Active Complexes on Engineered Crystal Facets of ... - ACS Publications

subdivided flats, and lack of windows and ventilation in most enclosed areas. Considering the. 91 toxicity of HCHO at very low concentrations, standar...
0 downloads 0 Views 3MB Size
Subscriber access provided by Uppsala universitetsbibliotek

Remediation and Control Technologies

Active Complexes on Engineered Crystal Facets of MnOx–CeO2 and Scaleup Demonstration on an Air Cleaner for Indoor Formaldehyde Removal Haiwei Li, Wing Kei Ho, Jun-ji Cao, Duckshin Park, Shun Cheng Lee, and Yu Huang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b03197 • Publication Date (Web): 23 Aug 2019 Downloaded from pubs.acs.org on August 23, 2019

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

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

Page 1 of 42

Environmental Science & Technology

1

Active Complexes on Engineered Crystal Facets of MnOx–CeO2 and Scale-up

2

Demonstration on an Air Cleaner for Indoor Formaldehyde Removal

3 4

Haiwei Li,† Wingkei Ho,‡,§ Junji Cao,‖ Duckshin Park,‖‖ Shun-cheng Lee,*,† Yu Huang *,‖

5 6

†Department

7

Hong Kong, China.

8

‡Department

9

Hong Kong, China.

of Civil and Environmental Engineering, The Hong Kong Polytechnic University,

of Science and Environmental Studies, The Education University of Hong Kong,

10

§State

11

China.

12

‖State

13

Aerosol Chemistry and Physics, Institute of Earth Environment, Chinese Academy of Sciences,

14

Xi’an 710061, China.

15

‖‖Transportation

16

Gyeonggi-do, South Korea.

Key Laboratory of Marine Pollution, The City University of Hong Kong, Hong Kong,

Key Laboratory of Loess and Quaternary Geology (SKLLQG) and Key Laboratory of

Environmental Research Team, The Korea Railroad Research Institute,

17 18 19

*Corresponding Authors: Prof. Shuncheng Lee, E-mail: [email protected]; Prof. Yu Huang,

20

E-mail: [email protected].

1

ACS Paragon Plus Environment

Environmental Science & Technology

21

TOC Art

22 23

2

ACS Paragon Plus Environment

Page 2 of 42

Page 3 of 42

Environmental Science & Technology

24

ABSTRACT: Crystal facet-dominated surfaces determine the formation of surface-active

25

complexes, and engineering specific facets is desirable for improving the catalytic activity of

26

routine transition metal oxides that often deactivate at low temperatures. Herein, MnOx–CeO2

27

was synthetically administered to tailor the exposure of three major facets, and their distinct

28

surface-active complexes concerning the formation and quantitative effects of oxygen vacancies,

29

catalytically active zones, and active-site behaviors were unraveled. Compared with two other

30

low-index facets {110} and {001}, MnOx–CeO2 with exposed {111} facet showed higher

31

activity for formaldehyde oxidation and CO2 selectivity. However, the {110} facet did not

32

increase activity despite generating additional oxygen vacancies. Oxygen vacancies were highly

33

stable on the {111} facet and its bulk lattice oxygen at rapid migration rates could replenish the

34

consumption of surface lattice oxygen, which was associated with activity and stability. High

35

catalytically active regions were exposed at the {111}-dominated surfaces, wherein the

36

predominated Lewis acid–base properties facilitated oxygen mobility and activation. The

37

mineralization pathways of formaldehyde were examined by a combination of in situ X-ray

38

photoemission spectroscopy and diffuse reflectance infrared Fourier transform spectrometry. The

39

MnOx–CeO2-111 catalysts were subsequently scaled up to work as filter substrates in a

40

household air cleaner. In in-field pilot tests, 8 h of exposure to an average concentration of

41

formaldehyde after start-up of the air cleaner attained the Excellent Class of Indoor Air Quality

42

Objectives in Hong Kong.

3

ACS Paragon Plus Environment

Environmental Science & Technology

43

Page 4 of 42

INTRODUCTION

44

Catalytic oxidation processes at low temperatures using either supported noble metals (e.g.,

45

Au, Ag, or Pt)1-3 or non-noble transition metal oxides (containing Mn, Co, or Zr)4-6 have been

46

extensively studied for air pollution control. Earth-abundant MnOx is often modified in hybrid

47

fabrications due to its high reversible capacitance and structural flexibility4, and an ensuing

48

redox loop with cyclic electron transfer can be sustained via a Mn dismutation when reduced

49

oxidation states of Mn are periodically remediated by the oxygen-carrier materials such as

50

CeO27, 8, Fe2O39, or Al2O310. Among them, MnOx–CeO2 mixed oxides have been examined in the

51

catalytic oxidation of priority gaseous pollutants, such as formaldehyde (HCHO)11-14, carbon

52

monoxide (CO)15, and nitrogen oxides (NOx)16, 17. Given reactive metal–particle interfaces, the

53

promoting effect of noble metal nanoparticles on MnOx–CeO2 mixed oxides (e.g., Ag/MnOx–

54

CeO218 and Pt/MnOx–CeO219) has been obtained during the nearly complete oxidation of HCHO

55

at low temperatures. To date, the industrial development of relatively inexpensive but effective

56

noble metal-free catalysts is needed, and catalyst deactivation occurs at ambient temperature11, 20.

57

Surface-active complexes concerning defect formation, activation of reactive oxygen

58

species, molecule adsorption, and heterogeneous oxidation are proportional to exposed crystal

59

facets. Engineering crystal facets induces distinct oxygen vacancy clusters within different

60

exposed facets and possesses various physical and chemical performances of structured crystal

61

atoms21, 22. From an industrial perspective, engineering surface structures for binary transition

62

metal oxides is more apt to their scale-up production relative than either single nanostructured

63

oxidized metals or ternary junctions4,

64

uncertain in the Mars–Van Krevelen redox mechanisms of heterogeneous catalysis11, 23 or in a

65

solid solution containing complex solvents or additives12,

23, 24,

because the reactivity and facet control are often

4

21.

ACS Paragon Plus Environment

As such, ceria with a fluorite

Page 5 of 42

Environmental Science & Technology

66

structure has three major low-index facets: nanocubes preferentially dominated by the {110}

67

facet and nanorods dominated by the {111} and {100} facets25, 26. The exposed {110} and {100}

68

facets can facilitate the migration of lattice oxygen from the bulk to the surface, but the process

69

is restricted on the {111} surfaces27. However, the dominant {110} and {100} facets are inclined

70

to deactivate with reaction time because of the low stability of the oxygen vacancies employed25-

71

28.

72

Mn, Ti, or Zr12, 27, 29, 30. By contrast, when manganese ions (Mnn+) with small ionic radius and

73

low compositions are stabilized in MnOx–CeO2 mixed oxides, the oxygen mobility is strongly

74

improved, and higher catalytic activities are achieved at low temperatures than single oxidized

75

metals4, 11-13, 19. In MnOx–CeO2, ceria is not directly involved in catalytic oxidation but functions

76

as an oxygen carrier to keep high-valence manganese oxide via a Mn dismutation reaction.

77

Moreover, the enhancement of reactivity is ascribed to the mobility and activation of lattice

78

oxygen26, 31. Exposing specific crystal facets remains challenges. The exposed facets with high

79

surface energy can decline with a fast growth rate in the bulk of crystals; hence, the

80

thermodynamically stable facets predominated on crystal surfaces could minimize the total

81

surface energy21. Key factors, such as solvents, additives, and impurities in a solid solution, can

82

also influence the final shape of crystals. Therefore, appropriate surfactant-assisted/capping

83

agents (e.g., citric acid and cetyltrimethylammonium bromide) have been chosen to tailor the

84

exposure of crystal facets11, 12, 21, 26. For example, the ceria {001} facet with fast growth was well

85

administered using decanoic acid as an organic ligand molecules32. Aside from the facet

86

reforming and growth, a large amplitude of catalytically active regions is indispensable for

87

exposed surface-active complexes within facet-dominated surfaces33, 34.

