., Ozone decomposition on Ag/SiO2 and Ag/clinoptilolite catalysts at ambient temperature.

435

J. Hazard. Mater. 2010, 184, (1), 16-19.

436

21. Hao, Z.; Cheng, D.; Guo, Y.; Liang, Y., Supported gold catalysts used for ozone decomposition and

437

simultaneous elimination of ozone and carbon monoxide at ambient temperature. Appl. Catal., B 2001, 33, (3),

438

217-222.

439

22. Tang, W.-X.; Liu, H.-D.; Wu, X.-F.; Chen, Y.-F., Higher oxidation state responsible for ozone decomposition at

440

room temperature over manganese and cobalt oxides: Effect of calcination temperature. Ozone: Sci. Eng. 2014, 36,

441

(5), 502-512.

442

23. Mehandjiev, D.; Naydenov, A.; Ivanov, G., Ozone decomposition, benzene and CO oxidation over

443

NiMnO3-ilmenite and NiMn2O4-spinel catalysts. Appl. Catal., A 2001, 206, (1), 13-18.

444

24. Jia, J.; Zhang, P.; Chen, L., Catalytic decomposition of gaseous ozone over manganese dioxides with different

445

crystal structures. Appl. Catal., B 2016, 189, 210-218.

446

25. Li, L.; Feng, X.; Nie, Y.; Chen, S.; Shi, F.; Xiong, K.; Ding, W.; Qi, X.; Hu, J.; Wei, Z.; Wan, L.-J.; Xia, M.,

447

Insight into the effect of oxygen vacancy concentration on the catalytic performance of MnO2. ACS Catal. 2015, 5,

448

(8), 4825-4832.

449

26. Wang, C.; Ma, J.; Liu, F.; He, H.; Zhang, R., The effects of Mn2+ precursors on the structure and ozone

450

decomposition activity of cryptomelane-type manganese oxide (OMS-2) Catalysts. J. Phys. Chem. C 2015, 119, (40),

451

23119-23126. 26

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

Environmental Science & Technology

452

27. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H. M.; Tadé, M. O.; Wang, S., Different crystallographic

453

one-dimensional MnO2 nanomaterials and their superior performance in catalytic phenol degradation. Environ. Sci.

454

Technol. 2013, 47, (11), 5882-5887.

455

28. Yuan, Y.; Zhan, C.; He, K.; Chen, H.; Yao, W.; Sharifi-Asl, S.; Song, B.; Yang, Z.; Nie, A.; Luo, X.; Wang, H.;

456

Wood, S. M.; Amine, K.; Islam, M. S.; Lu, J.; Shahbazian-Yassar, R., The influence of large cations on the

457

electrochemical properties of tunnel-structured metal oxides. Nat. Commun.2016, 7, 13374.

458

29. Tseng, L.-T.; Lu, Y.; Fan, H. M.; Wang, Y.; Luo, X.; Liu, T.; Munroe, P.; Li, S.; Yi, J., Magnetic properties in

459

α-MnO2 doped with alkaline elements. Sci. Rep. 2015, 5, 9094.

460

30. Hou, J.; Liu, L.; Li, Y.; Mao, M.; Lv, H.; Zhao, X., Tuning the K+ Concentration in the tunnel of OMS-2

461

nanorods leads to a significant enhancement of the catalytic activity for benzene oxidation. Environ. Sci. Technol.

462

2013, 47, (23), 13730-13736.

463

31. Li, L.; King, D. L., Synthesis and characterization of silver hollandite and its application in emission control.

464

Chem. Mater. 2005, 17, (17), 4335-4343.

465

32. Meng, Y.; Song, W.; Huang, H.; Ren, Z.; Chen, S.-Y.; Suib, S. L., Structure–property relationship of

466

bifunctional MnO2 nanostructures: Highly efficient, ultra-stable electrochemical water oxidation and oxygen

467

reduction reaction catalysts identified in alkaline media. J. Am. Chem. Soc. 2014, 136, (32), 11452-11464.

468

33. Luo, J.; Zhu, H. T.; Liang, J. K.; Rao, G. H.; Li, J. B.; Du, Z. M., Tuning magnetic properties of α-MnO2

469

nanotubes by K+ doping. J. Phys. Chem. C 2010, 114, (19), 8782-8786.

