Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

14
Mn-Modied CuO, CuFe 2 O 4 , and γFe 2 O 3 Three-Phase Strong Synergistic Coexistence Catalyst System for NO Reduction by CO with a Wider Active Window Xiaobing Shi, Bingxian Chu, Fan Wang, Xiaoling Wei, Lixia Teng, Minguang Fan, Bin Li,* ,Lihui Dong,* ,,and Lin Dong Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensication Technology, School of Chemistry and Chemical Engineering, Guangxi University, Nanning 530004, PR China School of the Environment, Jiangsu Key Laboratory of Vehicle Emissions Control, Center of Modern Analysis, Nanjing University, Nanjing 210093, PR China * S Supporting Information ABSTRACT: A series of samples with the precursors molar ratio of {KMn 8 O 16 }/{CuFe 2 O 4 } = 0, 0.008, 0.010, 0.016, and 0.020 were successfully synthesized for selective catalytic reduction of NO by CO. The physicochemical properties of all samples were studied in detail by combining the means of X-ray photoelectron spectroscopy, H 2 -temperature-pro- grammed reduction, scanning electron microscopy mapping, X-ray diraction (XRD), N 2 physisorption (BrunauerEmmettTeller), NO + CO model reaction, and in situ Fourier transform infrared spectroscopy techniques. The results show that three phases of γ-Fe 2 O 3 , CuFe 2 O 4 , and CuO, which have strong synergistic interaction, coexist in this catalyst system, and dierent phases play a leading role in dierent temperature ranges. Mn species are highly dispersed in the three-phase coexisting system in the form of Mn 2+ , Mn 3+ , and Mn 4+ . Because of the strong interaction between Mn 2+ and Fe species, a small amount of Cu 2+ precipitates from CuFe 2 O 4 and grows along the CuO(110) plane, which has better catalytic performance. Mn 3+ can inhibit the conversion of γ-Fe 2 O 3 to α- Fe 2 O 3 at high temperature and then increases the high-temperature activity. The synergistic eect between Mn 4+ and the surfaces of three phases generates active oxygen species Cu 2+ OMn 4+ and Mn 4+ OFe 3+ , which can be more easily reduced to some synergistic oxygen vacancies during the reaction. Furthermore, the formed synergistic oxygen vacancies can promote the dissociation of NO and are also propitious to the transfer of oxygen species. All of these factors make the appropriate manganese-modied three-phase coexisting system have better catalytic activity than the manganese-free catalyst, making NO conversion rate reach 100% at around 250 °C and maintain to 1000 °C. Combining comprehensive analysis of various characterization results and in situ infrared as well as XRD results in the equilibrium state, a new possible NO + CO model reaction mechanism was temporarily proposed to further understand the catalytic processes. KEYWORDS: three-phase coexisting, Mn-modied, NO reduction by CO, broad temperature window, in situ DRIFTS 1. INTRODUCTION NO x species from fuel combustion can cause serious problems, such as lung bronchial disease, greenhouse eect, ozone depletion, acid rain, and photochemical smog. These can have an immeasurable impact on our living environment, health, and national economy. As the emission regulation of NO x becomes more and more stringent, the reduction of NO x has caused widespread concern around the world. 1 Though the technology of selective catalytic reduction of NO x with NH 3 has been widely used for NO x reduction, its disadvantages cannot be ignored, 2,3 such as air preheater blocks and NH 3 leakage. The catalytic removal of NO x by CO is probably one of the most promising methods to meet the current requirements in various deNO x technologies because the internal combustion engines using fossil fuels often produce the two gases at the same time, and this deNO x technology can achieve the purpose of eliminating two kinds of harmful gases at the same time. 2 Many researchers have spent a lot of energy studying the catalytic performance and mechanism of various noble metal catalysts in the process of carbon monoxide decomposing NO x , such as Pd, Rh, and Pt supported on some carrier materials. These catalysts have high catalytic activity for the reduction of NO by CO, but their scarcity, high-temperature instability, high price, and easy toxicity have greatly limited their applications. As a result, cheap alternative materials are becoming more and more concerned by researchers. In recent years, many studies Received: August 3, 2018 Accepted: October 29, 2018 Published: October 29, 2018 Research Article www.acsami.org Cite This: ACS Appl. Mater. Interfaces 2018, 10, 40509-40522 © 2018 American Chemical Society 40509 DOI: 10.1021/acsami.8b13220 ACS Appl. Mater. Interfaces 2018, 10, 4050940522 Downloaded via NANJING UNIV on January 2, 2019 at 06:33:45 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Transcript of Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

Page 1: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

Mn-Modified CuO, CuFe2O4, and γ‑Fe2O3 Three-Phase StrongSynergistic Coexistence Catalyst System for NO Reduction by COwith a Wider Active WindowXiaobing Shi,† Bingxian Chu,† Fan Wang,† Xiaoling Wei,† Lixia Teng,† Minguang Fan,† Bin Li,*,†

Lihui Dong,*,†,‡ and Lin Dong‡

†Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensification Technology, School of Chemistry andChemical Engineering, Guangxi University, Nanning 530004, PR China‡School of the Environment, Jiangsu Key Laboratory of Vehicle Emissions Control, Center of Modern Analysis, Nanjing University,Nanjing 210093, PR China

*S Supporting Information

ABSTRACT: A series of samples with the precursor’s molarratio of {KMn8O16}/{CuFe2O4} = 0, 0.008, 0.010, 0.016, and0.020 were successfully synthesized for selective catalyticreduction of NO by CO. The physicochemical properties ofall samples were studied in detail by combining the means ofX-ray photoelectron spectroscopy, H2-temperature-pro-grammed reduction, scanning electron microscopy mapping,X-ray diffraction (XRD), N2 physisorption (Brunauer−Emmett−Teller), NO + CO model reaction, and in situFourier transform infrared spectroscopy techniques. Theresults show that three phases of γ-Fe2O3, CuFe2O4, andCuO, which have strong synergistic interaction, coexist in this catalyst system, and different phases play a leading role indifferent temperature ranges. Mn species are highly dispersed in the three-phase coexisting system in the form of Mn2+, Mn3+,and Mn4+. Because of the strong interaction between Mn2+ and Fe species, a small amount of Cu2+ precipitates from CuFe2O4and grows along the CuO(110) plane, which has better catalytic performance. Mn3+ can inhibit the conversion of γ-Fe2O3 to α-Fe2O3 at high temperature and then increases the high-temperature activity. The synergistic effect between Mn4+ and thesurfaces of three phases generates active oxygen species Cu2+−O−Mn4+ and Mn4+−O−Fe3+, which can be more easily reducedto some synergistic oxygen vacancies during the reaction. Furthermore, the formed synergistic oxygen vacancies can promotethe dissociation of NO and are also propitious to the transfer of oxygen species. All of these factors make the appropriatemanganese-modified three-phase coexisting system have better catalytic activity than the manganese-free catalyst, making NOconversion rate reach 100% at around 250 °C and maintain to 1000 °C. Combining comprehensive analysis of variouscharacterization results and in situ infrared as well as XRD results in the equilibrium state, a new possible NO + CO modelreaction mechanism was temporarily proposed to further understand the catalytic processes.

KEYWORDS: three-phase coexisting, Mn-modified, NO reduction by CO, broad temperature window, in situ DRIFTS

1. INTRODUCTION

NOx species from fuel combustion can cause serious problems,such as lung bronchial disease, greenhouse effect, ozonedepletion, acid rain, and photochemical smog. These can havean immeasurable impact on our living environment, health, andnational economy. As the emission regulation of NOx becomesmore and more stringent, the reduction of NOx has causedwidespread concern around the world.1 Though the technologyof selective catalytic reduction of NOxwith NH3 has been widelyused for NOx reduction, its disadvantages cannot be ignored,

2,3

such as air preheater blocks and NH3 leakage. The catalyticremoval of NOx by CO is probably one of the most promisingmethods to meet the current requirements in various deNOxtechnologies because the internal combustion engines usingfossil fuels often produce the two gases at the same time, and this

deNOx technology can achieve the purpose of eliminating twokinds of harmful gases at the same time.2

Many researchers have spent a lot of energy studying thecatalytic performance and mechanism of various noble metalcatalysts in the process of carbon monoxide decomposing NOx,such as Pd, Rh, and Pt supported on some carrier materials.These catalysts have high catalytic activity for the reduction ofNO by CO, but their scarcity, high-temperature instability, highprice, and easy toxicity have greatly limited their applications. Asa result, cheap alternative materials are becoming more andmore concerned by researchers. In recent years, many studies

Received: August 3, 2018Accepted: October 29, 2018Published: October 29, 2018

Research Article

www.acsami.orgCite This: ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

© 2018 American Chemical Society 40509 DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

Dow

nloa

ded

via

NA

NJI

NG

UN

IV o

n Ja

nuar

y 2,

201

9 at

06:

33:4

5 (U

TC

).

See

http

s://p

ubs.

acs.

org/

shar

ingg

uide

lines

for

opt

ions

on

how

to le

gitim

atel

y sh

are

publ

ishe

d ar

ticle

s.