88

The stability of CeO2 nanocrystals can be maintained by doping with stable metals, such as

Translating laboratory work into commercial value at large is the focus of our research. The

5

ACS Paragon Plus Environment

Environmental Science & Technology

89

situation of indoor air pollution is aggravated by numerous factors, including high-rise living,

90

subdivided flats, and lack of windows and ventilation in most enclosed areas. Considering the

91

toxicity of HCHO at very low concentrations, standards for the emission levels of HCHO have

92

become stringent in the Indoor Air Quality Certification Scheme in Hong Kong, i.e., Excellent

93

Class with 30 µg·m−3 and Good Class with 100 µg·m−3 for an 8 h of exposure35, 36. To eliminate

94

HCHO in indoor air, our group developed various collosol-film coating, purifier filter substrates,

95

or honeycomb-like reactors based on heterogeneous catalysis to use in different environments37,

96

38.

97

case scenario for the formulation and development of mitigation strategies.

Knowledge gained from laboratory tests and field campaigns can be used to estimate the best-

98

Herein, MnOx–CeO2 catalysts with different major exposed facets were synthesized using

99

varying morphology-controlling methods. The exposed three major {111}, {110}, and {001}

100

facet-dominated surfaces were systematically studied in terms of activity and selectivity,

101

formation and quantitative effects of oxygen vacancies, identification of catalytically active

102

regions, and intermediate pathways. Efficient MnOx–CeO2-111 catalysts were scaled up to

103

function as filter substrates of a household air cleaner. The materials obtained from the

104

laboratory-scale experiments were examined and validated in in-field pilot tests of the air cleaner

105

to evaluate their actual removal effectiveness.

106

MATERIALS AND METHODS

107

Catalyst Synthesis. MnOx–CeO2-111 (denoted as MCO-111) was synthesized by a

108

hydrothermal redox reaction containing 50 wt% Mn(NO3)2 ( 8 mM) solution, Ce(NO3)3·6H2O (8

109

mM), and (NH4)2S2O8 (16 mM). The resulting mixture was transferred into a 50 mL Teflon-lined

110

stainless-steel autoclave after stirring for 2 h and then heated at 140 °C for 12 h. The

111

hydrothermal products were washed with ultrapure water (Milli-Q system, Millipore Inc.), dried

6

ACS Paragon Plus Environment

Page 6 of 42

Page 7 of 42

Environmental Science & Technology

112

at 70 °C, and calcined at 350 °C for 6 h. Finally, the calcined products were treated with 100 mL

113

of citric acid solution (2 mM) before washing and drying.

114

MnOx–CeO2-110 (MCO-110) was prepared by co-precipitation. A mixture of the same 50

115

wt% Mn(NO3)2 (8 mM) solution and NH4Ce(NO3)4·6H2O (8 mM) solution was dissolved in

116

ultrapure water followed by the dropwise addition of the precipitant NaOH solution (20 mM).

117

The resulting solution was added to an aqueous solution of cetyltrimethylammonium bromide

118

(CTAB, Sigma ≥ 98%) under stirring for 2 h and then heated at 70 °C for 12 h. The harvested gel

119

precipitates were washed and filtered before drying at 70 °C and calcining at 500 °C for 6 h.

120

MnOx–CeO2-001 (MCO-001) was prepared with the synthesis route similar to that of MCO-

121

111 but the mixture was added with organic ligand molecules, i.e., decanoic acid (50 mg). After

122

the hydrothermal reaction at 400 °C for 1 h and quenching in a water bath to room temperature,

123

the organic ligand-modified products were extracted from a mixture of 3 mL of hexane and 15

124

mL of ethanol. The molar ratio of decanoic acid to ceria precursor did not generally exceed 6:1

125

to overcome the surfaces enclosed by the {111} planes. Considering that the precursor CeO2

126

{001} surface is less stable than the {111} surface, the organic ligand molecules can lead to the

127

formation of the exposed {001} surface toward the crystal growth in the [111] direction instead

128

of the [001] direction32.

129

Characterizations. The as-prepared catalysts were characterized by the following

130

techniques: X-ray powder diffraction (XRD; Philips X’pert Pro Super diffractometer), high-

131

resolution transmission electron microscopy (HRTEM; JEOL JEM-2010), Brunauer–Emmett–

132

Teller analysis (BET; Micromeritics Gemini VII 2390 Norcross, GA), thermogravimetric

133

analysis (TGA; Setaram Setsys 16/18 thermoanalyzer), inductively coupled plasma atomic

134

emission spectroscopy (ICP-AES; Varian), H2 temperature-programmed reduction (H2–TPR), O2

7

ACS Paragon Plus Environment

Environmental Science & Technology

135

temperature-programmed desorption (O2–TPD), X-ray photoemission spectroscopy (XPS;

136

Thermo ESCALAB 250Xi) calibrated with C 1s at 284.8 eV, and pyridine adsorbed IR

137

spectroscopy (Py-IR; Tensor 27, Bruker). The detailed characterization instruments and methods

138

are provided in Supporting Information (SI), and the schematic diagram and operating

139

parameters infrared thermography (IRT; FLIR camera SC7000 system) are demonstrated in SI

140

Figures S1–S2 and Table S1.

141

Laboratory Activity Test. Under the reaction condition of 5 ppm HCHO/21 vol% O2/N2

142

and gas hourly space velocity (GHSV) = 4 × 104 h–1, the catalytic activity for the conversion of

143

HCHO into CO2 was tested in a thermostatic fixed-bed reaction system, as described in SI. The

144

recycling catalytic activity was conducted in 5 batch modes of 90 minute measurements under

145

the similar reaction condition, each of which continued when the HCHO concentration was back

146

to its initial level.

147

Designed H2–TPR Experiments. The quantitative effects of oxygen vacancies within

148

different exposed facets were evaluated through the consumption of reductant H2 with lattice

149

oxygen in two types of designed H2–TPR experiments, i.e., the migration of bulk lattice oxygen

150

treated at 600 °C and the consumption of surface lattice oxygen treated at 450 °C, wherein the

151

two types of lattice oxygen of MCO catalysts were ran into release and decline, consistent with

152

the TGA profile (SI Figure S3). The samples were pretreated similarly to the H2–TPR

153

measurements (see in SI) and heated to 600 °C at a heating rate of 10 °C·min–1 for various

154

treatment periods (1, 5, and 10 h) under 10 vol% H2/Ar gas flow. The corresponding H2

155

consumption was monitored by using an online thermal conductivity detector to calculate the

156

migration amount of bulk lattice oxygen after reheating to 950 °C. The pretreated samples in the

157

second experiment were treated at 450 °C under identical treatment conditions and then cooled to

8

ACS Paragon Plus Environment

Page 8 of 42

Page 9 of 42

Environmental Science & Technology

158

100 °C under ultra-pure Ar feed. After reheating to 950 °C, the corresponding H2 consumption

159

was recorded in a similar manner to calculate the consumption amount of surface lattice oxygen.

160

In Situ XPS and Diffuse Reflectance Infrared Fourier Transform Spectrometry

161

(DRIFT) Analysis. The detailed in situ XPS and DRIFT experiments are described in SI. The

162

assignment of C 1s XPS photopeaks is listed in SI Table S2 in accordance with the binding

163

energies (BEs) of different possible chemical bonds.

164

RESULTS AND DISCUSSION

165

Structural Characterizations. As shown in Figure 1a, the entire characteristic peaks of

166

MCO with different exposed facets shifted toward lower 2θ angles relative to those of precursor

167

ceria {111}, {110}, and {001} (see also SI Figure S4), which was due to the facet growth after

168

some Ce4+ ions (0.087 nm) of ceria were replaced with Mnn+ of a small ionic radius (i. e., Mn4+

169

of 0.053 nm and Mn3+ of 0.065 nm) in the mixed oxides. The intensities of diffraction patterns

170

for MnOx within the three exposed facets, i.e., main Mn3O4 hausmannite phase (JCPDS 07-1841)

171

together with small Mn2O3 (JCPDS 07-0856) and Mn5O8 (JCPDS 39-1218), became weak

172

because their individual facet growth in lattice plane of ceria was restricted and the Mn2+ and

173

Mn3+ ions of Mn3O4 were partially oxidized into Mn4+ ions after acid treatment.12, 14, 26

174

The successful fabrications of the three major exposed facets were examined and validated

175

in HRTEM images (Figure 1b). In compliance with the downward shift of XRD patterns, the d-

176

spacings of MCO (110) and (110) planes were estimated at 0.49 and 0.31 nm, larger than those

177

of pristine ceria (111) and (110) of 0.31 and 0.19 nm (SI Figure S5), respectively. The increases

178

in d-spacings were associated with the formation of defect sites (e.g., oxygen vacancies or Ce4+

179

reductions)26, 31 within the exposed MCO facets (Figure 1c). The corresponding SAED pattern

180

shows a similar angle at 45o between (111) and (100) planes for MCO-111 compared to

9

ACS Paragon Plus Environment

Environmental Science & Technology

181

precursor ceria {111}, which agrees with the {111} surfaces were regularly displayed along a

182

direction. MCO-110 was identified according to an angle of 45o between (100) and (110)

183

planes, while the {001} facet exposure was related to an angle at 90o of between (001) and

184

(110)26,

185

unchanged despite increased facet growth, indicating that the facet and size control for the

186

engineered MCO surfaces were well regulated. These results could be correlated with the molar

187

ratio of Mn/(Mn+Ce) administered below 0.512,

188

specific surface areas but affect their immobility in the mixed oxides after rigorous aging

189

treatment4.