470

34. Rong, S.; Zhang, P.; Yang, Y.; Zhu, L.; Wang, J.; Liu, F., MnO2 framework for instantaneous mineralization of

471

carcinogenic airborne formaldehyde at room temperature. ACS Catal. 2017, 7, (2), 1057-1067.

472

35. Hu, B.; Chen, C.-h.; Frueh, S. J.; Jin, L.; Joesten, R.; Suib, S. L., Removal of aqueous phenol by adsorption

473

and oxidation with doped hydrophobic cryptomelane-type manganese oxide (K−OMS-2) nanofibers. J. Phys. Chem. 27

ACS Paragon Plus Environment

Environmental Science & Technology

474

C 2010, 114, (21), 9835-9844.

475

36. Luo, J.; Zhang, Q.; Garcia-Martinez, J.; Suib, S. L., Adsorptive and acidic properties, reversible lattice oxygen

476

evolution, and catalytic mechanism of cryptomelane-type manganese oxides as oxidation catalysts. J. Am. Chem.

477

Soc. 2008, 130, (10), 3198-3207.

478

37. Yang, Y.; Huang, J.; Wang, S.; Deng, S.; Wang, B.; Yu, G., Catalytic removal of gaseous unintentional POPs

479

on manganese oxide octahedral molecular sieves. Appl. Catal., B 2013, 142-143, 568-578.

480

38. Peng, X.; Guo, Y.; Yin, Q.; Wu, J.; Zhao, J.; Wang, C.; Tao, S.; Chu, W.; Wu, C.; Xie, Y., Double-exchange

481

effect in two-dimensional MnO2 nanomaterials. J. Am. Chem. Soc. 2017, 139, (14), 5242-5248.

482

39. Liu, Y.; Zhang, P., Removing surface hydroxyl groups of Ce-modified MnO2 to significantly improve its

483

stability for gaseous ozone decomposition. J. Phys. Chem. C 2017, 121, (42), 23488-23497.

484

40. Chen, H.; Stanier, C. O.; Young, M. A.; Grassian, V. H., A kinetic study of ozone decomposition on

485

illuminated oxide surfaces. J. Phys. Chem. A 2011, 115, (43), 11979-11987.

486

41. Tompsett, D. A.; Parker, S. C.; Islam, M. S., Rutile (β-)MnO2 surfaces and vacancy formation for high

487

electrochemical and catalytic performance. J. Am. Chem. Soc. 2014, 136, (4), 1418-1426.

488

42. Tompsett, D. A.; Parker, S. C.; Islam, M. S., Surface properties of α-MnO2: Relevance to catalytic and

489

supercapacitor behaviour. J. Mater. Chem. A 2014, 2, (37), 15509-15518.

490

43. Li, W.; Gibbs, G. V.; Oyama, S. T., Mechanism of ozone decomposition on a manganese oxide catalyst. 1. in

491

situ Raman spectroscopy and ab initio molecular orbital calculations. J. Am. Chem. Soc. 1998, 120, (35),

492

9041-9046.

493

44. Knoops, H. C. M.; Elam, J. W.; Libera, J. A.; Kessels, W. M. M., Surface loss in ozone-based atomic layer

494

deposition processes. Chem. Mater. 2011, 23, (9), 2381-2387.

495

45. Roscoe, J. M.; Abbatt, J. P. D., Diffuse reflectance FTIR study of the interaction of alumina surfaces with 28

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

Environmental Science & Technology

496

ozone and water vapor. J. Phys. Chem. A 2005, 109, (40), 9028-9034.

497

46. Yang, L.; Hu, C.; Nie, Y.; Qu, J., Catalytic ozonation of selected pharmaceuticals over mesoporous

498

alumina-supported manganese oxide. Environ. Sci. Technol. 2009, 43, (7), 2525-2529.

499

47. Xi, Y.; Reed, C.; Lee, Y.-K.; Oyama, S. T., Acetone oxidation using ozone on manganese oxide catalysts. J.

500

Phys. Chem. B 2005, 109, (37), 17587-17596.

501

48. Jing, X.-L.; Chen, Q.-C.; He, C.; Zhu, X.-Q.; Weng, W.-Z.; Xia, W.-S.; Wan, H.-L., Mechanistic aspects of

502

photo-induced formation of peroxide ions on the surface of cubic Ln2O3 (Ln = Nd, Sm, Gd) under oxygen. Phys.

503

Chem. Chem. Phys. 2012, 14, (19), 6898-6904.

504

29

ACS Paragon Plus Environment