You are on page 1of 39

Accepted Manuscript

Title: Facile synthesis of Fe-modified manganese oxide with


high content of oxygen vacancies for efficient airborne ozone
destruction

Authors: Jingbo Jia, Wenjuan Yang, Pengyi Zhang, Junying


Zhang

PII: S0926-860X(17)30384-8
DOI: http://dx.doi.org/doi:10.1016/j.apcata.2017.08.013
Reference: APCATA 16368

To appear in: Applied Catalysis A: General

Received date: 7-6-2017


Revised date: 4-8-2017
Accepted date: 8-8-2017

Please cite this article as: Jingbo Jia, Wenjuan Yang, Pengyi Zhang, Junying
Zhang, Facile synthesis of Fe-modified manganese oxide with high content of
oxygen vacancies for efficient airborne ozone destruction, Applied Catalysis A,
Generalhttp://dx.doi.org/10.1016/j.apcata.2017.08.013

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Facile synthesis of Fe-modified manganese oxide with high content of

oxygen vacancies for efficient airborne ozone destruction

Jingbo Jiaa,c#, Wenjuan Yangb#, Pengyi Zhanga*, Junying Zhangd

a
State Key Joint Laboratory of Environment Simulation and Pollution Control, School

of Environment, Tsinghua University, Beijing 100084, China

b
School of Chemistry and Chemical Engineering, Yulin University, Yulin City 719000,

Shaanxi, China.

c
Key Laboratory of Advanced Materials, School of Materials Science and Engineering,

Tsinghua University, Beijing 100084, China

d
Department of Physics, Beihang University, Beijing 100191, China

#
These authors contributed equally to this work.

*Corresponding author. Tel: +86 10 62773720; Fax: +86 10 62796840-602

E-mail: zpy@tsinghua.edu.cn (P.Y. Zhang)

1
Graphical Abstract

Highlights

The Fe-modified manganese oxide was synthesized by an economic and facile one-step
hydrothermal method.
The Fe-modified manganese oxide exhibited better catalytic activity for airborne ozone
decomposition under dry and humid conditions compared to pure MnO2.
The Fe3+ partially replaced the sites of K and Mn within MnO2, leading to the phase
transformation from -MnO2 to ramsdellite MnO2 and extremely few Fe2O3.
The concentration and depletion rate of intermediate O22- formed on the surface of Fe-
MnOx during ozone destrucion was higher than that on MnO2.
The high concentration and well dispersion of oxygen vacancies on the surface of Fe-
MnOx was identified to be mainly responsible for the superior activity.

Abstract

Oxygen vacancy engineering is as an efficient strategy to improve the catalytic

performance of nanomaterials. In this work, a highly active Fe-modified manganese

oxide (Fe-MnOx) was synthesized and used for airborne ozone decomposition. The

addition of Fe3+ during MnO2 synthesis led to higher specific surface area, greatly

increased content of oxygen vacancies, evidenced by XPS, H2-TPR analysis and lower

oxygen vacancy formation energy (decreased by ~1.2 eV) based on the density

functional theory calculations. The ozone conversion over Fe-MnOx kept 97% after
2
24h reaction, while it over MnO2 slowed down to 85% under dry condition.

Remarkably, under humid condition (RH=60%), the ozone conversion over Fe-MnOx

kept 73% after 6 h reaction, while ozone conversion over pure MnO2 decreased to

50% within 1 h under the conditions of 100 ppm inlet ozone concentration and

weight space velocity of 660 L g-1 h-1. The intermediate peroxide species (O22-)

formed on the surface oxygen vacancies of Fe-MnOx and MnO2 during ozone

decomposition reaction were observed using in situ Raman spectroscopy. The

concentration and depletion rate of O22- on the surface of Fe-MnOx was higher than

that on MnO2, illustrating that O22- acted as the key species to boost the catalytic

process. The content and dispersity of oxygen vacancies were identified to be mainly

responsible for the performance difference. This provides a promising idea for

designing novel nanomaterial catalyst for gaseous ozone decomposition.

Keywords: iron modified manganese dioxide; oxygen vacancy; ozone decomposition;


in situ Raman spectroscopy; peroxide

3
1 Introduction

Ozone exposure in closed environment (buildings and aircrafts) is usually

associated with increasing morbidity and mortality rates 1. Typically, ozone source in

indoor environment includes outdoor air, ozone-generating air cleaning devices

(OACD) and imaging equipment such as photocopier and laser-jet printer. Due to the

increasing popularity of OACD, they could contribute even more ozone than outdoor

air 2, 3. The elevated ozone in aircraft cabins when airlines fly at high altitudes

originates from high-level ozone naturally in the stratosphere 4. Besides the direct

impact of ozone on human health, the effect of ozone-initiated secondary pollutions

cant be ignored 5. A number of studies in indoor environments have shown that

secondary volatile organic compounds (VOCs) 6-8 and ultrafine particles 9, 10 are

generated in the presence of ozone. The safe and efficient removal of ozone needs

to be urgently addressed.