Page 2: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

have focused on the development of transition metals and rareearth complex oxides,2,3 especially cobalt, copper, nickel, andiron oxides, which have been reported to have NO + COreactivity activities.3 In these transition-metal oxides, iron-containing catalysts have been widely studied because of theirlow cost, extensive source, environmental friendliness, stability,and durability.4

Fe-based materials provide promising results for variouscatalytic reactions. For example, Ce−Mn/TiO2 catalystsmodified by iron have catalytic performance for the selectivereduction of NO by CO.5 Previous literature reported thatoxides containing some iron species provide catalytic activity inthe NO elimination.6−8 Study9 reported that γ-Fe2O3 has bettercapacity to reduce NO than α-Fe2O3 because of the more acidicsites on the surface, and magnetic γ-Fe2O3 has a relatively strongmagnetic adsorption capacity for NO molecules, therebypromoting the chemical adsorption of NO onmagnetic surfaces.It also has better magnetic susceptibility which is slightlyreduced after being used, and the catalyst still has the potentialfor magnetic recycling and reuse,6 but the defect is that thethermal stability is worse and more or less converted to α-Fe2O3under high-temperature conditions, resulting in a decrease incatalytic activity.9 Magnetite Fe3O4 is a kind of ferromagneticmaterial with a cubic antispine structure at room temperature,which has attracted considerable attention because of itselectromagnetic properties and a reactive structure for the COshift reaction.10 When Fe3O4 and γ-Fe2O3 coexist, they willexhibit better stability, magnetic properties, and catalyticproperties.11 However, one extremely important point cannotbe ignored that the efficiency of mono iron oxides at lowtemperature (<300 °C) is not high.11,12

Catalysts that contain CuO species have been identified asactive substances for the reduction of NO by CO.12,13 However,it was found that single copper oxide still exhibited a lower NOreduction activity at low temperatures.13,14 It is known that thepossible double exchange behavior between Mn3+ and Mn4+ inmanganese species can produce a large amount of surface activeoxygen exhibiting unique catalytic redox properties, which hasattracted a lot of attention in the study of NO reduced by the COmodel reaction.14

In this work, we successfully used a simple new method toprepare a synergistic three-phase coexistence catalyst system (γ-Fe2O3, CuFe2O4, and CuO modified by some Mn species).There is a strong synergy between these phases, and differentphase species take turn to play a leading role in differenttemperature ranges, resulting in considerable catalytic perform-ance for NO reduction by CO at low and high temperatures(100−1000 °C). This catalyst can effectively avoid the abovedisadvantages discussed above. Combining comprehensiveanalysis of various characterization results and in situ infraredas well as X-ray diffraction (XRD) results in the equilibriumstate, the reason for low-temperature considerable activity, high-temperature stability, and a new possible NO + CO modelreaction mechanism in different temperature regions weretemporarily proposed.

2. EXPERIMENTAL SECTION2.1. Preparation of Catalysts. CuFe2O4 magnetic nanoparticle

precursor was obtained by the method similar to the typical preparationreported in the literature.15 The detailed synthesis process is describedbelow: first, 2.5 mmol CuCl2·2H2O (0.426 g) and 5 mmol FeCl3·6H2O(1.351 g) were dissolved in ethylene glycol (40 mL). Subsequently, 1.0g of polyethylene glycol and 3.6 g of CH3COONa were added to the

mixture solution obtained above, and then the mixture was mixed atroom temperature and under magnetic stirring at a suitable speed for 1h, and thus, a green mixture system was observed. At last, a 100 mLTeflon-lined stainless-steel autoclave was used to fill the solutionobtained, and then it was placed in an oven at 200 °C for 12 h. Theresulting precipitated material was centrifuged and washed three timeswith deionized water and absolute ethanol and then dried for 12 h at 80°C.

KMn8O16 precursor was obtained by a typical method,16 22.05 gmanganese acetate tetrahydrate and 9.84 g potassium permanganatemixture were ground in an agate mill until homogeneous, and thepowder was then stored in a beaker and heated to 80 °C for 4 h. Afterwashing with deionized water to remove the unreacted material, a blackpowder was obtained, which was dried overnight at 80 °C.

The two precursors were chosen according to different molar ratios,and they were ground evenly in an agate mortar and then put in a mufflefurnace and calcined for 10 h at 400 °C.15 A series of samples with theprecursor’s molar ratio of {KMn8O16}/{CuFe2O4} = 0, 0.008, 0.010,0.016, and 0.020 were labeled as 0MnFeCu, 0.008MnFeCu, 0.010MnFeCu,0.016MnFeCu, and 0.020MnFeCu, respectively.

2.2. Catalysts Characterization. The X’Pert PRO diffractometer(PANalytical, The Netherlands) using Cu Kα radiation (λ = 1.54060Å) was used to obtain the powder XRD patterns of the samples. Thescan voltage and current were set to 40 kV and 40 mA, respectively, thescan rate was 8°·min−1, and the scan range 2θ was 10−80°.

The adsorption−desorption isotherm of N2 at 77 K was obtainedusing aMicrometrics TriStar I 3020 analyzer. The pore distribution andspecific surface area were calculated from the nitrogen sorptionisotherms by the Barrett−Joyner−Halenda and Brunauer−Emmett−Teller (BET) methods, respectively.

A HITACHI S-3400N electron microscope (Hitachi Company ofJapan) was used to measure the scanning electron microscopy (SEM)images of the samples at 20 kV. Before the field-emission SEM test, thesamples were suspended in ethanol, dispersed by ultrasonic, anddripped onto the aluminum sheet.

A Tecnai G2 F20 S-TWIN instrument (FEI Company America)with 200 kV acceleration voltage was used to obtain the transmissionelectron microscopy (TEM) images.

The X-ray photoelectron spectroscopy (XPS) spectra of the sampleswere obtained by using an ESCALAB 250Xi multifunctional imagingelectron spectrometer (Thermo Fisher Company America), which usedmonochromatic Al Kα radiation (hν = 1486.6 eV) at the power level of150 W. The electronic binding energy is calibrated on the basis of C 1s(284.8 eV). The irradiated area and detection depth of the samples were2 mm × 1 mm and 2−5 nm, respectively.

The H2-temperature-programmed reduction (TPR) profiles of thesamples were obtained by using a FINESORB-3010 automaticchemical adsorption apparatus (Finetec Corporation). Sample (20mg) was heated from room temperature to 110 °C (which was keptunder these conditions for 1 h before analysis) under a flow rate of 50mL·min−1 of N2 and was then switched to 10 mL·min−1 H2−Ar flow(7% H2 by volume) for 30 min after being cooled to room temperaturein a N2 atmosphere. After that, the temperature was raised from roomtemperature to 900 °C (10 °C·min−1), and the thermal conductivitydetector (TCD) was used to continuously analyze the H2 consumption.

The Nicolet iS50 Fourier transform infrared spectrometer equippedwith mercury cadmium telluride and cooled with liquid nitrogen, with aspectral resolution of 4 cm−1 (number of scans = 32), was used to obtainthe in situ diffuse reflectance infrared Fourier transform spectrum(DRIFTS) from 650 to 4000 cm−1. The catalyst powder was placed in asample cell and pretreated with purified N2 for 1 h at 400 °C toeliminate adsorbed impurities on the surface, and the catalystbackground spectra at different target temperatures were collectedduring cooling to room temperature. Subsequently, the catalyst wasrespectively exposed to 10 mL·min−1 NO−He (5 vol % NO and 95 vol% He) or 10 mL·min−1 CO−He (10 vol % CO and 90 vol % He) or 10mL·min−1 NO−He (5 vol % NO and 95 vol % He) + 10 mL·min−1

CO−He (10 vol % CO and 90 vol % He) gaseous flow until saturation.The Fourier transform infrared spectroscopy (FT-IR) of CO, CO +NO, and NO at each target temperature was collected at the heating

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40510

Page 3: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

rate of 5 °C·min−1 from 50 to 400 °C. Finally, the correspondingbackground reference spectra were subtracted at various targettemperatures to get the results.2.3. Catalytic Activity Measurements. The performance of the

catalyst for NO reduced by CO was obtained under simulated reactionconditions. The reaction conditions include feed steam composed ofHe−NO (95 vol %He and 5 vol %NO) andHe−CO (90 vol %He and10 vol % CO), 10% H2O (only when used), and 1000 ppm SO2 (onlywhen used) andmake the flow rate of the mixed gas flowing through thesample constant at 150 000 mL·g−1·h−1 (GHSV). Each sieved sample(15mg, 40−60mesh) was packed in a quartz tube and pretreated at 110°C in a high-purity N2 flow for 1 h to remove impurities. The reactiongas mixture is switched after the sample is cooled to room temperature.Two chromatographic columns (diameter = 3 mm, length = 1.75 m)and two TCDs (T = 100 °C) were used to analyze the catalytic reactionproducts. Data collection begins when the outlet concentration of thegas reaches a steady state. The NO conversion and N2 selectivity arecalculated as follows:12

=[ ] − [ ]

[ ]×

=× [ ]

[ ] − [ ]×

= [ ] × [ ]

NONO NO

NO100%

N2 N

NO NO100%

Conversion of NO to N NO N

conversionin out

in

2 selectivity2 out

in out

2 conversion 2 selectivity

ikjjjjj

y{zzzzz

3. RESULTS AND DISCUSSION

3.1. Catalytic Performance. The NO conversion, N2selectivity, and conversion of NO to N2 over a series of catalystsand the CO conversion on the catalyst 0.010MnFeCu as a functionof temperature are shown in Figure 1. From Figure 1a, it can beseen that the NO conversion on all catalysts is almost the same attemperatures below 150 °C, and it is relatively low. When thetemperature rises from 150 to 200 °C, the catalytic conversion ofthe more Mn-containing species is significantly higher thanthose of 0MnFeCu and 0.008MnFeCu, indicating that theincorporation of the appropriate amount of Mn species furtherenhances the performance of these catalysts. As the temperaturerises from 200 to 230 °C, the NO conversion increasesdramatically, which may be due to the surface of the catalystreaching the best stable catalytic state, and each reactant reactsquickly on the surface of the catalyst. When the temperaturecontinues to increase to 350 °C, the conversion of NO graduallyincreases to 100%. When the temperature eventually rises to1000 °C, the NO conversion rate remains constant at 100%.This shows that the catalyst not only has a good NO conversionat low temperatures but also exhibits excellent stability at hightemperatures. Figure 1b shows the selectivity of N2 of allsamples. It was found that all catalysts show the similar curve andthat N2 selectivity decreases sharply to a minimum with the