27.

The angle between specific crystal planes for the three exposed facets remained

14, 19.

High Mn contents could enlarge the

190

Mnn+ doping reformed the surfaces and improved the crystal structures (SI Table S3)

191

compared with pristine MnOx and CeO2. An apparent plane (211) with d-spacings of 0.38 nm

192

was assigned to Mn3O4 phase preferentially immobilized within the stable MCO {111} surfaces

193

after citric acid treatment12. Relative to the pristine {111} facet dominated in ceria nanoparticles

194

and the pristine {110} facet in ceria nanorods (see also SI Figure S5), the MCO-111 and MCO-

195

110 surfaces became cylindrical, and their framework architectures changed from facetted pores

196

and channels to smooth pores and channels within the well-defined {111} and {100}-dominated

197

surfaces (Figure 1b). Since Mn contents in the mixed oxides were approaching (SI Table S3),

198

both MCO-111 and MCO-110 demonstrated similar size with a diameter of approximately 50

199

nm. By contrast, engineering MCO {001} facets is often uncontrolled. The crystal growth of

200

MCO-001 was along the [111] direction became predominant and was confined to transform an

201

irregular cylindrical shape with the exposed {001} facets. MCO-001 exhibited a small diameter

202

of approximately 30 nm, but it was much less stable than the two other facets, because the crystal

203

growth of ceria {001} surfaces was quite fast in the [001] direction27. The synthesis and control

10

ACS Paragon Plus Environment

Page 10 of 42

Page 11 of 42

Environmental Science & Technology

204

of high-quality crystal facet {001} depend on organic ligand-assisted liquid-solid-solution phase

205

synthetic transfer routes32. Considering that the reversible transition of Mn4+ and Mn3+ is directly

206

coupled with the localization/delocalization of the 4f electron of cerium, a high percentage of

207

unsaturated atoms in the MnOx−CeO2 mixed oxides can possess superior reactivity. However,

208

this effect could increase the challenge in engineering high-purity and well-defined crystal

209

surfaces.

210 211

Figure 1. XRD patterns of pristine MnO2, pristine CeO2, and MCO catalysts with different

212

exposed facets (a); TEM and HRTEM images of MCO catalysts with different exposed facets

213

(b); and atomic structure of MCO-111, MCO-110, and MCO-001 (c).

11

ACS Paragon Plus Environment

Environmental Science & Technology

214

TPR and TPD Profiles. Two reduction peaks were detected in a broad temperature range

215

of 250 °C–500 °C, as shown in Figure 2a, indicating a reduction/oxidation cycle between

216

transition-metal oxides with high valence and those with low valence during the ensuing redox

217

loops of the oxidized Mn–Ce ions. By using ceria as oxygen carrier, oxygen migration was

218

improved, and the redox-looping processes were sustained by ceria in the cyclic valence

219

transitions of Ce4+–Mn3+ ↔ Ce3+–Mn4+12,

220

XPS results (SI Figures S6a–d). The catalytic activity and stability, highly correlated with high

221

oxidation states in the Mn dismutation reaction, would be affected by the reduced Mn oxides if

222

they cannot be timely remediated. The consumption of lattice oxygen species could be

223

replenished periodically in the ensuing redox loop of transition-metal oxides. In comparisons

224

with the two other facets, the first reduction peak of MCO-111 moved to 267 °C, which

225

corresponded to the consumption of surface oxygen species (Oads), because the adsorbed oxygen

226

species are easily dissociated on the catalyst surface31. The reduction of cerium and manganese

227

ions did not largely result in a downward shift of the reduction peaks for all the three MCO

228

samples. By contrast, more apparent downward shift to low temperatures of the two reduction

229

peaks of MCO-111 was due to the continuous reduction of Mn4+ to Mn2+40, 41. The intensities of

230

the reduction peaks at low temperatures for MCO-111 were the highest, ascribing to the mobility

231

of oxygen species and the generation of –OH species40. Larger mobility and consumption of

232

lattice oxygen within MCO-111 were observed relative to MCO-110 and MCO-001. Mn4+ and

233

Mn3+ were the two main Mn oxidation states in Mn 3s spectra obtained in SI Figure S7. Hence,

234

the contribution of the high-temperature peak to the total reduction profile was associated with

235

the manganese ions with high oxidation states (Mn4+) in the recycling redox loops. The central

236

position of the reduction peaks of MCO-111 approached those of the reported Pt/MCO19, whose

14.

These findings matched with the corresponding

12

ACS Paragon Plus Environment

Page 12 of 42

Page 13 of 42

Environmental Science & Technology

237

catalytic activity often stands out from the majority of room-temperature catalytic oxidation

238

(RCO) catalysts.

239 240

Figure 2. H2–TPR (a) and O2–TPD (b) profiles of MCO catalysts with different exposed facets.

241

The O2–TPD spectra (Figure 2b) were analyzed to initially understand the roles of oxygen

242

species within the different exposed facets, which corresponded to the formation of oxygen

243

vacancies and subsequently access the HCHO oxidation mechanisms. A large amount of O2

244

desorbs from the catalyst at T600 °C is named as γ-O2, indicating that the remaining lattice oxygen species

249

continue to decline, and high oxidation states in metal oxides are completely reduced at the high-

250

temperature range41-43. Here, the O2 desorption temperature of the three facets generally shifted

251

toward the low-temperature region, indicating that the MCO catalysts were favorable in low-

13

ACS Paragon Plus Environment

Environmental Science & Technology

252

temperature activity stemming from the desirable BET performance and bulk lattice oxygen

253

migration of cerium-based oxygen carriers27, 31. The adsorbed O2 could be readily transformed

254

into atomic oxygen (O*) and continue to form surface active oxygen (O– and O2–) at the catalyst

255

surfaces40. A large amount of oxygen adsorption and migration at low temperature could benefit

256

the conversion of O2 into O*, leading to HCHO oxidation. MCO-110 possessed strong O2

257

desorption at the α-O2 peak range, which corresponded to the increased formation of oxygen

258

vacancies on the surface relative to MCO-111 and MCO-001 and were consistent with the

259

corresponding O 1s XPS spectra (SI Figure S8).

260

RCO Activity and Selectivity. Figure 3a shows the turnover frequencies (TOFs), defined

261

as the HCHO consumption rates per unit of surface area (molHCHO·s−1·mcat−2), over the MCO

262

with different exposed facets. Under a GHSV of 4 × 104 h–1, the activities presented facet

263

dependence and followed the sequence of MCO-111>MCO-110>MCO-001. The consumption

264

rates of MCO-111 of approximately 7.5 × 10−4 μmol·s−1·m−2 at a low temperature range of

265

25 °C–45 °C were higher than those of MCO-110 (approximately 4.5 × 10−4 μmol·s−1·m−2) and

266

MCO-001 (3.0×10−4 μmol·s−1·m−2). In comparison with some typical transition metal-based

267

catalysts in SI Table S4, the temperatures for complete HCHO oxidation must exceed 70 °C.