Various methods have been widely studied for ozone decomposition, such as

adsorption, photocatalytic decomposition and room-temperature catalytic

decomposition. Among them, catalytic decomposition without need of excessive

energy has been recognized as one of the most promising method to remove ozone.

Manganese oxide (MnOx)- based materials have been used as active components for

ozone decomposition due to their economic and environmental benefits, as well as

their excellent tunable structural and physicochemical properties11-15. However, low

catalytic performance and poor humidity resistance of MnOx limits its large-scale

industrial application. In order to overcome the drawbacks, modification with


4
transition metals has been proved an effective method. Ma et al. 16 synthesized

metal-modified (cerium, cobalt, and iron) cryptomelane-type manganese oxide and

found that cerium-doped OMS-2 showed the highest ozone conversion under high

relative humidity due to its rich Mn3+ content and surface defects. Lian et al. 17

studied the modification of MnOx by adding different iron precursors and iron

content, and they found that MnFe0.5Ox-Fe(NO3)3 exhibited the best catalytic activity

for ozone decomposition. However, the motivation of metal modification and how

the surface defects affect the catalytic activity are not clear. The effect of oxygen

vacancy on catalytic activity of Fe-modified MnOx for ozone reduction is rarely

reported. Whats more, the influence of O3 and water vapor on oxygen vacancies still

remains unknown. These questions are worthy of thoroughly research and explore.

Herein, the Fe-modified manganese oxide (Fe-MnOx) was synthesized by a facile

and economic one-step hydrothermal method and its catalytic activity for ozone

elimination was investigated. The introduction of Fe3+ into MnO2 facilitated better

dispersity and higher content of oxygen vacancies, which not only favors formation

of more peroxide species, but also resists deactivation caused by water vapor. Thus,

Fe-MnOx exhibited greatly enhanced catalytic activity for ozone decomposition

under dry and humid conditions than pure MnO2. Furthermore, the density function

theory (DFT) calculations verified that oxygen vacancies would be generated more

easily on the surface of Fe-MnOx owing to lower oxygen vacancy formation energy.

2 Experimental and computational part

5
2.1. Catalyst preparation

The Fe-MnOx was synthesized by a hydrothermal method. In a typical procedure,

the manganese acetate and Fe(NO3)3 (Mn/Fe molar ratio= 20, 10, 4, 3, 2, 1) mixing

solution was added dropwise into the KMnO4 solution with the addition of

cetyltrimethyl ammonium bromide (CTAB) under continuous stirring. The slurry was

then transferred into a Teflon lined stainless steel autoclave and kept at 140C for 2

h. After being cooled to room temperature, the solid was isolated by filtration,

washed and dried at 85 C. As reference, pure MnO2 was synthesized following a

similar process without the addition of Fe(NO3)3. Commercial ozone scrubber (COS)

was purchased from Nanjing Duer Environmental Protection Company.

2.2. Catalyst characterization

Powder X-ray diffraction (XRD) patterns were obtained with an O8 discover

(Siemens) instrument equipped with Cu K radiation at accelerating voltage and

current of 40 kV and 40 mA, respectively. The atomic composition of MnO2 and Fe-

MnOx was determined using an inductively coupled plasma - optical emission

spectrometer (ICP-OES, Thermo IRIS Intrepid II XSP). The specific surface areas of the

catalysts were determined via Brunauer-Emmett-Teller (BET) model measurements

from N2 adsorption-desorption isotherms at 77 K using Quantachrome Quadra Sorb

SI analyzer. The morphology was observed with FE-SEM (S-5500, Hitachi) and HRTEM

(Tecnai G2 F30 S-TWIN). The X-ray photoelectron spectroscopy (XPS) studies were

performed using a Thermo ESCALAB250xi to obtain the chemical state of Fe and Mn.

6
H2-TPR studies were conducted on Auto ChemII 2920 (Micromeritics) instrument. In

a typical experiment, 40 mg sample was placed in a U-shaped quartz tube. After

being pretreated with helium gas for 0.5 h at 105C, the sample was reduced in the

stream of a mixture consisting of 5% H2/Ar (50 mL/min) from 40C to 600C

(5C/min). The H2 concentration was monitored by a thermal conductivity detector

(TCD).

2.3. Catalytic activity test

The ozone decomposition activity measurements were carried out in a fixed-bed

reactor (ID 6 mm) loaded with fractionized (4060 mesh) catalyst at 25C under

different relative humidity. The total flow rate through the reactor was maintained

at 1.1 L/min and the weight space velocity was set at 660 or 1320 L g-1 h-1. The ozone

concentration was continuously recorded with an ozone analyzer (Model 49i,

Thermo Scientific, USA).