Figure 1. Results of (a) NO conversion (%); (b) N2 selectivity (%); (c) conversion of NO to N2 (%) over various catalysts as a function of reactiontemperatures; and (d) NO conversion or CO to CO2 conversion of the sample 0.010MnFeCu. Reaction conditions: 5% NO, 10% CO, and balanced He;SV = 150 000 mL·g−1·h−1. (e) Conversion of NO to N2 (%) over the catalyst 0.010MnFeCu at different GHSV= 30 000, 150 000, 300 000 mL·g−1·h−1.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40511

Page 4: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

temperature increasing from 150 to 200 °C and then quicklyrises again to 100% around 300 °C. However, the 0.010MnFeCucatalyst has higher selectivity in the whole temperature region,which suggests that the appropriate amount of Mn manganeseincorporation can further improve the selectivity of theproducts.17 From Figure 1c, it can be seen that when thetemperature rises from 100 to 200 °C, the conversions of NO toN2 of 0MnFeCu, 0.008MnFeCu, and 0.020MnFeCu catalysts remain at alower value. However, those of 0.010MnFeCu and 0.016MnFeCusamples were much higher, which indicated that the appropriateamount of Mn manganese incorporation can improve thecatalytic performance. As can be seen from Figure 1d, theconversion rate of NO and CO at 200−220 and 220−230 °Cincreased drastically compared to those at other temperatureranges, which may be due to the rapid formation of some activecatalytic species on the catalyst surface. In addition, it can also beobserved that the curve of the conversion of CO and NO isapproximately the same, but their values rise to 50 and 100%,respectively, at 300 °C, which indicates that NO and CO reach adynamic equilibrium on the catalyst.From Figure 1e, it can be known that when the space velocity

is 30 000 mL·g−1·h−1, the N2 generation rate reaches 100% at230 °C and stabilizes to 1000 °C. Although the 100% NOconversion temperature increases with the increase of the spacevelocity, because of the high-temperature activity stability of thecatalyst, NO can be completely eliminated as long as thetemperature is sufficiently high. In order to further explore thereasons for the differences in catalytic performance, weperformed a series of characterizations of these samples andpresented the corresponding results in the following sections.The effects of SO2, SO2 +H2O, and H2O on the activity of the

0.010MnFeCu catalyst at 320 °C were also studied (shown inFigure S1). The results show that the catalyst has a certain waterand sulfur resistance.

3.2. Texture Characterization (N2 Physisorption). Thecorresponding pore size distribution of these catalysts shows anarrow peak centered at around 16−30 nm, which is in goodconsistent with the pore distribution region of mesopores(Figure S2), indicating that the samples have uniformmesoporous size distribution. The summary of the BET surfacearea, pore size, and pore volume of these catalysts is shown inTable S1. It can be seen from Table S1 that the surface area andpore volume of the samples have no obvious change, which mayindicate that the influence of the deposition of trace Mn oxideswill be insignificant to the structure of all samples. Figure S3shows that the average size of the particles is about 10−30 nm.

3.3. XPS Results.The XPS spectra for O 1s of all samples arealso shown in Figure 2a, and three individual peaks of all sampleswere distinguished. The peak observed at around 529.5 eVcorresponds to the lattice oxygen O2− (denoted as Oβ). Thesecond peak having a higher binding energy with the position ofaround 531.1 eV (denoted as Oα) is assigned to the chemicaladsorbed oxygen (O−, O2

2−), and the binding energy occurringat 533.1 eV (denoted as Oγ) is contributed to the physicaladsorbed oxygen (H2O, OH

−).18,19 Surface adsorbed oxygenOα

is generally considered to be more reactive in oxidation becauseit has a higher mobility than the lattice oxygen Oβ,

20 it has beenwidely reported that the surface chemical oxygen has a positiveeffect on the catalytic effect of NO and CO reaction, and thismechanism may be attributed to the fact that it can not onlyoxidize CO, but also promote the adsorption of NO on activesites.19,21 The higher Oα/O(α+β+γ) value means that there is moreactive oxygen on the surface of the catalyst.22,23 The 0.010MnFeCusample has the highest content (0.46), which is in line with itsbetter performance, and the other three samples are 0.07, 0.40,and 0.32, respectively. According to the XPS spectrum, whenproper trace manganese oxides were added to catalysts, chemicaloxygen increased and then the redox performance was

Figure 2. (a) XPS results of O 1s of 0MnFeCu, 0.010MnFeCu, 0.020MnFeCu, and KMn8O16 catalysts; (b) XPS results of Cu 2p of 0MnFeCu, 0.010MnFeCu, and0.020MnFeCu catalysts; (c) XPS results of Fe 2p of 0MnFeCu, 0.010MnFeCu, and 0.020MnFeCu catalysts; and (d) XPS analysis of Mn 2p in 0.010MnFeCu,0.020MnFeCu catalyst, and KMn8O16.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40512

Page 5: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

promoted, which is in good consistent with the results of theactivities.In Figure 2b, the two main peaks at 933.6−933.9 and 952.7−

953.7 eV are classified as Cu 2p3/2 and Cu 2p1/2, respectively.20

One can clearly find that the peaks of Cu 2p3/2 and Cu 2p1/2 haveobvious vibration at the binding energies of about 944 and 963eV assigned to shakeup satellites, which are similar to theprevious literature,12,24−26 and the position and shape trend ofthe peak are also exactly the same as that of pure CuO.Therefore, all the above analysis shows that Cu species in all ofour catalysts exist in the form of Cu2+.All Fe 2p spectra in Figure 2c show two main peaks with the

binding energies of 710.7−711.1 and 724.2−724.3 eV, whichbelong to Fe 2p3/2 and Fe 2p1/2, respectively. One can clearlyfind that two accompanying satellite peaks at the bindingenergies of about 719.3−719.2 and 732.7−732.8 eV are thecharacteristic of Fe3+ cations, which is quite similar to thesereports,27−29 which showed that the Fe 2p XPS spectra of γ-Fe2O3, Fe3O4, α-Fe2O3, and the shape trend and position of thepeak are also exactly the same as that of pure Fe2O3. Therefore,the iron species exist in the state of Fe3+ over the sample, and thevalence state of the surface iron cannot be changed by dopingtrace manganese, which is highly consistent with the followingXRD results. However, the binding energy of Fe 2p in the0.010MnFeCu catalyst was a little higher than that of others, whichmay indicate that the higher interaction between Fe ions andneighbor high-valence Mn4+ species results in the formation ofFe species in a more positively charged environment.22,30

As shown in Figure 2d, one can find that two distinct peakscentered at 641.5 and 653.5 eV can be clearly observed, whichare attributed to Mn 2p3/2 and Mn 2p1/2, respectively. Byperforming peak-fitting deconvolution, the Mn 2p3/2 signal canbe decomposed into three components at the binding energiesof 640.0, 642.2, and 643.5 eV with three corresponding satellitepeaks at 648.0, 653.3, and 655.2 eV, respectively, which can be

assigned to Mn2+, Mn3+, and Mn4+ species on the surface of the0.010MnFeCu catalyst, similar to the previous report.19,21,31 As canbe seen from Figure 2d, both samples 0.010MnFeCu and0.020MnFeCu contain the newly formed Mn2+ species comparedto KMn8O16, and 0.010MnFeCu has more Mn4+ and Mn2+. Thismay be due to the part of Mn2+ being generated because of theinteraction among iron, copper, andMn species in the process ofbeing calcined. It is precisely because there are more Mn4+ in0.010MnFeCu (Mn4+/(Mn4+ + Mn3+ + Mn2+) = 0.45) that thesurrounding copper or iron is in a more positively chargedenvironment, which is consistent with the shifting of theirbinding energy to high binding energy.

3.4. XRD and HRTEM Results. In order to further identifythe detailed crystallinity and the crystalline phase of thesynthesized fresh samples, the XRD analysis was performedand the results of all catalysts are shown in Figure 3a. One canfind that all of the samples show the similar peaks. The peaksappearing at 2θ = 18.5, 30.2, 35.6, 37.2, 43.0, 57.1, 62.8, 74.5, and79.5o match well with the peaks of CuFe2O4 (JCPDS #25-0283),which are ascribed to the (220), (311), (400), (107), (422),(511), (440), and (533) lattice planes of the cubic spinelstructure, respectively. The small peaks can be clearly observedat 2θ = 32.5, 35.4, 35.5, 38.7, 38.9, 48.7, 58.3, 61.5, 66.2, and68.1°, corresponding to the reflection of (110), (002), (11−1),(111), (200), (20−2), (11−3), (31−1), and (220) of CuO(JCPDS #48-1548), which may be derived from thedecomposition of CuFe2O4 during the thermal treatmentprocess.32 Previous study7 shows that γ-Fe2O3 is an antispinelstructure with all iron existing in the Fe3+ form, and it exists intetrahedrally coordinated Fe3+ and octahedral Fe3+ structuresand contains cationic vacancies. Although γ-Fe2O3, CuFe2O4,and Fe3O4 have a face-centered cubic structure and belong to acubic system,28 the crystal structure is basically the same, but thedifference is that γ-Fe2O3 has cationic vacancies,

10 so its XRDresults are slightly different from those of the latter two.