268

The increase in reaction temperature can increase the selectivity toward CO2 and decrease the

269

production of formic acid intermediates19. The consumption rates of MCO-111 and MCO-110

270

catalysts at low temperatures were parallel with those of reported transition metal oxides at high

271

temperatures. The catalytic activity of MCO-111 even approached that of some Ag-supported

272

catalysts2, 44. Notably, noble metal-supported catalysts (e.g., Pt/TiO245 and Pd/TiO246) can result

273

in complete HCHO conversion at near-room temperature. Considering the limitations in the

274

scaling-up cost and mass dispersion control of the noble metal, transition metal oxides may be

14

ACS Paragon Plus Environment

Page 14 of 42

Page 15 of 42

Environmental Science & Technology

275

promising substitutes if lattice defects or oxygen vacancies are engineered to act as additional

276

active sites. The recycling activity and stability of MCO with the different exposed facets were

277

determined in Figure 3b. MCO-111 converted 56% HCHO into CO2 at 35 °C, followed by

278

MCO-110 (50%) and MCO-001 (42%). No apparent deactivation was observed in 360 min

279

reaction periods. In addition, the half reaction time (t50%) was independent on the initial reactant

280

concentration (SI Figure S9), suggesting that HCHO catalytic oxidation followed a pseudo-first-

281

order reaction and the RCO system could be adapted to HCHO removal at the sub-ppm level47,

282

48.

283

Oxygen vacancy sites are responsible for adsorption and catalytic activity. Interestingly, the

284

formation of oxygen vacancies was feasible on the {110} and {001} facets but did not incur

285

comparatively high activity (Figure 3). Except for the distorted electronic structure of crystals

286

with different exposed facets, the activity was also influenced by the migration of definite

287

surface oxygen and stability of the formed oxygen vacancies31. The quantitative effects of

288

oxygen vacancies regarding to oxygen storage capacity are analyzed to demonstrate the

289

migration of bulk lattice oxygen in the following section.

290 15

ACS Paragon Plus Environment

Environmental Science & Technology

291

Figure 3. HCHO consumption rate as a function of reaction temperature (a) and recycling

292

catalytic activity and stability in 90 min of each run (b) over MCO catalysts with different

293

exposed facets at 35 °C

294

Quantitative Effects of Oxygen Vacancies. The oxygen vacancy defects were engineered

295

to work as active centers that were correlated with adsorption, catalytic activity, and stability.

296

The O2–TPD and O 1s XPS (SI Figure S8) studies roughly demonstrated oxygen speciation in

297

the chemical-looping processes of metal oxides but failed to clearly elucidate the roles and

298

oxygen storage capacity of definite lattice oxygen, i.e., surface lattice oxygen OS–L and bulk

299

lattice oxygen OB–L31. The OS–L, rather than the surface adsorbed oxygen, was found to be

300

responsible for the catalytic activity and replenished from the migration of OB–L31. Understanding

301

the quantitative effects of oxygen vacancies is of practical importance in the catalyst design

302

based on high-performance oxygen mobility. The mobility between the migration and

303

consumption of lattice oxygen and the formation of oxygen vacancies within the exposed facets

304

were investigated through the designed H2–TPR experiments.

305

The corresponding H2 consumption amounts of OB–L at 600 °C for different treatment times

306

are summarized in SI Table S5. The H2 consumption amounts of OB–L were approaching to a

307

minimum level when the MCO samples were treated beyond 5 h. MCO-110 had the lowest H2

308

consumption during the whole treatment, which was correlated with the formation of the

309

maximum oxygen vacancies within MCO-110 obtained in the O2–TPD profiles (Figure 2b). The

310

corresponding migration rate was the slowest at 8.0 × 10−3 µmol·g−1·s−1. The fastest migration

311

rate up to 1.1 × 10–2 µmol·g−1·s−1 was found for MCO-001 but failed to result in the highest

312

catalytic activity, because the {001} facet growth was somewhat suppressed by the ceria (111)

313

and (200) planes32. Oxygen vacancies were more stable within pristine ceria {111} facet than

16

ACS Paragon Plus Environment

Page 16 of 42

Page 17 of 42

Environmental Science & Technology

314

within pristine ceria {110} and {001}, they occurred more easily within the {110} and {001}

315

though25, 27. Aside from the stable defect structure of crystal facet, improvements in catalytic

316

activity were ascribed to the fact that the OB–L mobility at faster migration rate can replenish OS–L,

317

but the excessively fast rate within the {001} facet can decrease activity31. Oxygen vacancies

318

could result in excellent catalytic activity and stability if they were more stable on catalyst

319

surfaces.

320

As shown in Table 1, the designed H2–TPR experiment at 450 °C was performed and

321

examined at different treatment times to demonstrate the total consumption based on the

322

migration of OB–L to OS–L. The actual consumption rates of OS–L within MCO-111 were larger

323

than those within MCO-110 and MCO-001 and remained stable after 5 h treatment relative to the

324

migration rate of OB–L. However, the consumptions of OS–L cannot be completely ruled out from

325

the migration of OB–L, possibly because of some mass loss in OB–L mobility. Thus, the largest

326

consumption amounts of OS–L were not well associated with the largest consumption amounts of

327

OB–L. Therefore, OB–L through its faster migration can not only participate in the catalytic

328

reaction but also replenish the consumed OS–L. These events directly supported the importance of

329

stable oxygen vacancies within the MCO-111 facet.

330

Table 1. H2 consumption of OB–L and OS–L of MCO samples treated at 450 °C with 10% H2/Ar

331

for different treatment times. H2 consumption Samples of OB–L

Actual H2 consumption of OS–L

OB–L migration

OS–L

(µmol·g−1)

(µmol·g−1) a

(µmol·g−1) b

(µmol·g−1) Untreated MCO-111

645

consumption rate

consumption of OS–L (µmol·g−1·s−1) c

568

0

17

ACS Paragon Plus Environment

0

0

Environmental Science & Technology

Page 18 of 42

Untreated MCO-110

549

516

0

0

0

Untreated MCO-001

676

593

0

0

0

Treated MCO-111 (1 h)

633

409

6

79.5

2.0 × 10–2

Treated MCO-110 (1 h)

539

373

5

71.5

1.8 × 10–2

Treated MCO-001 (1 h)

661

441

7.5

76

1.9 × 10–2

Treated MCO-111 (5 h)

315

166

165

201

2.0 × 10–3

Treated MCO-110 (5 h)

241

159

154

178.5

1.4 × 10–3

Treated MCO-001 (5 h)

279

163

198.5

215

0.9 × 10–3

332

a

333

300 °C–600 °C.

334

b

335

treated sample)/2.

336

c

337

L)/treatment

H2 consumption of OS–L was calculated according to TPR profiles at temperature range of

Consumption amount of OS–L = (H2 consumption of untreated sample−H2 consumption of

Actual consumption rate of OS–L = (consumption amount of OS–L−migration amount of OB– time.

338

Identification of Catalytically Active Regions and Active-Site Behaviors at Low

339

Temperatures. The activation of oxygen species and catalytic oxidation are favorable to high

340

reaction temperatures. However, the performance of the temperature-dependent catalysts that can

341

trigger their activity at low temperatures has been rarely studied. Here high-throughput screening

342

studies with IR thermography were performed to distinguish and identify catalytically active

343

regions33, 34. As shown in Figure 4, the catalytically active regions corresponding to temperature

344

variations were characterized and compared in various isothermal feeds. First, the IR absorbance

345

intensities of the different exposed facets were obtained from 25 °C–175 °C, which were closely

346

consistent with different surface active regions at varying feed temperatures (Figure 4a). The IR

18

ACS Paragon Plus Environment

Page 19 of 42

Environmental Science & Technology

347

absorbance was monitored under the average of the emerging temperatures when a hot spot

348

region of IR thermography was formed on the catalyst surfaces. Small changes in the IR

349

absorbance were observed in a high temperature range of 125 °C–175 °C, despite a gradual

350

decline with increasing feed temperatures. Moreover, no huge difference in the IR absorbance

351

and retention was found among the three exposed facets at the high-temperature range. Below

352

75°C, comparatively large IR absorbance intensities were attained at MCO-111 compared with

353

MCO-110 and MCO-001. MCO-111 maintained enhanced intensity and retention of the IR

354

absorbance under the emerging temperatures at 25 °C–50 °C. The IR thermography images were

355

then recorded to visualize the active regions at low temperatures after the emerging temperature

356

reached 50 °C in Figure 4b. The amplitude (i.e., retention of the maximum emerging

357

temperature after hot spot formation) of the temperature gradient could be clearly compared at

358

the different crystal facets. The amplitude of the maximum temperature gradient was broader and

359

stronger at MCO-111 than at the two other facets, although the IR absorbance was released and

360

extinguished during cooling to 25°C. The highly active zones displayed at MCO-111 could act as

361

a reactive catalytic bed to sustain the exposure of abundant active sites and participation of

362

oxygen species at low temperatures, thereby providing the enhanced catalytic activity of MCO-

363

111 under ambient conditions.