2.4. Ozone decomposition intermediates analysis

The Raman spectra were acquired with a Renishaw inVia Microscope with

resolution of 2 cm-1 and an argonion laser at the wavelength of 532 nm as

excitation source. The in situ Raman experiments were conducted in a temperature-

controlled CCR1000 cell reactor system (Linkam scientific instruments) to directly

observe and identify intermediates during ozone decomposition. Approximately 60

mg of catalyst was pressed into thin wafer and mounted into the cell. The ozone was

generated by an electric field discharge ozone generator (Tonglin Technology,

7
China). If necessary, the feed gas was humidified prior to the cell reactor. The Raman

spectra was acquired under oxygen flow and under ozone/oxygen mixture flow at

different temperatures.

2.5. Computational methodology

The DFT calculations were performed to investigate oxygen vacancy formation

energy of MnO2 and Fe-MnOx, respectively. The exchange-correlation energy was in

the form of Perdew-Burke-Erzernhof (PBE) with the projected augmented wave

potentials (PAW) as implemented in the Vienna Ab initio Simulation Package (VASP)

18-21
. Kohn-Sham single electron wave functions are expanded by plane waves with

an energy cutoff 450 eV and a uniform 313 centered k-points mesh was utilized

to sample the Brillion zone. The atomic coordinates were fully relaxed using the

conjugate gradient method until the force on each atom was converged to less than

0.02 eV/ 22.

3. Results and Discussion

The preliminary experiments on Fe-Mn oxides with different initial Mn/Fe ratios

(20, 10, 4, 3, 2 and 1) revealed that among all the compositions, the sample with

Mn/Fe = 2 showed the highest activity for ozone conversion (Fig. S1).Therefore, we

focused mainly on the structure-activity relationship of this sample, named as Fe-

MnOx.

3.1. Structural and elemental composition

8
The XRD patterns of obtained MnO2 and Fe-MnOx samples are presented in Fig. 1.

It can be seen that the diffraction peaks of MnO2 occur at 12.8, 18.4, 25.8, 28.7,

37.5, 41.7, 49.3 and 59.8, which agree well with (110), (200), (220), (310), (211),

(301), (411) and (521) of tetragonal -MnO2 (JCPDS No. 44-0141). The -MnO2

consists of a series of [22] tunnels, which are formed by double chains of edge-

sharing MnO6 octahedra. Broadening of the diffraction peaks suggests that as-

synthesized -MnO2 is poor crystallized. The XRD patterns of Fe-MnOx display mixed

phases of orthorhombic ramsdellite MnO2 (JCPDS No. 43-1455) and Fe2O3 (JCPDS No.

39-0238). The diffraction peaks of Fe-MnOx occur at 21.8, 39.7, 41.8, 55.3 and

66.0, which agree well with (101), (301), (211) of ramsdellite MnO2 and (440), (622)

of Fe2O3, respectively. It is clear that after the addition of Fe ions during synthesis

procedure, the crystal growth of original -MnO2 is disturbed. During Fe cation

insertion, anisotropic expansion and contraction of unit cell parameters result in the

structural damage. Finally, -MnO2 with [2 2] tunnels transforms into ramsdellite

MnO2 with [1 2] tunnels and little Fe2O3 is formed.

In order to learn the composition of metal ions within -MnO2 and Fe-MnOx, the

contents of K, Fe and Mn were determined by ICP-OES analysis. As shown in Table 1,

the K/(Mn+Fe+K) (0.39%) and Mn/(Mn+Fe+K) (75.05%) of Fe-MnOx sample are lower

than K/(Mn+Fe+K) (7.78%) and Mn/(Mn+Fe+K) (92.22%) of pure -MnO2. This result

implies that during crystal growth process, Fe ions may enter the framework of -

MnO2 and occupy the sites of K and Mn. When the sites of K+ are partly occupied by

Fe3+, the [2 2] tunnels of -MnO2 are deformed, collapsed and reconstructed in new

9
configurations due to the smaller ionic radius of Fe3+ (0.785 ) than K+(1.65 ). At the

same time, due to the similar atomic radius of Fe3+ with Mn3+ (high spin, 0.785 ; low

spin, 0.72 ), the sites of Mn3+ would be preferable to be substituted by Fe3+.

Similarly, Ma et al. pointed out that when cerium ions added into the OMS-2

material, the substitution of both Mn ions in the framework and K in the tunnel sites

took place 16.