Figure 3. (a) Powder XRD patterns of all samples: 0MnFeCu, 0.008MnFeCu, 0.010MnFeCu, 0.016MnFeCu, and 0.020MnFeCu; (b) partial enlargement of XRDresults; and (c) HRTEM images of the 0.010MnFeCu sample.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40513

Page 6: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

The difference between γ-Fe2O3 and CuFe2O4 can be clearlyobserved through partial enlarged drawing in Figure 3b. Theresults show that γ-Fe2O3 and CuFe2O4 coexist in our presentsamples. Meanwhile, it is interesting to find that the intensity ofthe characteristic peak for (110) of monoclinic CuO in the0.010MnFeCu sample is much higher than that of others.According to previous literature studies,33,34 it is reported thatMn species is easily combined with Fe species and then weakensthe interaction between Cu and Fe, leading to the precipitationof the CuO species, which are highly consistent with the results.Combining the analysis in Figure S4 (separate XRD result of the0.010MnFeCu sample), it can be suggested that some Mn2+

promotes the precipitation of Cu2+ fromCuFe2O4 and promotesCuO to grow along the CuO (110) (as shown in Figure 3c)crystal plane during the heat treatment.35,36 With the increase ofthe manganese loading, some manganese substances gatherthemselves to form tiny crystals or amorphous substances, whichcannot be observed from XRD results. Therefore, it can be seenthat the strength of CuO (110) gradually weakens. As reportedpreviously,37 the catalytic activity of CuO nanoparticles with theexposed active (110) facet for CO oxidation was significantlyhigher than that of the original nanoflower. It was also reportedthat the molecule NO preferentially binds to the oxygen atompointing upward to the (110) plane.37,38

On the basis of the above XRD results, it can be concludedthat three types of crystalline substances coexist in the presentcatalysts (which can be further proved by Figure 3c), and theexisting states of their cations are also highly consistent with theXPS results discussed above. The 0.010MnFeCu sample hasrelativelymore CuO (110) crystal plane (as shown in Figure 3c),and these CuO species are closely related to the surrounding γ-Fe2O3 and CuFe2O4, which may produce a number of defectsaccompanied by a large number of active oxygen species (O 1sXPS results of the 0.010MnFeCu sample shown in Figure 2a, andthese active oxygen species can be easily reduced to oxygenvacancies) at the two-phase contact surface, so it has higheractivity.3.5. SEM Mapping Results. In order to study the detailed

distribution and morphology of various elements on the surfaceof the catalyst, SEM mapping analysis was used. Figure 4 showselemental maps of Fe, O, Mn, and Cu on the surface of the0.010MnFeCu sample. It can be clearly seen that Mn, Fe, and O

elements are highly dispersed on the surface of the catalyst, butthe Cu element is relatively concentrated in a part of the area andin other parts is consistent with the distribution of otherelements. From the above observations, it is tentatively believedthat the Mn element is highly dispersed on the surface of thethree phases, possibly in the form of microcrystals orindeterminate forms, which is consistent with the XRD analysisresults. Some aggregated Cu element comes from CuO. In theregions where only iron and oxygen are distributed, these Fespecies are derived from γ-Fe2O3 nanocrystals. In the regionwhere Cu and Fe coexist, two elements exist in CuFe2O4nanoparticles as Cu−O−Fe. These CuO species are closelysurrounded by CuFe2O4 [can also be proved by the high-resolution TEM (HRTEM) images of the 0.010MnFeCu sampleshown in Figure 3c]. It can be seen from the figure that the Mnelement is highly uniformly dispersed, which indicates that someof the manganese species are doped on the surface of all phases.These manganese species may interact strongly with the surfacecations, resulting in a large number of defects and a large numberof active oxygen species (Cu2+−O−Mn4+, Cu2+−O−Mn3+, andMn4+−O−Fe3+ oxygen species are more easily reduced toCu+−□−Mn3+, Cu+−□−Mn2+, and Mn3+−□−Fe2+).17 Theabove SEM mapping result is in good consistent with the XRDand XPS results, further proving the coexistence of three phases(γ-Fe2O3, CuFe2O4, and CuO three phases are close and notindependent), which are difficultly distinguished only by theXRD results.

3.6. TPR Results and Corresponding XRD Results. Theredox performance is the key factor to influence the catalyticactivity of NO reduction by CO. Therefore, the reduction of allcatalysts was evaluated by the H2-TPR experiment, and theresults are shown in Figure 5. According to the literature,39 thefirst peak was attributed to the reduction of CuFe2O4 to Cu andFe3O4, and the second one is ascribed to the reduction of CuOto Cu0 (as shown in Figure 5a). However, studies40,41 reportedthat Cu−Fe spinel composite catalysts have similar composi-tions and TPR profiles, which attributed the three reductionpeaks from low temperature to high temperature to the highlydispersed CuO to Cu, CuFe2O4 to Cu and Fe3O4, and iron oxideto FeO or Fe, respectively.In order to understand the corresponding reduction peaks of

different substances in these catalysts, here we have performed asufficiently long time H2 reduction treatment of the 0.010MnFeCusample at different temperatures (shown in Figure 5b) so thatthe reduction reaction reaches a steady state. Their XRD analysisresults are shown in Figure 5b, which can be clearly seen that thecorresponding peaks of CuO disappeared completely at first(reduced at 160 °C). At the same time, a strong metal Cu peakappeared, probably because the larger particles of CuO arecompletely reduced or reduced to smaller particles. Here, theH2-TPR peak can be attributed to the reduction of CuO to Cu.Subsequently, as the reduction temperature continues toincrease (from 180 to 300 °C), the characteristic peak of γ-Fe2O3 gradually becomes smaller and eventually disappears,accompanied by the appearance of the Fe3O4 diffraction peak(Figure 5c) because γ-Fe2O3 is reduced to Fe3O4. As thetemperature continues to rise (from 300 to 380 °C), thediffraction peak of CuFe2O4 decreases until it disappears.Instead, the peak of Fe3O4 becomes sharper (Figure 5c),indicating that the two species (γ-Fe2O3 and CuFe2O4) arereduced to form larger Fe3O4 particles. When the temperaturereaches 800 °C, all the diffraction peaks disappear except for CuFigure 4. SEMmapping of various elements in the 0.010MnFeCu sample.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40514

Page 7: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

and Fe metals. This is because all substances are reduced to Cu0

and Fe0 at the high temperature.Similar to the previous literature,39,41 it can be proposed that

the highly dispersed CuO is reduced first and then, γ-Fe2O3 isreduced to Fe3O4; subsequently, CuFe2O4 is reduced to Fe3O4and Cu; and finally, Fe3O4 is reduced to FeO or Fe. Thereduction peaks associated with manganese cannot be observedhere, probably because Mn levels are too low to be observed.On the basis of the above facts, the overlapped peaks below

350 °C were tentatively fitted into three peaks by Gaussian−Lorentzian curves, as shown in Figure 5a. Three peaks located ataround 230, 270, and 300 °C correspond to the reduction ofhighly dispersed CuO to Cu, γ-Fe2O3 to Fe3O4, and CuFe2O4 toFe3O4 and Cu, respectively, which could be confirmed by theXRD of the catalysts gradually reduced by hydrogen. Theirrelative ratios can be obtained by comparing the peak area of H2-TPR profiles, as shown in Table 1. From the data in the table, itcan be seen that (1) the peak area corresponding to thereduction of CuO is approximately 3 times that of γ-Fe2O3. Thisis due to the fact that the amount of Fe in the precursor is twicethat of Cu, and there are three species (γ-Fe2O3, CuFe2O4, andCuO) in the catalyst. Theoretically, the ratio of the amount ofCuO and γ-Fe2O3 substances is 1:1, and the reduction of 1 molCuO to Cu needs 1 mol H2, whereas the reduction of 1 mol γ-Fe2O3 to Fe3O4 will consume 1/3 mol H2. Therefore, the area ofthe former is approximately 3 times of the latter, which furtherproves that our catalyst contains these three species. (2) The

value of CuO/CuFe2O4 is the highest in the 0.010MnFeCu sample,indicating that a proper proportion of the trace amount ofmanganese species loading promotes Cu2+ precipitation fromCuFe2O4 to form CuO and grow along the (110) plane, which isin good consistent with the XRD results in Figure 3a and theHRTEM images shown in Figure 3c. (3) All samples show thesimilar TPR profile (a small amount of manganese has a littleeffect on the relative content of each phase), except for0.010MnFeCu shifting to a lower temperature, which is in goodconsistent with its better catalytic performance. This may bebecause when an appropriate amount of manganese isincorporated, some of the newly added reactive oxygen species,such as Cu2+−O−Mn4+, Cu2+−O−Mn3+, and Mn4+−O−Fe3+,are formed on the surfaces of the phases so that the catalyst ismore easily reduced, making the reduction peak move to lowtemperature.It has been widely accepted that the synergistic interactions

among CuFe2O4, CuO, and γ-Fe2O3 can be appropriatelyreflected by their reduction behaviors. As reference, two puresamples of CuO and γ-Fe2O3 were also prepared (Figure 6).Here, it can be seen that pure CuO has a strong peak located ataround 350 °C corresponding to the reduction of CuO to Cu,which is similar to previous report.42 One can clearly find thatthe peaks ascribed to the reduction of CuO and γ-Fe2O3,occurring at 240 and 270 °C in our samples, shift to lowertemperature compared with those of the pure references, whichshows strong interaction and synergistic effect between the

Figure 5. (a) H2-TPR profiles of these catalysts and the overlapped peaks of all catalysts fitted by Gaussian−Lorentzian curves; (b) XRD results for the0.010MnFeCu sample fully reduced by H2 at different temperatures; and (c) partial enlargement of (b).