19

ACS Paragon Plus Environment

Environmental Science & Technology

364 365

Figure 4. IR absorbance intensities of MCO catalysts with different exposed facets at varying

366

feed temperatures from 25 °C to 175 °C (a) and identification of catalytically active zones after

367

the emerging temperatures reached 50 °C via IR thermography images (b).

368

The types and properties of surface active sites at different exposed facets were investigated

369

through Py-IR measurements49,

50.

370

kinds of surface metal sites, i.e., Lewis acid (L-acid) and Brønsted acid (B-acid) sites, are related

371

to free electron and proton exchange at metal oxides. For transition-metal oxides, the Lewis

372

acid–base properties depend on cyclic electron transfer in the chemical-looping processes and

373

play a major role in the activation of reactive oxygen species51. As shown in Figures 5, the peaks

374

at ca. 1456 and 1610 cm−1 bands originated from pyridine adsorbed onto L-acid sites, whereas

375

the peaks at ca. 1545 and 1635 cm−1 bands were related to B-acid sites. The peak at ca. 1489

376

cm−1 originated from pyridine adsorbed onto L-acid and B-acid sites. The exposure of

Metal oxides are used for their acid–base properties. Two

20

ACS Paragon Plus Environment

Page 20 of 42

Page 21 of 42

Environmental Science & Technology

377

predominant L-acid sites was observed at the three major exposed facets. The Lewis/Brønsted

378

(L/B) ratio at MCO-111 was 2.85 larger than that at MCO-110 and MCO-001 with 2.11 and

379

1.94, respectively. A high degree of nucleophilic substitution was assigned to an exchange of

380

surface metal sites and atomic O*, and the hopping of hydrocarbon compounds to atomic O* was

381

preferential onto L-acid sites. B-acid sites acted as predominant active sites in favor of high

382

reaction temperatures52. Furthermore, the exposure of L-acid and B-acid sites was influenced by

383

high ambient humidity49, 50. However, many routine metallic catalysts were water-repellent that

384

affects HCHO adsorption and catalytic oxidation under ambient conditions to a certain extent.

385

Therefore, Lewis acid sites involving the superior selective adsorption of electrolyte cations were

386

presented for polar HCHO molecules induced by cyclic electron transfer. The acid–base

387

properties provided explicit evidence of strength and distribution of surface-active sites. The

388

predominance of L-acid sites at low temperatures could facilitate the activation of surface

389

oxygen species to reactive oxygen species and catalytic activity of RCO catalysts.

21

ACS Paragon Plus Environment

Environmental Science & Technology

390 391

Figure 5. Py-IR measurements of acid–base active sites at MCO with different exposed facets.

392

In Situ XPS and DRIFT Studies. The adsorption and reaction of HCHO over MCO

393

catalysts with different exposed facets were studied using in situ C 1s XPS measurements. As

394

shown in Figure 6a, the XPS peaks in the C 1s region were more precisely deconvoluted to

395

adventitious carbon (CA) at 284.5 eV from C–C species and at 285.5 eV from C–O–C species, to

396

formaldehyde/mono-dentate formate species (CF) at 288.2 eV, and O–C=O (carboxylate species

397

= CCB) at 289.1 eV53, 54. These peaks were consistent with the assignment of chemical bonds to

398

the BE of C 1s peaks in SI Table S2. CF signals were retained on the three exposed facets, and

399

more evolutions of CA were found for MCO-111. The C 1s core level envelope presented an

400

important contribution at the BE of CF at temperatures below 50 °C. The intensities of CCB 22

ACS Paragon Plus Environment

Page 22 of 42

Page 23 of 42

Environmental Science & Technology

401

photopeaks were lower than those of CF and CA photopeaks for the three samples. The formation

402

of CA favored increased reaction temperatures, which was preferential to CO oxidation and CO2

403

yield. Both formaldehyde and formate species were bonded end-on through the oxygen atom,

404

and the rapid conversion of CCB species continued. Therefore, physisorption was an initial

405

preadsorption step, chemisorption started with physisorption, and the strong interaction between

406

the adsorbate and sorbent surface created new types of electronic bonds (ionic or covalent). The

407

chemisorption and oxidation of HCHO were derived from H bonding of formaldehyde and

408

formate groups and bridging of CCB preferentially with oxygen atom in the MCO surfaces50, 55.

409

These findings matched well with the quantitative effects of oxygen vacancies in which excess

410

defects could not sustain the oxygen species-oriented active sites involved in the adsorption and

411

oxidation.

412

The formation of intermediate species for HCHO decomposition at 35 °C was examined

413

through in situ DRIFT spectra (Figure 6b). All transient reaction results represented the main

414

intermediates that were ascribed to formate species (including carboxylate), carbonate, CO2, and

415

water vapor. The absorbance intensities of the CO2 peak (ca. 2380 cm−1) and surface hydroxyl

416

species (a broad OH stretching region located at ca. 3580 cm−1, denoted as υs [OH]) increased

417

significantly for MCO-111 compared with exposure of MCO-110 and MCO-001 to HCHO/O2.

418

The symmetric stretching OH possesses a strongly hydrogen bonded water structure and could

419

facilitate the chemisorption of HCHO in terms of hydroxyl bonding preferable with the methyl

420

groups under low temperature and relative humidity50. MCO-111 achieved comparatively high

421

catalytic activity for HCHO conversion, producing CO2 and water vapor as final products. In

422

brief, HCHO peaks (ca. 1680 cm−1) weakened with reaction time and new typical bands

423

indicated the gradual generation of formate species (quite weak at ca. 2840 cm−1 for υs [CH], ca.

23

ACS Paragon Plus Environment

Environmental Science & Technology

424

1466 cm−1 for δ [CH2], ca. 1601 cm−1 for υs [COO], and ca. 1350 and 1314 cm−1 for υas

425

[COO])55,

426

attack of the reactive oxygen atom. The intensities of –CH2 and –OH were stronger than those of

427

the –CH match. In fact, the ensuing conversions from the intermediate CH2OO are more

428

complex than those from HCHO initially, because distinct and complex chemical speciation and

429

reactivity pathways of the CH2OO isomers (i.e., formic acid, dioxirane, and CH2OO Criegee)

430

have been found50. The conversions of HCHO into formic acid and dioxirane preferentially occur

431

through monodentate binding and bidentate coordination with metal sites of catalysts,

432

respectively. The production and removal of CH2OO Criegee are rather complex and probably

433

involved in side or secondary processes. CH2OO Criegee has some single bond characters, which

434

are assumed to a rapid dissociation of C–C and O–O bonds from the oxidation of unsaturated

435

carboxyl groups.

56.

The decomposition of –COO and –CH2 species resulted from the nucleophilic

24

ACS Paragon Plus Environment

Page 24 of 42

Page 25 of 42

Environmental Science & Technology

436 437

Figure 6. In situ C 1s XP spectra as a function of reaction temperature of the HCHO-exposed

438

MCO catalysts with different exposed facets (a) and in situ DRIFT spectra of bond cleavage at

439

35 °C (b).