3.2. Textural property and morphology

The morphologies of MnO2 and Fe-MnOx were investigated by SEM and TEM. The

-MnO2 sample shown in Fig. 2 (A, B) possess disorder nanofibers structure with the

diameter of 5-20 nm and the length of 30-300 nm. When Fe3+ is introduced into the

synthetic solution, the growth behavior of Fe-MnOx is different and the morphology

of as-obtained Fe-MnOx (Fig. 2 D, E) transforms into agglomerative particles,

consisting of nanoparticles with the size of ~10-50 nm. Shen et al. also found out that

after Fe doping, the morphology of Fe-OMS-2 changed from nanoneedles to

nanorods 23. To further learn the distribution of Fe within the Fe-MnOx, elemental

mapping was conducted. The Fe and Mn mapping in Fig. 2 H and I show that Fe

element dispersed as homogeneously as Mn elements, which indicates ferric ions

homogeneously incorporated into the crystal lattice of MnO2.

The HRTEM images were obtained to further identify the microstructure of

catalysts. The well-resolved lattice fringe distance of MnO2 sample (Fig. 2 C) is 0.31

nm, which corresponds to the (310) plane of -MnO2. After the addition of Fe into

10
synthesis solution, as shown in Fig. 2F, two different lattice fringes and interface can

be clearly observed. Among them, the clear lattice fringe distance of 0.25 nm

matches with the (301) crystal plane of ramsdellite MnO2. Besides, the other fringe

distance of 0.17 nm corresponds to the plane (440) of Fe2O3. The above results

further confirm that Fe addition leads to the disruption of [2 2] tunnels of -MnO2

and formation of [1 2] tunnels of ramsdellite MnO2. That is to say, the ferric ions

partly enter the crystal lattice of MnO2 and partly form separate Fe2O3 phase. An

interface between MnO2 and Fe2O3 can be clearly observed, which is consistent with

XRD results.

The BET surface area of -MnO2 (92.9 m2/g) was considerably improved after the

addition of iron (190 m2/g). This is consistent with morphology change as indicated

with SEM. The larger surface area of Fe-MnOx can be explained by the well-

dispersion of particles and formation of smaller particles, which is beneficial for the

formation and exposure of defect sites.

3.3. Surface chemical states

XPS was employed to reveal the chemical states of Mn and Fe on the surface of as-

obtained MnO2 and Fe-MnOx samples. The XPS survey spectrum and Fe 2p, Mn 3s

and Mn 2p3/2 spectra were recorded and displayed in Fig. 3. The binding energies of

the most intense Fe 2p3/2 and Fe 2p1/2 appeared at 710.9 eV and 724.5 eV,

respectively, which corresponded to typical Fe3+ ions in an environment of oxides.

11
The weak peak located at approximately 719.2 eV was attributed to the satellite

peak of Fe 2p3/2 for Fe2O3 24.

The deconvolution analysis of Mn 2p3/2 spectra is summarized in Table 1. The

molar ratio of surface Mn3+/Mn4+ on -MnO2 was 0.87 while that on Fe-MnOx was

1.14. The enhancement of molar ratio of surface Mn3+/Mn4+ shows that more Mn4+

transformed into Mn3+ within Fe-MnOx. Oxygen vacancies would be generated to

maintain the charge balance once Mn3+ and Fe3+ existed within the MnO2. Thus,

higher amount of surface oxygen vacancies was formed as a result of the addition of

ferric ion in the synthetic solution during sample preparation.

The average oxidation state (AOS) of Mn is calculated according to the following

formula: AOS=8.956-1.126E, where E is the energy difference between the

multiplet splitting peaks of Mn 3s. As shown in Table 1, the calculated AOS of Mn in

MnO2 and Fe-MnOx were 3.48 and 3.38, respectively.

Thus, based on the K/Mn (0.12 for MnO2 and 0.0074 for Fe-MnOx) and AOS by XPS

analysis, the oxygen vacancy contents of MnO2 and Fe-MnOx are calculated according

to the following equation 4 adapted from the literature 25.

+ + (2 ) (Equation 1)

Where Ovacancy represents an oxygen vacancy site.

As shown in Table1, the oxygen vacancy content of Fe-MnOx (0.31) is much higher

than that of MnO2 (0.20). The above analysis, combined with XRD results, reveal that

12
the introduction of Fe results in the reconstruction of surface atoms and generation

of a large number of oxygen vacancies.

3.4. Temperature programmed studies

To further investigate the effect of Fe addition on the amount of oxygen

vacancies, H2-TPR and TG analysis were conducted. The H2-TPR profiles of as-

obtained -MnO2 and Fe-MnOx samples are shown in Fig. 4. As for -MnO2, the main

reduction peak with a shoulder was observed at 328C and a small peak occurred at

410C. The calculated H2 consumption of the reduction of MnO2 to MnO is 6.58

mmol/g, which is lower than the theoretical H2 consumption for reduction of MnO2

to MnO (11.50 mmol/g). This suggests that oxygen vacancies exist within the as-

prepared MnO2 material. As for Fe-MnOx, the reduction could be separated into two

sections, including the reduction of MnOx and Fe2O3. In the left section, two

overlapped reduction peaks located at 276 and 302C were attributed to the

reduction of MnOx to MnO. Usually the lower starting reduction temperature, the

more easily to be reduced. The lower starting reduction temperature of Fe-MnOx can

be ascribed to its higher content of oxygen vacancies induced by Fe3+ addition. Liu et

al. also found that the formation of more desired oxygen vacancies would result in

enhanced reducibility of nanosized ceria 26. In the right section, the latter two peaks

at 426 and 589 C were assigned to the reduction of Fe3+ to Fe. The H2 consumption

of Fe species is 18.7mmol/g, which agrees well with the Fe content in Fe-MnOx

determined by ICP-OES.