Table 1. H2 Consumption of Each Phase in Various Catalysts and the Proportion of Consumption

sample/MnFeCu 0 0.008 0.010 0.016 0.020H2 consumption (CuO) (CuO → Cu0) 3430 4255 4141 4434 3595H2 consumption (γ-Fe2O3) (γ-Fe2O3 → Fe3O4) 1167 1433 1323 1549 1200H2 consumption (CuFe2O4) (CuFe2O4 → Cu0 + Fe3O4) 9438 10 221 8574 9707 8465H2 (CuO)/H2 (CuFe2O4) 0.36 0.41 0.48 0.45 0.42H2 (CuO)/H2 (γ-Fe2O3) 2.94 2.97 3.13 2.86 2.99

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40515

Page 8: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

highly dispersed oxide particles.43 Previous researcher re-ported2,44 that the oxygen species in Cu2+−O−Mn4+, Cu2+−O−Mn3+, Mn4+−O−Fe3+, and Cu2+−O−Fe3+ can be moreeasily reduced than Cu2+−O−Cu2+ and Fe3+−O−Fe3+. There-fore, we propose that there is a strong interaction among thethree phases, and Cu2+−O−Fe3+ can be generated, as well as theincorporation of appropriate amount of Mn manganese canproduce another active oxygen (Cu2+−O−Mn4+, Cu2+−O−Mn3+,Mn4+−O−Fe3+, andMn3+−O−Fe3+) on the surface of thedifferent phases, which makes the reduction peaks shift to lowertemperature and further results in the excellent catalyticperformance. This further proves that there is a stronginteraction among CuO, CuFe2O4, and γ-Fe2O3 three-phasecoexistence catalyst systems.Through the above analysis (XRD, XPS, TPR, SEMmapping,

and HRTEM), it can be proposed that γ-Fe2O3, CuO, andCuFe2O4 three phases coexist in the catalyst, and stronginteraction and synergistic effect can generate among them(Figure 6), which is in line with the good performance of the0MnFeCu catalyst. Some loaded Mn species which are highlydispersed on the surface or between the contact interface of eachphase can further interact with the cations of the surface of thethree-phase coexisting catalyst so as to producemore defects andeventually produce a large number of active oxygen that can beeasily reduced to oxygen vacancies (confirmed by XPS and TPRresults). As a result, the 0.010MnFeCu catalyst has better catalyticactivity than the 0MnFeCu catalyst.

3.7. CO and/or NO Interaction with the Catalysts (inSitu FT-IR) and Possible Low-Temperature ReactionMechanism. To understand the adsorption and reductionperformance of these catalysts, in situ infrared spectroscopy(FT-IR) tests of CO adsorption were performed at a series oftemperatures, and the corresponding results of 0.010MnFeCu and0MnFeCu samples are given in Figure 7. One can clearly see thatthe two samples had identical peaks at 2115 and 2173 cm−1.Previous report18 assigned them to gas-phase CO vibration, butthey were also attributed to the peaks of adsorbed CO.4 In orderto clarify their assignment under the current conditions, wepurposely conducted a blank control experiment shown inFigure S5b (these two peaks have the identical intensity at thewhole temperature ranges), and the two peaks should be gas-phase CO. An interesting phenomenon can be observed in thepeak of around 2105 cm−1 (which can be clearly distinguishedfrom the peaks of the gas-phase CO shown in Figure S5b),appearing around 100 °C, gradually becoming stronger as thetemperature rises,13 and then gradually weakening until itdisappeared, which was assigned to Cu+−CO.45,46 Literature47reported that CO was preferentially adsorbed Cu+, and Cu+−CO was very stable, so Cu+ species can be probed by CO.However, this peak from the 0.010MnFeCu sample hasdisappeared completely at 200 °C, which is lower than that ofthe 0MnFeCu sample. This may be because when a small amount ofMn species are incorporated, Cu2+−O−Mn4+, Cu2+−O−Mn3+,Mn4+−O−Fe3+, and other active oxygen species are generated atthe contact interface of each phase2,18 so that these active oxygenspecies can be more easily reduced by CO, and eventually Cuspecies are more easily reduced to Cu+ (Cu+−□−Mn3+,Cu+−□−Mn2+, and Mn3+−□−Fe2+) and then to Cu0 (ingood consistent with the TPR results), which is in line withbetter catalytic performance. For both samples, the peaksobserved at about 2341 and 2360 cm−1 correspond to gas-phaseCO2,

48 whose intensity increases rapidly at 150 °C. At 150 °C,there is a large amount of Cu+−CO which can rapidly combinewith the adjacent oxygen to form CO2.

13,48

In order to gain an insight into the interaction betweenadsorbed NO and the surface components of the samples, theNO adsorption FT-IR spectra were recorded at varioustemperatures from 25 to 400 °C on 0MnFeCu and 0.010MnFeCucatalysts. As shown in Figure 8, for the 0.010MnFeCu sample, thepeaks observed at 1004 and 1628 cm−1 are attributed to thestretching mode of NO from bridging bidentate nitrate,

Figure 6. H2-TPR profiles of pure CuO, γ-Fe2O3, and 0.010MnFeCusamples.

Figure 7. In situ FT-IR results of CO (10% in volume) adsorption on 0.010MnFeCu and 0MnFeCu catalysts.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40516

Page 9: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

gradually decreasing with the increase of temperature and beinglost at high temperature because of their poor stability.2,49 Bandsappearing at 1205 and 1585 cm−1 correspond to symmetric andasymmetric vibration modes, respectively, of chelating bidentatenitrate,2,49 which first increase and then decrease slowly withraising the temperature up to 400 °C, which is in line with itsrelatively good stability compared with others. The peak at 1242cm−1 is assigned to the vibration of chelating bidentate nitrite,having been reported by other researchers,13,50 which has betterthermal stability than that of others. It is worth noticing that thepeak located at 1277 cm−1 is larger than others at roomtemperature and decreases rapidly with increasing temperatureand disappears completely below 200 °C because of itsdesorption or decomposition or transformation. The peak isassigned to the vibration of the linear nitrite and has a relativelypoor stability.48,50 Bands appearing at 1551 and 1603 cm−1

correspond to monodentate nitrate and bridging monodentatenitrate, respectively.2 As the temperature rises, the first onedecreases gradually and absolutely vanishes below 350 °C andthe second one shows a tendency to decrease in intensity, but itnever disappears even at 400 °C.2 The bands at 1844 and 1903cm−1 are noticed for both samples, and their intensity decreasesgradually with an increase in temperature but never vanishes.The literature4,51 reported that the two peaks are assigned to theweakly adsorbed NO or gas-phase NO, in order to get trueinformation at our present condition. A blank control FT-IR testwas tentatively conducted (without samples, the otherconditions are exactly the same, shown in Figure S5a), fromwhich we can accurately attribute them to the gas-phase NOthrough the experimental results.It is interesting to find that the 0MnFeCu sample has the similar

vibration bands at 1903, 1844, 1630, 1608, 1587, 1545, 1279,1249, 1211, and 1006 cm−1 corresponding to gas-phase NO(1903 and 1844 cm−1), bridging monodentate nitrate (1608

cm−1), linear nitrite (1279 cm−1), chelating bidentate nitrite(1249 cm−1), chelating bidentate nitrate (1211 and 1587 cm−1),monodentate nitrate (1545 cm−1), and bridging bidentatenitrate (1630 and 1006 cm−1), respectively. In addition,compared with 0MnFeCu, it can be obviously noticed that all ofthe vibration bands of adsorbedNO species on the surface of the0.010MnFeCu catalyst except for two peaks at 1844 and 1903 cm

−1

(both the catalysts were surrounded by the same gasatmosphere) shift toward the lower wavenumbers. Thisphenomenon may be interpreted by the more stronglysynergistic interaction among CuO, CuFe2O4, and γ-Fe2O3interfaces, which are promoted by some Mn species highlydispersed between these contact surfaces. Therefore, it can causethe d-electron’ back-donation from the metal cation to theantibonding orbital of NO, and then the N−O bond wasweakened, whichmay be accounted for its better performance.48

Comparing the change regulation of intensity of NOx adsorbedon the two samples, some interesting results can be obtained,and the weakening degree of the peak intensity of nitrate andnitrite over 0.010MnFeCu is greater than that of 0MnFeCu, indicatingthat the decomposition, adsorption, and desorption abilities of0.010MnFeCu are better than those of the 0MnFeCu catalyst.

52

From the FT-IR spectra results discussed above, it can beconcluded that when the catalyst is placed in an atmosphere ofNO, the oxygen species on its surface combine with NO andoxidize it to NOx species, which shows that the present catalysthas strong oxygen storage and release characteristics.50 As thetemperature rises, some unstable NO begins to desorb,decompose, or transform. When the temperature reaches 400°C, most of the NOx species has been desorbed or decomposed,except for stable nitrates.50 When a proper amount of Mnsubstance is added to the contact interface among CuO, γ-Fe2O3, and the surrounding CuFe2O4, some defects may occurbecause of strong interactions so that more cations are in an

Figure 8. In situ DRIFTS results of NO (5% in volume) adsorption on 0MnFeCu and 0.010MnFeCu catalysts.

Figure 9. In situ FT-IR results of NO and CO (5% NO and 10% CO in volume, respectively) coadsorption on 0.010MnFeCu and 0MnFeCu catalysts.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40517

Page 10: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

unsaturated state (can adsorb a great amount of oxygen species,supported by XPS results) and thus have a stronger adsorbedability on NO and eventually make all infrared absorption peaksof NOx move toward lower wavenumbers.52

In order to better understand themechanism ofNO reductionby CO, the properties and relative quantities of surface speciesover 0.010MnFeCu and 0MnFeCu samples were studied by CO andNO coadsorption in situ FT-IR techniques under simulatedreaction conditions, as shown in Figure 9. One can clearly findthat the interaction of CO and NO with samples producesseveral kinds of nitrates and nitrite species, which is similar to theresults of NO adsorbed alone on the surface at roomtemperature. NO was preferentially adsorbed on the surface ofthe samples because of unpaired electrons, and then theadsorption of CO species on the surface of sample was inhibited,which is in line with the previous studies.50 However, somedifferences from those of the NO alone adsorption can be found,and the intensities of all the NOx species decrease more quicklyand they vanish completely at lower temperature, whichdemonstrate that some nitro-intermediate species have reactedwith CO. The peaks observed at 2341 and 2360 cm−1

correspond to gas-phase CO2,48 which shows the oxidation of

CO even at relatively low temperature.Here, a particularly different phenomenon is noticed, that is,

when the temperature rises to 175 °C, all NOx species disappearcompletely, and two special peaks located at about 1770−1790and 2105 cm−1 begin to appear (the intensity of the peak at 2105cm−1 over 0.010MnFeCu is much higher than that of 0MnFeCu. Thismay be because when an appropriate amount of manganese isincorporated, some active oxygen species such as Cu2+−O−Mn4+, Cu2+−O−Mn3+, and Mn4+−O−Fe3+ (Figure 2d showsthat the sample 0.010MnFeCu has more Mn4+ species andtherefore can produce more Cu2+−O−Mn4+ and Mn4+−O−Fe3+) can be generated and then can be rapidly reduced, andsome Cu+−□−Mn3+, Cu+−□−Mn2+, and Mn3+−□−Fe2+ areimmediately generated and then immediately a large amount ofCu+−CO is produced).As the temperature rises, their intensity increases sharply.