440

Practical Application in Household Air Cleaner and In-field Pilot Test. MnOx–CeO2 25

ACS Paragon Plus Environment

Environmental Science & Technology

441

samples with the exposed {111} facet were examined with respect to room-temperature catalytic

442

activity in laboratory scale and were about to be validated in an in-field test for actual HCHO

443

removal. First, a prototype of a household air cleaner was designed and equipped with a filter

444

cylinder cartridge in Figure 7, whose main dimensions and operating conditions are listed in SI

445

Table S6. The cylinder cartridge consisted of five sets of porous filtration units, that is, non-

446

woven fabric filter, high efficiency particulate air (HEPA) filter, two layers of nylon fabric filter,

447

and polytetrafluoroethylene (PTFE) keel shirt, and operated through outside-in filtration. Test

448

MCO-111 particles in a sieve fraction of 140 meshes (approximately 106 μm) were scaled-up

449

and immobilized with a loading density of approximately 0.30 g·cm−2 between two layers of

450

nylon fabric filter. The fabric filter mainly removed fine particulate matter and priority gaseous

451

pollutants in indoor air (e.g., HCHO and NOx). Airflow was controlled with a built-in air pump

452

(maximum airflow, 400 m3·h−1), and purified air was released from the outlet at the top of the air

453

cleaner. After the majority of particulate matter was pre-filtered through the multi-layer

454

filtration, gaseous pollutants were in contact with the catalyst and then oxidized into harmless

455

substances under ambient conditions.

26

ACS Paragon Plus Environment

Page 26 of 42

Page 27 of 42

Environmental Science & Technology

456 457

Figure 7. Scaling-up MCO-111 filter substrates filled in the filter cylinder cartridge of a

458

household air cleaner.

459

To evaluate the HCHO removal efficiency of the MCO-111 catalyst, we conducted an in-

460

field pilot test of the household air cleaner at a newly decorated office without the use of

461

ventilation under the ambient condition of averaging RH = 68% and T = 27.5 oC, as shown in

462

Figure 8. The concentration of HCHO during a round switch-on (8 h) and -off (8 h) of the air

463

cleaner was continuously measured using a portable analyzer (Interscan 4160, USA) and double-

464

checked in the on-site measurement for 5 days. According to the air cleaner standard (GB/T

465

18801–2015, China), the sampling inlet of the analyzer was kept at a vertical distance higher

466

than 0.8 m with the ground and a horizontal distance of 1.75 m from the purified air outlet of the

467

air cleaner in a 30 m3 test site. When the air cleaner started to work, the concentration of HCHO

468

dramatically decreased and remained as low as 20 ppb over an 8 h period, which satisfied with

469

the Excellent Class (30 µg·m−3, equal with std. 24 ppb of an 8 h average) of Indoor Air Quality

27

ACS Paragon Plus Environment

Environmental Science & Technology

470

Certification Scheme for Offices and Public Places issued in Hong Kong. By contrast, when

471

switched off, the concentration of HCHO gradually recovered to its initial level at approximately

472

120 ppb, exceeding a maximum exposure of 100 ppb (equal with std. 81 ppb of an 8 h average)

473

indoors. HCHO emissions in indoor air are long-lasting and result from a source complex, such

474

as building materials, furnishings, and occupant activities. The situation is aggravated during a

475

lack of ventilation in most enclosed areas, thereby seriously burdening long-term removal

476

efficiency. The routine household air cleaner, mainly equipped with porous-media filters, shows

477

a very insufficient ability in removing gaseous pollutants. A very few commercial air cleaners

478

are available in the combined use of catalytic oxidation and mechanical filtration. Therefore, the

479

test MCO-111 catalyst can work synergically to optimize and maximize the air cleaner

480

performance for the removal of HCHO. Furthermore, this work has a proper strategy

481

recommendation for translating laboratory research into commercial value in large scale to

482

eliminate HCHO in indoor air via the efficient and cost-effective catalytic oxidation to attain the

483

standard protocols of indoor air quality.

28

ACS Paragon Plus Environment

Page 28 of 42

Page 29 of 42

Environmental Science & Technology

484 485

Figure 8. In-filed pilot tests of the household air cleaner for HCHO removal in 5 days. Inset

486

photo images: on-site measurements at a newly decorated office (approximately 30 m3) without

487

ventilation.

488

ASSOCIATED CONTENT

489

Supporting Information. The Supporting Information associated with this article can be found

490

in this section, including the detailed methods of characterizations by H2–TPR, O2–TPD, Py-IR,

491

infrared thermography, in situ XPS and DRIFT, catalytic activity measurement, assignment of

492

chemical bonds to BEs of C 1s peaks, BET performance, TGA of MCO-111, TEM and HRTEM

493

of precursor ceria {111} and {110}, survey of catalytic activity for HCHO oxidation over typical

494

RCO catalysts, migration of OB–L, high-resolution XPS spectra of surface chemical compositions,

495

determination of the half reaction time (t50%), and specification of the filter cylinder cartridge in a

29

ACS Paragon Plus Environment

Environmental Science & Technology

496

test prototype of the household air cleaner.

497

AUTHOR INFORMATION

498

Corresponding Authors

499

*Prof. Shun-cheng Lee (Email: [email protected])

500

*Prof. Yu Huang (Email: [email protected])

501

Notes

502

The authors declare no competing interests.

503

Acknowledgements

504

This work was supported by the National Key Research and Development Program of China

505

(2016YFA0203000), the Research Grants Council of Hong Kong Government (Project No.

506

T24/504/17), the Research Grants Council of Hong Kong Government (Project No.

507

PolyU152083/14E, PolyU152090/15E, and 18301117), and the Hong Kong RGC Collaborative

508

Research Fund (C5022-14G). Yu Huang was also supported by the “Hundred Talent Program”

509

of the Chinese Academy of Sciences.

30

ACS Paragon Plus Environment

Page 30 of 42

Page 31 of 42

Environmental Science & Technology

510

REFERENCES

511

(1) Ma, C.; Wang, D.; Xue, W.; Dou, B.; Wang, H.; Hao, Z., Investigation of formaldehyde

512

oxidation over Co3O4–CeO2 and Au/Co3O4–CeO2 catalysts at room temperature: effective

513

removal and determination of reaction mechanism. Environ. Sci. Technol. 2011, 45, (8),

514

3628–3634.

515 516

(2) Ma, L.; Wang, D.; Li, J.; Bai, B.; Fu, L.; Li, Y., Ag/CeO2 nanospheres: efficient catalysts for formaldehyde oxidation. Appl. Catal. B-Environ. 2014, 148, 36–43.

517

(3) Huang, M.; Li, Y.; Li, M.; Zhao, J.; Zhu, Y.; Wang, C.; Sharma, V. K., Active site-directed

518

tandem catalysis on single platinum nanoparticles for efficient and stable oxidation of

519

formaldehyde at room temperature. Environ. Sci. Technol. 2019, 53, (7), 3610–3619.

520

(4) Xu, H.; Yan, N.; Qu, Z.; Liu, W.; Mei, J.; Huang, W.; Zhao, S., Gaseous heterogeneous

521

catalytic reactions over Mn-Based oxides for environmental applications: a critical review.

522

Environ. Sci. Technol. 2017, 51, (16), 8879–8892.

523

(5) Wang, Y.; Zhu, A.; Chen, B.; Crocker, M.; Shi, C., Three-dimensional ordered mesoporous

524

Co–Mn oxide: A highly active catalyst for “storage-oxidation” cycling for the removal of

525

formaldehyde. Catal. Commun. 2013, 36, 52–57.

526

(6) Bai, L.; Wyrwalski, F.; Lamonier, J.-F.; Khodakov, A. Y.; Monflier, E.; Ponchel, A.,

527

Effects of β-cyclodextrin introduction to zirconia supported-cobalt oxide catalysts: from

528

molecule-ion associations to complete oxidation of formaldehyde. Appl. Catal. B-Environ.

529

2013, 138, 381–390.

530

(7) Jampaiah, D.; Velisoju, V. K.; Devaiah, D.; Singh, M.; Mayes, E. L.; Coyle, V. E.; Reddy,

531

B. M.; Bansal, V.; Bhargava, S. K., Flower-like Mn3O4/CeO2 microspheres as an efficient

532

catalyst for diesel soot and CO oxidation: synergistic effects for enhanced catalytic

31

ACS Paragon Plus Environment

Environmental Science & Technology

533

performance. Appl. Surf. Sci. 2019, 473, 209–221.

534

(8) Jampaiah, D.; Venkataswamy, P.; Tur, K. M.; Ippolito, S. J.; Bhargava, S. K.; Reddy, B.

535

M., Effect of MnOx loading on structural, surface, and catalytic properties of CeO2–MnOx

536

mixed oxides prepared by Sol‐Gel method. Z. Anorg. Allg. Chem. 2015, 641, (6), 1141–

537

1149.