13
Thermal gravity analysis was conducted to determine the content of oxygen

vacancies within MnO2 and Fe-MnOx and the results were as shown in Fig. 5. The

weight loss below 120 C can be ascribed to physical adsorbed water, and the first

weight loss of the MnO2 was 1.41%, while it was 3.44% for Fe-MnOx. The loss in the

range 120-350 C can be attributed to the loss of chemisorbed water and surface

active oxygen species. The loss in this range can indirectly reflect the amount of

oxygen vacancies because the oxygen vacancies generally would be occupied by

water and oxygen in the atmosphere. And the corresponding second weight loss of

MnO2 and Fe-MnOx was 5.10% and 10.5%, respectively, implying that the amount of

oxygen vacancies in Fe-MnOx was greatly increased. The weight loss above 400 C

can be attributed to the lattice oxygen desorption.

3.5. DFT calculation

To further evaluate the improvement of lattice oxygen activity of Fe-MnOx

compared to that of MnO2, the oxygen vacancy formation energy of pure -MnO2

and Fe-MnOx were theoretically studied using DFT calculations. According to HRTEM

characterization, -MnO2 (310) surface plane and ramsdellite-MnO2 (301) plane are

selected as exposed crystal planes. The oxygen vacancy formation energy could be

calculated with following equation 2:

EVO Esys o Esys 1/ 2EO2 (Equation 2)

Where Esys-o and Esys are the total energies of MnO2 with and without oxygen

vacancy, respectively, and EO2 is the energy of an O2 molecule. In Fig. 6, EVO of the

14
relaxed -MnO2 structure was 3.50 eV. However, when Fe was doped into MnO2,

EVO of Fe-MnOx decreased to 2.30 eV. The lower oxygen vacancy formation energy

implies that at ambient condition the Fe-MnOx surface probably contain a significant

amount of oxygen vacancies. That is to say that the Fe-MnOx surface is predicted to

be more defective under reaction conditions, which is theoretically consistent with

the experimental XPS and H2-TPR results of Fe-MnOx.

3.6. Catalytic activity

The performance of as-synthesized -MnO2 and Fe-MnOx for dry ozone

decomposition at 25 C is presented in Fig. 7a. The Fe-MnOx exhibits excellent

activity, maintaining as high as 97% ozone conversion after 24 h reaction; while for

-MnO2, the ozone conversion decreases to ~ 85% in the same reaction period. In a

typical heterogeneous catalysis, the first step is the adsorption of reactants and the

surface area of catalyst play an important role. From Fig. 7a we can see that in the

first 4 h, the performance of -MnO2 and Fe-MnOx seem no obvious difference,

indicating that surface area is not the determining factor during ozone destruction. It

is generally regarded that catalytic ozone decomposition over metal oxides includes

three steps, i.e. dissociative adsorption of ozone, formation of peroxide

intermediate, and the decomposition of peroxide 27. Oxygen vacancies could act as

active sites for ozone adsorption and decomposition, and played a crucial role in

decomposing ozone. Herein Fe-MnOx owns larger surface area which is beneficial for

the exposure of active sites, and higher amount of oxygen vacancies which provides

more reaction sites. Therefore, Fe-MnOx shows better activity for ozone
15
decomposition than that of -MnO2. This result demonstrates a useful strategy for

improving the catalytic activity by doping a fraction of different cations in metal

oxide.

Since water vapor ubiquitously exists in air, the performance of catalysts for

ozone decomposition are also tested in humid air condition (25 C, RH 60%). As

shown in Fig. 7b, ozone conversion over Fe-MnOx in humid air still maintains 73% in 6

h, while ozone conversion over -MnO2 decreases dramatically to 50% in 1 h. Many

researchers 28-31 has observed the negative impact of water vapor on catalytic

activity for ozone decomposition over metal oxides. Depositing Pd on MnO2 or Pt on

TiO2 is an alternative way to resist the water suppression on the O3 conversion

activity 32, 33. However, due to the high cost of noble metals, researchers have been

exploring the economic and effective catalyst for high humidity ozone gas

decomposition. The above result indicates that as-synthesized Fe-MnOx provides an

alternative to noble metals for alleviating the influence of water vapor.