When the temperature rises to 200 °C, their strength reaches themaximum value with the conversion of CO and NO beginningto increase rapidly (which is well in line with the conversion ofNO and CO shown in Figure 1a,d), and as the temperaturecontinues to rise, their strength gradually decreases until itdisappears. On the basis of the current FT-IR analysis results andprevious studies,53 the first peak can be attributed to thevibration of M−NO (M represents some cations), while it isassigned to Cu+−NO (the detailed FT-IR analysis in Figure S6).The second one has been assigned to CO bonded to Cu+, whichis generally accepted by many previous literature studies.45,50

Another slightly different peak from the above two, located at2237 cm−1, was attributed to the gas phase of N2O,

48 whichbegan to appear at 200 °C and gradually decreased until itdisappeared with the increase in temperature. This is highlyconsistent with the selectivity and catalytic activity of the catalystshown in Figure 1b,c. We may intentionally propose that atroom temperature, NO preferentially adsorbs on some metalcations from the surface of the catalyst to cover some active sites,thereby inhibiting the adsorption of CO;50 With the gradualincrease of temperature, part of NOx reacts directly with CO,and the other part desorbs or decomposes. Eventually, NOxdisappears completely at 175 °C, and at this point, CO candirectly react with the surface active oxygen (Cu2+−O−Mn4+,Cu2+−O−Mn3+, and Cu2+−O−Fe3+) to make Cu2+ on the

catalyst surface reduced to Cu0 or Cu+ and generate a largenumber of oxygen vacancies (Cu+−□−Mn3+, Cu+−□−Mn2+,and Cu+−□−Fe2+).18,45 When the temperature rises to 200 °C,a large amount of Cu+ (some Cu2+−O−Cu2+ are also reduced toCu+, resulting in further increase in the content of Cu+) isgenerated on the surface of the catalyst so that a certain amountof NO and a large amount of CO are adsorbed. Meanwhile,surface oxygen vacancies can weaken the N−O bond andpromote the decomposition of the adsorbed NO species into Nand O atoms,13 and then O combines with the adjacent COadsorbed on Cu+ to form CO2. Most of the N radicals combinewith neighboring Cu+−NOmolecules to formN2O, and the restcombined with each other to form N2, which is in goodconsistent with the low N2 selectivity at 200 °C (Figure 1b).Thereafter, as the temperature continues to rise to 300 °C, mostof the Cu+ is reduced to Cu0 accompanied by the disappearanceof Cu+−CO and Cu+−NO. Because of the disappearance ofCu+−NO, most NO may be directly decomposed by oxygenvacancies to O and N, and two N atoms are combined into N2,which can be proved by the selectivity of N2 increasing sharply asshown in Figure 1b. Finally, all the peaks disappeared at 300 °C,except for CO and NO gas-phase peaks. Therefore, the FT-IRspectra results above 300 °C will not be shown.One can find several special phenomena. (1) No COx

absorption peaks can be seen throughout the temperaturerange; however, the peak of CO2 can be seen, which indicatesthat CO can quickly react with the catalyst and desorb so that itdoes not occupy the reaction site (as shown in Figure 7). (2)When the temperature rises to 175 °C, all NOx species disappearcompletely, and two special peaks located at about 1770−1790(Cu+−NO) and 2105 cm−1 (Cu+−CO) begin to appear (asshown in Figure 9). (3) When the temperature rises by about300 °C, all the peaks disappear, except for the characteristic peakof the gas-phase CO2 (as shown in Figure 9). These phenomenacannot be found in the previous literature. The mechanism inthis paper is slightly different from the above litera-ture.2−4,6,12−14,18,44,45,49,50

3.8. Possible High-Temperature Reaction Mechanismover the Catalyst. When the temperature rises to 300 °C,because of high-temperature desorption or rapid reaction, allpeaks vanish completely except for some peaks of gas-phasespecies (CO, NO, and CO2). Therefore, we cannot get anyinformation from FT-IR spectra results, but Cu2+ species havebeen reduced to Cu0 before 300 °C, which can be proved by thecomplete disappearance of peaks for Cu+−CO.To further prove that the Cu species exist in the form of Cu0

on the surface of the catalyst under our present condition athigher temperature (≥300 °C), we collected the catalyst whichhad reacted for a long time at 400 °C and been swept by nitrogento room temperature. The XRD results (Figure S7a,b) showedthat all the peaks assigned to CuO disappeared completelyexcept the peaks for γ-Fe2O3, accompanied by the appearance ofsome strong Cu peaks and Fe3O4 peaks, which furtherdemonstrates that some Cu2+ species may have been reducedto Cu0. The detailed discussion in the figure also proves this.Previous studies13,46 reported the relationship between Cu+

content and temperature in Cu-containing composite oxides,and it was found that when the temperature was lower than 300°C, Cu+ was completely reduced to Cu0, which is in line with theresults.The XRD results of samples having reacted long enough

under 400, 600, and 800 °C (shown in Figure S8) show that it issimilar to the crystal-phase species at 400 °C. This indicates that

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40518

Page 11: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

there may be no significant change in the active components ofthe catalyst in the temperature range of 400 to 1000 °C.Therefore, here, it was tentatively believed that under high-temperature conditions, γ-Fe2O3 and Fe3O4 may be catalyticallyactive substances. It was reported that both Fe3O4 and γ-Fe2O3had similar cubic dense stacked oxide ion arrays,11 whichexplains the interchangeability of the two compounds in redoxand reduction because these reactions have relatively littlechange on the whole structure. The conversion between Fe3O4and γ-Fe2O3 is quite easy, which is in good consistent with theexcellent high-temperature performance of our catalyst,29 and itis also explained that Fe3O4 and γ-Fe2O3 phases are found tocoexist in the process of all high-temperature reactions.9

Studies54,55 have shown that the catalytic activity of γ-Fe2O3 issignificantly higher than that of α-Fe2O3, which is similar withthe present study (Figure 10). However, the thermal stability of

α-Fe2O3 is better than that of γ-Fe2O3.56,57 Therefore, it is also

possible that part of γ-Fe2O3 is converted to α-Fe2O3 at hightemperatures. The XRD results shown in Figure 11 indicate that

α-Fe2O3 is generated in the 0MnFeCu sample when thetemperature rises to 400 °C. In contrast, samples containingtraces of Mn species do not have this corresponding diffractionpeak. This may be because when the temperature is increased,theMn species can suppress the γ-Fe2O3 to α-Fe2O3 conversion,which is in good consistent with the previous report,58 in whichit is reported that the Mn3+-doped γ-Fe2O3 displays super-paramagnetic behavior at room temperature, and the conversionof γ-Fe2O3 to α-Fe2O3 at high temperatures can also beprevented. Combined with experimental results and literature, itcan be inferred that Mn3+ inhibits the conversion of γ-Fe2O3 toα-Fe2O3.The possible high-temperature reaction mechanism was

proposed. When surface oxygen ions are removed by CO,accompanied by the formation of Fe2+, the polycrystalline

transition of γ-Fe2O3 to Fe3O4 begins, and then the newlyformed Fe2+ diffuses into the bulk phase and forms Fe3O4.Because the crystal structures of γ-Fe2O3 and Fe3O4 are verysimilar, the phase transition between them is actually atopological process and needs no crystallographic changes,which makes the migration of Fe2+ need less energy. Therefore,γ-Fe2O3 can be easily reduced to Fe3O4. Similarly, the transitionof Fe3O4 to γ-Fe2O3 starts when NO contacts with oxygenvacancy. The electrons are transferred from bulk Fe2+ toadjacent Fe3+ and finally to the surface to bind with NO afterseveral transfers so that NO is decomposed intoN andO (N andadjacent N combine to form N2). At the same time, the newlygenerated Fe3+ expands from the body phase to the outside,leaving a cationic vacancy to form γ-Fe2O3.

54,59,60

To further prove that different crystal phases in the strongsynergistic coexistence catalyst system play dominant catalyticroles at different temperatures,61 we show the diagram of therelationship between the catalytic activity and temperature fordifferent pure phases. It can be clearly seen from Figure 12 that

when the temperature rises from 180 to 200 °C, the catalyticactivity of pure CuO increases abruptly because of the excellentlow-temperature catalytic activity, and as the temperaturecontinues to rise, its activity decreases sharply because of mostCu2+ being reduced to Cu0. However, under the sameconditions, when the temperature is less than 250 °C, thecatalytic conversion of pure Fe2O3 is almost 0, and the catalyticactivity gradually increases to 100% as the temperaturecontinues to increase up to 600 °C because of the excellenthigh-temperature catalytic activity. It is clear that the curves ofthe physical mixture of γ-Fe2O3, CuO, and CuFe2O4 are justsimple summations of pure species. The difference from theabove three is that the present sample not only has a low-temperature catalytic activity such as pure CuO but also has ahigh-temperature stability such as γ-Fe2O3 and also completelybetter than physical mixing. Combined with the XRD, XPS, FT-IR, SEM mapping, and in situ FT-IR analyses, it is believed thatwhen the temperature is lower than 250 °C, the CuO withrelatively more Mn species on the surface in present three-phasecoexisting catalyst play a dominant catalytic role, andmost of theCu2+ is reduced to Cu0 as the temperature increases to 300 °C.Meanwhile, Fe3O4 (CuFe2O4 was reduced to Fe3O4) begins toappear.When the temperature is higher than 400 °C (400−1000°C), Fe3O4/γ-Fe2O3 becomes a highly active substance. Apossible NO reduction by the CO reaction mechanism(schematic diagram) is proposed in Figure 13.