538

(9) Zhang, S.; Fan, Q.; Gao, H.; Huang, Y.; Liu, X.; Li, J.; Xu, X.; Wang, X., Formation of

539

Fe3O4@MnO2 ball-in-ball hollow spheres as a high performance catalyst with enhanced

540

catalytic performances. J. Mater. Chem. A 2016, 4, (4), 1414–1422.

541

(10) Sang, M. P.; Jeon, S. W.; Sang, H. K., Formaldehyde oxidation over manganese–cerium–

542

aluminum mixed oxides supported on cordierite monoliths at low temperatures. Catal. Lett.

543

2014, 144, (4), 756–766.

544

(11) Quiroz Torres, J.; Royer, S.; Bellat, J. P.; Giraudon, J. M.; Lamonier, J. F., Formaldehyde:

545

catalytic oxidation as a promising soft way of elimination. ChemSusChem 2013, 6, (4),

546

578–592.

547

(12) Quiroz, J.; Giraudon, J. M.; Gervasini, A.; Dujardin, C.; Lancelot, C.; Trentesaux, M.;

548

Lamonier, J. F., Total oxidation of formaldehyde over MnOx–CeO2 catalysts: the effect of

549

acid treatment. ACS Catal. 2015, 5, 2260-2269.

550

(13) Wang, J.; Li, J.; Jiang, C.; Zhou, P.; Zhang, P.; Yu, J., The effect of manganese vacancy in

551

birnessite-type MnO2 on room-temperature oxidation of formaldehyde in air. Appl. Catal.

552

B: Environ. 2017, 204, 147–155.

553

(14) Li, H.; Huang, T.; Lu, Y.; Cui, L.; Wang, Z.; Zhang, C.; Lee, S.; Huang, Y.; Cao, J.; Ho,

554

W., Unraveling the mechanisms of room-temperature catalytic degradation of indoor

555

formaldehyde and its biocompatibility on colloidal TiO2-supported MnOx–CeO2. Environ.

32

ACS Paragon Plus Environment

Page 32 of 42

Page 33 of 42

556

Environmental Science & Technology

Sci-Nano 2018, 5, (5), 1130–1139.

557

(15) Wang, C.; Wen, C.; Lauterbach, J.; Sasmaz, E., Superior oxygen transfer ability of

558

Pd/MnOx–CeO2 for enhanced low temperature CO oxidation activity. Appl. Catal. B-

559

Environ. 2017, 206, 1–8.

560

(16) Zha, K.; Cai, S.; Hu, H.; Li, H.; Yan, T.; Shi, L.; Zhang, D., In Situ DRIFTs investigation of

561

promotional effects of tungsten on MnOx–CeO2/meso-TiO2 Catalysts for NOx Reduction. J.

562

Phys. Chem. C 2017, 121, (45), 25243–25254.

563

(17) Zhang, L.; Zhang, D.; Zhang, J.; Cai, S.; Fang, C.; Huang, L.; Li, H.; Gao, R.; Shi, L.,

564

Design of meso-TiO2@MnOx–CeOx/CNTs with a core–shell structure as DeNOx catalysts:

565

promotion of activity, stability and SO2-tolerance. Nanoscale 2013, 5, (20), 9821–9829.

566

(18) Tang, X.; Chen, J.; Li, Y.; Li, Y.; Xu, Y.; Shen, W., Complete oxidation of formaldehyde

567

over Ag/MnOx–CeO2 catalysts. Chem. Eng. J. 2006, 118, (1-2), 119–125.

568

(19) Tang, X.; Chen, J.; Huang, X.; Xu, Y.; Shen, W., Pt/MnOx–CeO2 catalysts for the complete

569

oxidation of formaldehyde at ambient temperature. Appl. Catal. B-Environ. 2008, 81, (1),

570

115–121.

571 572

(20) Miao, L.; Wang, J.; Zhang, P., Review on manganese dioxide for catalytic oxidation of airborne formaldehyde. Appl. Surf. Sci. 2018, 466, 441–453.

573

(21) Rong, S.; Zhang, P.; Liu, F.; Yang, Y., Engineering crystal facet of α-MnO2 nanowire for

574

highly efficient catalytic oxidation of carcinogenic airborne formaldehyde. ACS Catal.

575

2018, 8, (4), 3435–3446.

576

(22) Weon, S.; Choi, E.; Kim, H.; Kim, J. Y.; Park, H.-J.; Kim, S.-m.; Kim, W.; Choi, W.,

577

Active {001} facet exposed TiO2 nanotubes photocatalyst filter for volatile organic

578

compounds removal: from material development to commercial indoor air cleaner

33

ACS Paragon Plus Environment

Environmental Science & Technology

579

application. Environ. Sci. Technol. 2018, 52, (16), 9330–9340.

580

(23) Patnaik, S.; Sahoo, D. P.; Mohapatra, L.; Martha, S.; Parida, K., ZnCr2O4@ ZnO/g‐C3N4: a

581

triple‐junction nanostructured material for effective hydrogen and oxygen evolution under

582

visible light. Energy Technol. 2017, 5, (9), 1687–1701.

583

(24) Ali, A. M.; Daous, M. A.; Khamis, A. A.; Driss, H.; Burch, R.; Petrov, L. A., Strong

584

synergism between gold and manganese in an Au–Mn/triple-oxide-support (TOS) oxidation

585

catalyst. Appl. Catal. A-Gen. 2015, 489, 24–31.

586

(25) Esch, F.; Fabris, S.; Zhou, L.; Montini, T.; Africh, C.; Fornasiero, P.; Comelli, G.; Rosei,

587

R., Electron localization determines defect formation on ceria substrates. Science 2005, 309,

588

(5735), 752–755.

589

(26) Sudarsanam, P.; Hillary, B.; Amin, M. H.; Hamid, S. B. A.; Bhargava, S. K., Structure-

590

activity relationships of nanoscale MnOx/CeO2 heterostructured catalysts for selective

591

oxidation of amines under eco-friendly conditions. Appl. Catal. B-Environ. 2016, 185, 213–

592

224.

593 594 595 596

(27) Sun, C.; Li, H.; Chen, L., Nanostructured ceria-based materials: synthesis, properties, and applications. Energ. Environ. Sci. 2012, 5, (9), 8475–8505. (28) Zhang, D.; Du, X.; Shi, L.; Gao, R., Shape-controlled synthesis and catalytic application of ceria nanomaterials. Dalton T. 2012, 41, (48), 14455–14475.

597

(29) Wang, J.; Sun, J.; Jing, Q.; Liu, B.; Zhang, H.; Yongsheng, Y.; Yuan, J.; Dong, S.; Zhou,

598

X.; Cao, X., Phase stability and thermo-physical properties of ZrO2–CeO2–TiO2 ceramics

599

for thermal barrier coatings. J. Eur. Ceram. Soc. 2018, 38, (7), 2841-2850.

600

(30) Xiong, Y.; Tang, C.; Yao, X.; Zhang, L.; Li, L.; Wang, X.; Deng, Y.; Gao, F.; Dong, L.,

601

Effect of metal ions doping (M = Ti4+, Sn4+) on the catalytic performance of MnOx/CeO2

34

ACS Paragon Plus Environment

Page 34 of 42

Page 35 of 42

Environmental Science & Technology

602

catalyst for low temperature selective catalytic reduction of NO with NH3. Appl. Catal. A-

603

Gen. 2015, 495, 206–216.

604

(31) Chen, D.; He, D.; Lu, J.; Zhong, L.; Liu, F.; Liu, J.; Yu, J.; Wan, G.; He, S.; Luo, Y.,

605

Investigation of the role of surface lattice oxygen and bulk lattice oxygen migration of

606

cerium-based oxygen carriers: XPS and designed H2–TPR characterization. Appl. Catal. B-

607

Environ. 2017, 218, 249–259.

608

(32) Zhang, J.; Ohara, S.; Umetsu, M.; Naka, T.; Hatakeyama, Y.; Adschiri, T., Colloidal ceria

609

nanocrystals: a tailor‐made crystal morphology in supercritical water. Adv. Mater. 2007, 19,

610

(2), 203–206.