Furthermore, in order to highlight the performance advantages of obtained Fe-

MnOx, the activity of commercial ozone scrubber (COS) was compared in the same

conditions. The data in Fig. 7 demonstrates that Fe-MnOx shows better activity not

only in dry air but also in humid air than commercial ozone destruction catalyst.

Therefore, Fe-MnOx is expected to be a promising catalyst for ozone removal under

low or high humidity condition.

3.7. The role of oxygen vacancy

16
To gain deep insights into the effect of oxygen vacancies on ozone decomposition,

the reaction process was observed using in situ Raman spectroscopy. When the

catalysts are exposed to the O2 flow at 11C (Fig. 8 a, b), the Raman spectrum of

fresh MnO2 consists of a strong band at 579 and 640 cm-1 attributed to the vibration

mode of Ag spectroscopic species of MnO6 octahedra; while for Fe-MnOx, the peak at

579 cm-1 becomes broad and not sharp and the peak at 640 cm-1 seems invisible. It is

clear that the doped Fe inducing oxygen vacancies change the surface structure and

coordination of Mn cations.

When -MnO2 and Fe-MnOx are exposed to the dry O3 flow at 11C, a new band at

874 cm-1 appears, which is assigned to the O-O stretching vibration of peroxide

species resulting from the reaction of ozone with oxygen vacancies on the surface of

catalysts 27, 34. However, the intensity varies with temperature. The intensity of

intermediate peroxide species depended on both their formation rate (R1 and R2)

and decomposition rate (R3). As pointed out in previous study, decomposition of

peroxide (R3) is the rate-limiting step. Therefore, the intensity of peroxide species at

11C is mainly attributed to the formation of peroxide species resulted from the

reaction of ozone and oxygen vacancies on the surface of catalysts.

k1
O3 + VO O 2 + O 2- (R1)

k2
O3 + O 2- O 2 + O 2-2 (R2)

k3
O2-2 O2 + VO (R3)

17
To normalize the intensity of peroxide, the intensity ratio (I874/I579) between the

874 cm-1 and 579 cm-1 band of MnO2 and Fe-MnOx are calculated when the reaction

temperature increases from 11C to 125 C (Fig. S2). As shown in Fig. 8c, the I874/I579

on the surface of Fe-MnOx (14.9) is much higher than that of MnO2 (1.45). This

difference can be ascribed to the much higher content of oxygen vacancies over Fe-

MnOx. The ratio of I874/I579 decreases when the reaction temperature increases. This

implies that increasing temperature would accelerate the decomposition rate of

peroxide species more than the formation rate. However, the reduction rate of

I874/I579 on Fe-MnOx is faster than that on MnO2. This indicates that increasing

temperature is more favorable for peroxide decomposition on the surface of Fe-

MnOx. This result is consistent with their activities for ozone decomposition.

On the poisoning of catalyst resulted from water vapor, it is generally accepted for

a long time that water vapor gradually blocked the active sites leading to the

inhibition of ozone adsorption and decomposition on the active sites 28-30, 35. Little

research could provide strong evidence. Artur Braun reported that oxygen vacancy

could fill under water vapor in ceramic proton conductors in situ with ambient

pressure XPS 36. Oier Bikondoa et al reported that they direct visualize the water

dissociation of water on TiO2 (001) mediated by defect 37. To get insights into the

effect of water vapor, in situ Raman spectroscopy was employed to observe the

reaction process in humid gas. When the MnO2 sample is exposed to the humid

ozone (O3+H2O) flow (Fig. 7), the I874/I579 at 11C (0.37) becomes much weaker

compared to dry gas. As for the Fe-MnOx sample, the I874/I579 at 11C (3.28) also

18
becomes weaker and then reaches highest at 50C (8.85). The above results indicate

that the formation of peroxide species are obviously suppressed by water vapor,

which occupied the oxygen vacancies competitively, thus hindering the binding

between ozone and oxygen vacancies. Whats more, once water contacts with

oxygen vacancy, it would dissociate into OH species. When the oxygen vacancies are

occupied by water molecule or dissociated OH species, adsorption and

decomposition of ozone on oxygen vacancies are inhibited. Thus the performance of

Fe-MnOx under humid condition shows decreasing trend compared with that in dry

condition. However, the formation of peroxide on the surface of Fe-MnOx is less

impacted compared to MnO2. It is reasonable that oxygen vacancies disperse more

uniformly on the surface of Fe-MnOx due to its much higher specific surface area

than MnO2, which may be helpful to alleviate the negative impact of water vapor.