Figure 10. Conversion of NO to N2 on γ-Fe2O3, α-Fe2O3, and Fe3O4.

Figure 11. XRD results of 0.010MnFeCu and 0MnFeCu samples.

Figure 12. NO conversion on CuO, 0.010MnFeCu, γ-Fe2O3, and thephysical mixture.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40519

Page 12: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

4. CONCLUSIONSThe three phases of γ-Fe2O3, CuFe2O4, and CuO coexist in ourcatalyst system, resulting in a large number of active oxygenspecies in their interface, which can be reduced to oxygenvacancies by CO, resulting in a good activity of the 0MnFeCucatalyst even without manganese. When a suitable amount ofmanganese is incorporated into the catalyst, the Mn2+ speciesinteracts strongly with the Fe species, allowing Cu2+ toprecipitate from CuFe2O4 and preferentially grow along theCuO(110) plane surface, which has better catalytic perform-ance. At high temperature (300−1000 °C), Mn3+ can suppressthe transformation of γ-Fe2O3 to α-Fe2O3, which leads to a stablehigh-temperature catalytic performance. Mn4+ on the surface ofthe catalyst generates reactive oxygen species, which can bemore easily reduced to form some oxygen vacancies during thereaction. These factors eventually lead to the excellent catalyticperformance of the 0.010MnFeCu catalyst with a broad temper-ature window.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acsami.8b13220.

Water and sulfur resistance of the 0.010 sample; XRDpatterns; N2 adsorption−desorption isotherms of all thesamples and pore size distribution of the samples; TEMimages; and in situ FT-IR results (PDF)Influence of SO2, SO2 + H2O, and H2O on NOconversion of the sample 0.010MnFeCu; N2 adsorption−desorption isotherms of all samples and pore sizedistribution of the samples; surface area and porestructure parameters of samples; TEM image and XRDresult of the catalyst 0.010MnFeCu and at differenttemperatures; in situ FT-IR results for NO and CO inblank control experiments without the sample; partiallymagnified in situ DRIFTS results of NO and COcoadsorption on the catalyst 0.010MnCuFe and 0MnCuFe;and XRD results of the catalyst 0MnCuFe and 0.010MnCuFetreated by CO+NOunder 400 °C and partial enlargementof XRD results (PDF)

■ AUTHOR INFORMATIONCorresponding Authors*E-mail: [email protected] (B.L.).*E-mail: [email protected] (L.D.).ORCIDBin Li: 0000-0002-4080-3268Lihui Dong: 0000-0002-1184-7317Lin Dong: 0000-0002-8393-6669NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThis work was supported by the National Natural ScienceFoundation of China (nos. 21507014, 21663006, and21763003) and the Program for Science and TechnologyDevelopment Plan of Nanning (no. 20163146).

■ REFERENCES(1) Liang, Z.; Ma, X.; Lin, H.; Tang, Y. The energy consumption andenvironmental impacts of SCR technology in China. Appl. Energy 2011,88, 1120−1129.(2) Xiong, Y.; Yao, X.; Tang, C.; Zhang, L.; Cao, Y.; Deng, Y.; Gao, F.;Dong, L. Effect of CO-pretreatment on the CuO−V2O5/γ-Al2O3catalyst for NO reduction by CO. Catal. Sci. Technol. 2014, 4, 4416−4425.(3) Dong, L.; Zhang, B.; Tang, C.; Li, B.; Zhou, L.; Gong, F.; Sun, B.;Gao, F.; Dong, L.; Chen, Y. Influence of CeO2 modification on theproperties of Fe2O3−Ti0.5Sn0.5O2 catalyst for NO reduction by CO.Catal. Sci. Technol. 2013, 4, 482.(4) Wang, L.; Cheng, X.; Wang, Z.; Zhang, X.; Ma, C. NO reductionby CO over iron-based catalysts supported by activated semi-coke.Can.J. Chem. Eng. 2017, 95, 449−458.(5) Shen, B.; Liu, T.; Zhao, N.; Yang, X.; Deng, L. Iron-dopedMn-Ce/TiO2 catalyst for low temperature selective catalytic reduction of NOwith NH3. J. Environ. Sci. 2010, 22, 1447−1454.(6) Li, J.; Wang, S.; Zhou, L.; Luo, G.; Wei, F. NO reduction by COover a Fe-based catalyst in FCC regenerator conditions. Chem. Eng. J.2014, 255, 126−133.(7) Qu, W.; Chen, Y.; Huang, Z.; Gao, J.; Zhou, M.; Chen, J.; Li, C.;Ma, Z.; Chen, J.; Tang, X. Active Tetrahedral Iron Sites of γ-Fe2O3Catalyzing NO Reduction by NH3. Environ. Sci. Technol. Lett. 2017, 4,246−250.(8) Fan, Z.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Niu, C. RationallyDesigned Porous MnOx−FeOx Nanoneedles for Low-TemperatureSelective Catalytic Reduction of NOx by NH3. ACS Appl. Mater.Interfaces 2017, 9, 16117−16127.

Figure 13. Possible reactionmechanism of NO reduction by CO on the0.010MnCuFe catalyst (schematic diagram).

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40520

Page 13: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

(9) Capone, S.; Manera, M. G.; Taurino, A.; Siciliano, P.; Rella, R.;Luby, S.; Benkovicova, M.; Siffalovic, P.; Majkova, E. Fe3O4/γ-Fe2O3Nanoparticle Multilayers Deposited by the Langmuir−BlodgettTechnique for Gas Sensors Application. Langmuir 2014, 30, 1190−1197.(10) Nica, S.-L.; Nica, V.; Grigoras, V. C.; Varganici, C.-D.; Popovici,D.; Hulubei, C.; Ioan, S. Influence of two structural phases of Fe3O4 andγ-Fe2O3 on the properties of polyimide/iron oxide composites. Polym.Int. 2015, 64, 1172−1181.(11) Daniel, P.; Shylin, S. I.; Lu, H.; Tahir, M. N.; Panthofer, M.;Weidner, T.; Moller, A.; Ksenofontov, V.; Tremel, W. The surfacechemistry of iron oxide nanocrystals: surface reduction of γ-Fe2O3 toFe3O4 by redox-active catechol surface ligands. J. Mater. Chem. C 2018,6, 326−333.(12) Dasireddy, V. D. B. C.; Likozar, B. Selective catalytic reduction ofNOx by CO over bimetallic transition metals supported by multi-walledcarbon nanotubes (MWCNT). Chem. Eng. J. 2017, 326, 886−900.(13) Zhang, L.; Yao, X.; Lu, Y.; Sun, C.; Tang, C.; Gao, F.; Dong, L.Effect of precursors on the structure and activity of CuO-CoOx/γ-Al2O3catalysts for NO reduction by CO. J. Colloid Interface Sci. 2017, 509,334−345.(14) Liu, T.; Yao, Y.; Wei, L.; Shi, Z.; Han, L.; Yuan, H.; Li, B.; Dong,L.;Wang, F.; Sun, C. Preparation and Evaluation of Copper-ManganeseOxide as a High-Efficiency Catalyst for CO Oxidation and NOReduction by CO. J. Phys. Chem. C 2017, 121, 12757−12770.(15) Zhuang, Y.-T.; Gao, W.; Yu, Y.-L.; Wang, J.-H. Facile fabricationof three-dimensional porous CuFe2O4 cages as highly efficient andrecyclable heterogeneous catalyst. Mater. Des. 2017, 130, 294−301.(16) Dharmarathna, S.; King’ondu, C. K.; Pahalagedara, L.; Kuo, C.-H.; Zhang, Y.; Suib, S. L. Manganese octahedral molecular sieve (OMS-2) catalysts for selective aerobic oxidation of thiols to disulfides. Appl.Catal., B 2014, 147, 124−131.(17) Zha, K.; Cai, S.; Hu, H.; Li, H.; Yan, T.; Shi, L.; Zhang, D. In SituDRIFTs Investigation of Promotional Effects of Tungsten on MnOx-CeO2/meso-TiO2 Catalysts for NOx Reduction. J. Phys. Chem. C 2017,121, 25243−25254.(18) Wang, L.; Cheng, X.; Wang, Z.; Ma, C.; Qin, Y. Investigation onFe-Co binary metal oxides supported on activated semi-coke for NOreduction by CO. Appl. Catal., B 2017, 201, 636−651.(19) Shen, Q.; Zhang, L.; Sun, N.; Wang, H.; Zhong, L.; He, C.; Wei,W.; Sun, Y. Hollow MnOx-CeO2 mixed oxides as highly efficientcatalysts in NO oxidation. Chem. Eng. J. 2017, 322, 46−55.(20) Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. A superior Ce-W-Timixed oxide catalyst for the selective catalytic reduction of NOx withNH3. Appl. Catal., B 2012, 115−116, 100−106.(21) Yan, L.; Liu, Y.; Zha, K.; Li, H.; Shi, L.; Zhang, D. Scale-ActivityRelationship ofMnOx-FeOy Nanocage Catalysts Derived from PrussianBlue Analogues for Low-Temperature NO Reduction: Experimentaland DFT Studies. ACS Appl. Mater. Interfaces 2017, 9, 2581−2593.(22) Zhang, Y.; Qin, Z.; Wang, G.; Zhu, H.; Dong, M.; Li, S.; Wu, Z.;Li, Z.; Wu, Z.; Zhang, J.; Hu, T.; Fan, W.; Wang, J. Catalyticperformance of MnOx-NiO composite oxide in lean methanecombustion at low temperature. Appl. Catal., B 2013, 129, 172−181.(23) Li, J.; Liang, X.; Xu, S.; Hao, J. Catalytic performance ofmanganese cobalt oxides on methane combustion at low temperature.Appl. Catal., B 2009, 90, 307−312.(24) Zhang, W.; Quan, B.; Lee, C.; Park, S.-K.; Li, X.; Choi, E.; Diao,G.; Piao, Y. One-Step Facile Solvothermal Synthesis of Copper Ferrite-Graphene Composite as a High-Performance Supercapacitor Material.ACS Appl. Mater. Interfaces 2015, 7, 2404−2414.(25) Zhang, Q.; Wang, H.; Ning, P.; Song, Z.; Liu, X.; Duan, Y. In situDRIFTS studies on CuO-Fe2O3 catalysts for low temperature selectivecatalytic oxidation of ammonia to nitrogen. Appl. Surf. Sci. 2017, 419,733−743.(26) Hussain, S.; Hussain, S.; Waleed, A.; Tavakoli, M. M.; Wang, Z.;Yang, S.; Fan, Z.; Nadeem, M. A. Fabrication of CuFe2O4/α-Fe2O3Composite Thin Films on FTO Coated Glass and 3-D NanospikeStructures for Efficient Photoelectrochemical Water Splitting. ACSAppl. Mater. Interfaces 2016, 8, 35315−35322.