611

(33) Gänzler, A. M.; Casapu, M.; Boubnov, A.; Müller, O.; Conrad, S.; Lichtenberg, H.; Frahm,

612

R.; Grunwaldt, J.-D., Operando spatially and time-resolved X-ray absorption spectroscopy

613

and infrared thermography during oscillatory CO oxidation. J. Catal. 2015, 328, 216–224.

614

(34) Ramirez, A.; Hueso, J. L.; Mallada, R.; Santamaria, J., In situ temperature measurements in

615

microwave-heated gas-solid catalytic systems. Detection of hot spots and solid-fluid

616

temperature gradients in the ethylene epoxidation reaction. Chem. Eng. J. 2017, 316, 50–60.

617

(35) Abdul-Wahab, S. A.; En, S. C. F.; Elkamel, A.; Ahmadi, L.; Yetilmezsoy, K., A review of

618

standards and guidelines set by international bodies for the parameters of indoor air quality.

619

Atmos. Pollut. Res. 2015, 6, (5), 751–767.

620 621

(36) Burnett, J., Indoor air quality certification scheme for Hong Kong buildings. Indoor Built Environ. 2005, 14, (3–4), 201–208.

622

(37) Huang, Y.; Wang, W.; Zhang, Y.; Cao, J.; Huang, R.; Wang, X., Synthesis and applications

623

of nanomaterials with high photocatalytic activity on air purification. In Novel

624

Nanomaterials for Biomedical, Environmental and Energy Applications, Elsevier: 2019; pp

35

ACS Paragon Plus Environment

Environmental Science & Technology

625

299–325.

626

(38) Huang, Y.; Wang, P.; Wang, Z.; Rao, Y.; Cao, J.-j.; Pu, S.; Ho, W.; Lee, S. C., Protonated

627

g-C3N4/Ti3+ self-doped TiO2 nanocomposite films: room-temperature preparation,

628

hydrophilicity, and application for photocatalytic NOx removal. Appl. Catal. B-Environ.

629

2019, 240, 122–131.

630

(39) Bhatta, U. M.; Ross, I. M.; Sayle, T. X.; Sayle, D. C.; Parker, S. C.; Reid, D.; Seal, S.;

631

Kumar, A.; Möbus, G. n., Cationic surface reconstructions on cerium oxide nanocrystals: an

632

aberration-corrected HRTEM study. ACS Nano 2012, 6, (1), 421–430.

633

(40) Wang, J.; Zhang, P.; Li, J.; Jiang, C.; Yunus, R.; Kim, J., Room-temperature oxidation of

634

formaldehyde by layered manganese oxide: effect of water. Environ. Sci. Technol. 2015, 49,

635

(20), 12372–12379.

636

(41) Zhang, T.; Li, H.; Yang, Z.; Cao, F.; Li, L.; Chen, H.; Liu, H.; Xiong, K.; Wu, J.; Hong, Z.,

637

Electrospun YMn2O5 nanofibers: a highly catalytic activity for NO oxidation. Appl. Catal.

638

B-Environ. 2019, 247, 133–141.

639

(42) Lee, Y. N.; Lago, R. M.; Fierro, J. L. G.; Cortés, V.; Sapiña, F.; Martı́nez, E., Surface

640

properties and catalytic performance for ethane combustion of La1−xKxMnO3+δ perovskites.

641

Appl. Catal. A-Gen. 2001, 207, (1–2), 17–24.

642 643

(43) Fino, D.; Russo, N.; Saracco, G.; Specchia, V., The role of suprafacial oxygen in some perovskites for the catalytic combustion of soot. J. Catal. 2003, 217, (2), 367–375.

644

(44) Bai, B.; Qiao, Q.; Arandiyan, H.; Li, J.; Hao, J., Three-dimensional ordered mesoporous

645

MnO2-supported Ag nanoparticles for catalytic removal of formaldehyde. Environ. Sci.

646

Technol. 2016, 50, (5), 2635–2640.

647

(45) Kwon, D. W.; Seo, P. W.; Kim, G. J.; Hong, S. C., Characteristics of the HCHO oxidation

36

ACS Paragon Plus Environment

Page 36 of 42

Page 37 of 42

Environmental Science & Technology

648

reaction over Pt/TiO2 catalysts at room temperature: the effect of relative humidity on

649

catalytic activity. Appl. Catal. B-Environ. 2015, 163, 436-443.

650 651

(46) Zhang, C.; Li, Y.; Wang, Y.; He, H., Sodium-promoted Pd/TiO2 for catalytic oxidation of formaldehyde at ambient temperature. Environ. Sci. Technol. 2014, 48, (10), 5816–5822.

652

(47) Wang, Y.; Zhu, X.; Crocker, M.; Chen, B.; Shi, C., A comparative study of the catalytic

653

oxidation of HCHO and CO over Mn0.75Co2.25O4 catalyst: the effect of moisture. Appl.

654

Catal. B-Environ. 2014, 160, 542–551.

655

(48) Debono, O.; Thevenet, F.; Gravejat, P.; Hequet, V.; Raillard, C.; Lecoq, L.; Locoge, N.,

656

Toluene photocatalytic oxidation at ppbv levels: kinetic investigation and carbon balance

657

determination. Appl. Catal. B-Environ. 2016, 106, (3), 600–608.

658

(49) Weng, X.; Sun, P.; Long, Y.; Meng, Q.; Wu, Z., Catalytic oxidation of chlorobenzene over

659

MnxCe1–xO2/HZSM-5 catalysts: a study with practical implications. Environ. Sci. Technol.

660

2017, 51, (14), 8057–8066.

661

(50) Li, H.; Cui, L.; Lu, Y.; Huang, Y.; Cao, J.-j.; Park, D.; Lee, S. C.; Ho, W. K., In situ

662

intermediates determination and cytotoxicological assessment in catalytic oxidation of

663

formaldehyde: implications for catalyst design and selectivity enhancement under ambient

664

conditions. Environ. Sci. Technol. 2019, 53, (9), 5230–5240.

665

(51) Van de Vyver, S.; Odermatt, C.; Romero, K.; Prasomsri, T.; Román-Leshkov, Y., Solid

666

Lewis acids catalyze the carbon–carbon coupling between carbohydrates and formaldehyde.

667

ACS Catal. 2015, 5, (2), 972–977.

668

(52) Albonetti, S.; Blasioli, S.; Bonelli, R.; Mengou, J. E.; Scirè, S.; Trifirò, F., The role of

669

acidity in the decomposition of 1,2-dichlorobenzene over TiO2-based V2O5/WO3 catalysts.

670

Appl. Catal. A-Gen. 2008, 341, (1–2), 18–25.

37

ACS Paragon Plus Environment

Environmental Science & Technology

671

(53) Selvakumar, S.; Nuns, N.; Trentesaux, M.; Batra, V.; Giraudon, J.-M.; Lamonier, J.-F.,

672

Reaction of formaldehyde over birnessite catalyst: a combined XPS and ToF-SIMS study.

673

Appl. Catal. B-Environ. 2018, 223, 192–200.

674

(54) Islas, L.; Ruiz, J.-C.; Muñoz-Muñoz, F.; Isoshima, T.; Burillo, G., Surface characterization

675

of poly (vinyl chloride) urinary catheters functionalized with acrylic acid and poly (ethylene

676

glycol) methacrylate using gamma-radiation. Appl. Surf. Sci. 2016, 384, 135–142.

677

(55) Jia, X.; Ma, J.; Xia, F.; Xu, Y.; Gao, J.; Xu, J., Carboxylic acid-modified metal oxide

678

catalyst for selectivity-tunable aerobic ammoxidation. Nat. Commun. 2018, 9, (1), 933–939.

679

(56) Chen, H.; Tang, M.; Rui, Z.; Wang, X.; Ji, H., ZnO modified TiO2 nanotube array supported

680

Pt catalyst for HCHO removal under mild conditions. Catal. Today 2016, 264, 23–30.

38

ACS Paragon Plus Environment

Page 38 of 42

Page 39 of 42

Environmental Science & Technology

Figure 1 127x101mm (600 x 600 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 2 152x76mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 40 of 42

Page 41 of 42

Environmental Science & Technology

Figure 4 88x63mm (600 x 600 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 5 84x65mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 42 of 42