4 Conclusions

In this study, the catalytic activity of MnO2 for ozone removal was greatly

improved with introducing Fe3+ by generating oxygen vacancies. The obtained Fe-

MnOx catalyst contains more oxygen vacancies, which not only enables higher

activity for ozone decomposition in the dry air, but also favors resistance to water

vapor in humid ozone-containing air. This finding provides some hints to develop

efficient ozone-destruction catalysts with more oxygen vacancies via doping

transitional metals, and deepens the understanding of the role of oxygen vacancy

during ozone decomposition.

19
Acknowledgments

This study was funded by National Nature Science Foundation of China (Nos.

21521064, 51651201) and Suzhou-Tsinghua Innovation Guiding Program

(2016SZ0104).

20
References

[1] C. Chen, B. Zhao, C.J. Weschler, Environ. Health Persp. 120 (2012) 235-240.

[2] M.O. Fadeyi, Sustain. Cities Soc. 18 (2015) 78-94.

[3] K.-P. Yu, G.W.-M. Lee, C.-P. Hsieh, C.-C. Lin, Atmos. Environ. 45 (2011) 35-42.

[4] S. Bhangar, S.C. Cowlin, B.C. Singer, R.G. Sextro, W.W. Nazaroff, Environ. Sci.

Technol. 42 (2008) 3938-3943.

[5] C.J. Weschler, Environ. Health Persp. 114 (2006) 1489-1496.

[6] C. Weisel, C.J. Weschler, K. Mohan, J. Vallarino, J.D. Spengler, Environ. Sci.

Technol. 47 (2013) 4711-4717.

[7] B.K. Coleman, H. Destaillats, A.T. Hodgson, W.W. Nazaroff, Atmos. Environ. 42

(2008) 642-654.

[8] S. Yang, K. Gao, X. Yang, Build. Environ. 103 (2016) 146-154.

[9] C.C.D. Fung, S. Shu, Y. Zhu, Indoor Air 24 (2014) 503-510.

[10] A.C. Rai, C.-H. Lin, Q. Chen, Atmos. Environ. 102 (2015) 145-155.

[11] C. Wang, J. Ma, F. Liu, H. He, R. Zhang, J. Phys. Chem. C 119 (2015) 23119-

23126.

[12] R. R., O.S. T., C. Jingguang, A. K., J. Phys. Chem. B 105 (2001) 4245-4253.

[13] C. Jiang, P. Zhang, B. Zhang, J. Li, M. Wang, Ozone Sci. Eng. 35 (2013) 308-315.

[14] M. Wang, P. Zhang, J. Li, C. Jiang, Chin. J. Catal. 35 (2014) 335-341.

[15] J. Jia, P. Zhang, L. Chen, Catal. Sci. Technol. 6 (2016) 5841-5847.

[16] J. Ma, C. Wang, H. He, Appl. Catal. B Environ. 201 (2017) 503-510.

[17] Z. Lian, J. Ma, H. He, Catal. Comm. 59 (2015) 156-160.

21
[18] J.P. Perdew, K. Burke, M. Ernzerhof, Phy. Rev. Lett. 77 (1996) 3865-3868.

[19] P.E. Blochl, Phys. Rev. B 50 (1994) 17953-17979.

[20] G. Kresse, J. Furthmuller, Comput. Mater. Sci. 6 (1996) 15-50.

[21] G. Kresse, J. Furthmuller, Phys. Rev. B 54 (1996) 11169-11186.

[22] William, H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical

Recipes 3rd Edition: The Art of Scientific Computing, Cambridge University

Press, New York, 2007.

[23] X. Shen, A.M. Morey, J. Liu, Y. Ding, J. Cai, J. Durand, Q. Wang, W. Wen, W.A.

Hines, J.C. Hanson, J. Bai, A.I. Frenkel, W. Reiff, M. Aindow, S.L. Suib, J. Phy.

Chem. C 115 (2011) 21610-21619.

[24] T. Yamashita, P. Hayes, Appl. Surf. Sci. 254 (2008) 2441-2449.

[25] J. Hou, Y. Li, M. Mao, L. Ren, X. Zhao, ACS Appl. Mater. Inter. 6 (2014) 14981-

14987.

[26] X. Liu, K. Zhou, L. Wang, B. Wang, Y. Li, J. Am. Chem. Soc. 131 (2009) 3140-

3141.

[27] W. Li, G.V. Gibbs, S.T. Oyama, J. Am. Chem. Soc. 120 (1998) 9041-9046.

[28] H. Einaga, M. Harada, S. Futamura, Chem. Phys. Lett. 408 (2005) 377-380.

[29] W.T. Tsai, C.Y. Chang, F.H. Jung, C.Y. Chiu, W.H. Huang, Y.H. Yu, H.T. Liou, Y.

Ku, J.N. Chen, C.F. Mao, J. Environ. Sci. Health A Tox. Hazard. Subst. Environ.

Eng. 33 (1998) 1705-1717.

[30]. C.A. Liu, D.Z. Sun, H. Wang, W. Li, J. Environ. Sci. 15 (2003) 779-782.

[31] H. Huang, X. Ye, W. Huang, J. Chen, Y. Xu, M. Wu, Q. Shao, Z. Peng, G. Ou, J.

22
Shi, X. Feng, Q. Feng, H. Huang, P. Hu, D.Y.C. Leung, Chem. Eng. J. 264 (2015)

24-31.

[32] T. Kameya, K. Urano, J. Environ. Eng. Asce, 128 (2002) 286-292.

[33] C. Ren, L. Zhou, H. Shang, Y. Chen, Chin. J. Catal. 35 (2014) 1883-1891.

[34] J. Jia, P. Zhang, L. Chen, Appl. Catal. B Environ. 189 (2016) 210-218.

[35] H. Einaga, S. Futamura, J. Catal. 227 (2004) 304-312.

[36] Q. Chen, F. El Gabaly, F.A. Akgul, Z. Liu, B.S. Mun, S. Yamaguchi, A. Braun,

Chem. Mater. 25 (2013) 4690-4696.

[37] O. Bikondoa, C.L. Pang, R. Ithnin, C.A. Muryn, H. Onishi, G. Thornton, Nat.

Mater. 5 (2006) 189-192.

23
Figure captions

Fig.1.XRD patterns of MnO2 and Fe-MnOx catalysts (Symbols , and represent


the characteristic peaks of tetragonal -MnO2, ramsdelliteMnO2 and Fe2O3,
respectively).
Fig. 2. SEM and TEM images of MnO2 (A-C), Fe-MnOx (D-G) and elemental mapping
(H-I) of rectangle areas in the F.
Fig. 3.XPS survey spectrum (a) and Fe 2p (b), Mn 2p3/2(c), Mn 3s (d) spectrumof MnO2
and Fe-MnOx.
Fig. 4. H2-TPR profiles of MnO2 and Fe-MnOx.
Fig. 5.TG curves of MnO2 and Fe-MnOx.
Fig. 6.The relaxed structure of MnO2 and Fe-MnOx with oxygen vacancy.
Fig. 7.Performance of different catalysts: MnO2, Fe-MnOx and commercial ozone
scrubber (COS) for decomposing ozone in dry (a) and humid gas (b).
Fig. 8. In situ Raman spectra of MnO2 (a) and Fe-MnOx (b) exposed to dry and humid
O2 and O3/O2 flow at 11oC. And the intensity ratio (I874/I579) of MnO2 and Fe-MnOx vs.
reaction temperature (c).
Fig. 9. Schematic representation of reaction between O3 and Fe-MnOx under dry or
humid condition.

24
Fig. 1. XRD patterns of MnO2 and Fe-MnOx catalysts (Symbols
, and represent the
characteristic peaks of tetragonal -MnO2, ramsdellite MnO2 and Fe2O3, respectively).

25
Fig. 2. SEM and TEM images of MnO2 (A-C), Fe-MnOx (D-G) and elemental mapping (H-I) of
rectangle areas in the F.

26
Fig. 3. XPS survey spectrum (a) and Fe 2p (b), Mn 2p3/2 (c), Mn 3s (d) spectrum of MnO2 and
Fe-MnOx.

27
Fig. 4. H2-TPR profiles of MnO2 and Fe-MnOx.

28
Fig. 5. TG curves of MnO2 and Fe-MnOx.

29
F

Fig. 6. The relaxed structure of MnO2 and Fe-MnOx with oxygen vacancy.

30
Fig. 7. Performance of different catalysts: MnO2, Fe-MnOx and commercial ozone scrubber
(COS) for decomposing ozone in dry (a) and humid gas (b).

31
32
Fig. 8. In situ Raman spectra of MnO2 (a) and Fe-MnOx (b) exposed to dry and humid O2 and
O3/O2 flow at 11oC. And the intensity ratio (I874/I579) of MnO2 and Fe-MnOx vs. reaction
temperature (c).

33
Fig. 9. Schematic representation of reaction between O3 and Fe-MnOx under dry or humid
condition.

34
35
Table captions

Table 1 BET surface area, element composition (ICP-OES) and XPS results of MnO2

and Fe-MnOx.

36
Table 1 BET surface area, element composition (ICP-OES) and XPS results of MnO2 and Fe-
MnOx.

BET ICP-OES XPS


surface AOS Oxygen
Catalysts Mn/(Mn+Fe+K K/(Mn+Fe+K Fe/(Mn+Fe+K vacancy
area Mn3+/Mn4+ of
) (%) ) (%) ) (%) content
2
(m /g) Mn
MnO2 92.9 92.22 7.78 \ 0.87 3.48 0.20
Fe-MnOx 190 75.05 0.39 24.56 1.14 3.38 0.31

37
38

You might also like