(27) Sun, G.; Dong, B.; Cao, M.; Wei, B.; Hu, C. HierarchicalDendrite-Like Magnetic Materials of Fe3O4, γ-Fe2O3, and Fe with HighPerformance of Microwave Absorption. Chem. Mater. 2011, 23, 1587−1593.(28) Lu, J.; Jiao, X.; Chen, D.; Li, W. Solvothermal Synthesis andCharacterization of Fe3O4 and γ-Fe2O3 Nanoplates. J. Phys. Chem. C2009, 113, 4012−4017.(29) Zhang, X.; Niu, Y.; Meng, X.; Li, Y.; Zhao, J. Structural evolutionand characteristics of the phase transformations between α-Fe2O3,Fe3O4 and γ-Fe2O3 nanoparticles under reducing and oxidizingatmospheres. CrystEngComm 2013, 15, 8166−8172.(30) Liu, X.-S.; Jin, Z.-N.; Lu, J.-Q.; Wang, X.-X.; Luo, M.-F. Highlyactive CuO/OMS-2 catalysts for low-temperature CO oxidation.Chem.Eng. J. 2010, 162, 151−157.(31) Li, Y.; Wan, Y.; Li, Y.; Zhan, S.; Guan, Q.; Tian, Y. Low-Temperature Selective Catalytic Reduction of NO with NH3 overMn2O3-Doped Fe2O3 Hexagonal Microsheets. ACS Appl. Mater.Interfaces 2016, 8, 5224−5233.(32) Dhanda, R.; Kidwai, M. Magnetically separable CuFe2O4/reduced graphene oxide nanocomposites: as a highly active catalyst forsolvent free oxidative coupling of amines to imines. RSC Adv. 2016, 6,53430−53437.(33) Ding, M.; Qiu, M.; Wang, T.; Ma, L.; Wu, C.; Liu, J. Effect of ironpromoter on structure and performance of CuMnZnO catalyst forhigher alcohols synthesis. Appl. Energy 2012, 97, 543−547.(34) Ding, M.; Qiu, M.; Liu, J.; Li, Y.; Wang, T.; Ma, L.; Wu, C.Influence of manganese promoter on co-precipitated Fe-Cu basedcatalysts for higher alcohols synthesis. Fuel 2013, 109, 21−27.(35) Lu, X.; Bian, W.; Min, C.; Fu, Z.; Zhang, Q.; Zhu, H. Cationdistribution of high-performance Mn-substituted ZnGa2O4 microwavedielectric ceramics. Ceram. Int. 2018, 44, 10028−10034.(36) Liang, X.; Zhong, Y.; Zhu, S.; He, H.; Yuan, P.; Zhu, J.; Jiang, Z.The valence and site occupancy of substituting metals in magnetitespinel structure Fe3−xMxO4 (M = Cr, Mn, Co and Ni) and theirinfluence on thermal stability: An XANES and TG-DSC investigation.Solid State Sci. 2013, 15, 115−122.(37) Peng, H.; Liu, Y.; Guo, Y.; Zhang, J.; Zhang, L.; Zhou, S.; Xu, X.;Liu, W.; Zhang, N.; Wang, X. Treating Copper(II) Oxide Nanoflowerswith Hydrogen Peroxide: A Novel and Facile Strategy To PrepareHigh-Performance Copper(II) Oxide Nanosheets with Exposed (1 1 0)Facets. ChemCatChem 2016, 8, 3714−3719.(38) Zhao, Y.; Shi, H.; Chen, M.; Teng, F. Splitting growth of novelCuO straw sheaves and their improved photocatalytic activity due toexposed active {110} facets and crystallinity. CrystEngComm 2014, 16,2417−2423.(39) Yang, S.-C.; Su, W.-N.; Lin, S. D.; Rick, J.; Cheng, J.-H.; Liu, J.-Y.;Pan, C.-J.; Liu, D.-G.; Lee, J.-F.; Chan, T.-S.; Sheu, H.-S.; Hwang, B.-J.Preparation of nano-sized Cu from a rod-like CuFe2O4: Suitable forhigh performance catalytic applications.Appl. Catal., B 2011, 106, 650−656.(40) Faungnawakij, K.; Kikuchi, R.; Fukunaga, T.; Eguchi, K. Catalytichydrogen production from dimethyl ether over CuFe2O4 spinel-basedcomposites: Hydrogen reduction and metal dopant effects. Catal.Today 2008, 138, 157−161.(41) Lin, X.; Zhang, Y.; Yin, L.; Chen, C.; Zhan, Y.; Li, D.Characterization and catalytic performance of copper-based WGScatalysts derived from copper ferrite. Int. J. Hydrogen Energy 2014, 39,6424−6432.(42) Perin, G.; Fabro, J.; Guiotto, M.; Xin, Q.; Natile, M. M.; Cool, P.;Canu, P.; Glisenti, A. Cu@LaNiO3 based nanocomposites in TWCapplications. Appl. Catal., B 2017, 209, 214−227.(43) Chi, G.; Shen, B.; Yu, R.; He, C.; Zhang, X. Simultaneous removalof NO and Hg0 over Ce-Cu modified V2O5/TiO2 based commercialSCR catalysts. J. Hazard. Mater. 2017, 330, 83−92.(44) Sun, C.; Tang, Y.; Gao, F.; Sun, J.; Ma, K.; Tang, C.; Dong, L.Effects of different manganese precursors as promoters on catalyticperformance of CuO-MnOx/TiO2 catalysts for NO removal by CO.Phys. Chem. Chem. Phys. 2015, 17, 15996−16006.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40521

Page 14: Mn-Modified CuO, CuFe2O4, and γ-Fe2O3 Three-Phase Strong ...

(45) Xiong, Y.; Yao, X.; Tang, C.; Zhang, L.; Cao, Y.; Deng, Y.; Gao,F.; Dong, L. Effect of CO-pretreatment on the CuO-V2O5/γ-Al2O3catalyst for NO reduction by CO. Catal. Sci. Technol. 2014, 4, 4416−4425.(46) Qi, L.; Yu, Q.; Dai, Y.; Tang, C.; Liu, L.; Zhang, H.; Gao, F.;Dong, L.; Chen, Y. Influence of cerium precursors on the structure andreducibility of mesoporous CuO-CeO2 catalysts for CO oxidation.Appl. Catal., B 2012, 119−120, 308−320.(47) Wan, H.; Wang, Z.; Zhu, J.; Li, X.; Liu, B.; Gao, F.; Dong, L.;Chen, Y. Influence of CO pretreatment on the activities of CuO/γ-Al2O3 catalysts in CO+O2 reaction. Appl. Catal., B 2008, 79, 254−261.(48) Yao, X.; Gao, F.; Yu, Q.; Qi, L.; Tang, C.; Dong, L.; Chen, Y. NOreduction by CO over CuO-CeO2 catalysts: effect of preparationmethods. Catal. Sci. Technol. 2013, 3, 1355−1366.(49) Liu, T.; Qian, J.; Yao, Y.; Shi, Z.; Han, L.; Liang, C.; Li, B.; Dong,L.; Fan, M.; Zhang, L. Research on SCR of NO with CO over theCu0.1La0.1Ce0.8O mixed-oxide catalysts: Effect of the grinding. J. Mol.Catal. A: Chem. 2016, 430, 43.(50) Liu, L.; Liu, B.; Dong, L.; Zhu, J.; Wan, H.; Sun, K.; Zhao, B.;Zhu, H.; Dong, L.; Chen, Y. In situ FT-infrared investigation of CO or/and NO interaction with CuO/Ce0.67Zr0.33O2 catalysts. Appl. Catal., B2009, 90, 578−586.(51) Qi, G.; Yang, R. T.; Chang, R. MnOx-CeO2 mixed oxidesprepared by co-precipitation for selective catalytic reduction of NOwith NH3 at low temperatures. Appl. Catal., B 2004, 51, 93−106.(52) Deng, C.; Huang, Q.; Zhu, X.; Hu, Q.; Su, W.; Qian, J.; Dong, L.;Li, B.; Fan, M.; Liang, C. The influence of Mn-doped CeO2 on theactivity of CuO/CeO2 in CO oxidation and NO + CO model reaction.Appl. Surf. Sci. 2016, 389, 1033−1049.(53) Kantcheva, M. FT-IR spectroscopic investigation of the reactivityof NOx species adsorbed on Cu2+/ZrO2 and CuSO4/ZrO2 catalyststoward decane. Appl. Catal., B 2003, 42, 89−109.(54)Wei, X.; Zhou, Y.; Li, Y.; Shen, W. Polymorphous transformationof rod-shaped iron oxides and their catalytic properties in selectivereduction of NO by NH3. RSC Adv. 2015, 5, 66141−66146.(55) Zhao, N.; Ma, W.; Cui, Z.; Song, W.; Xu, C.; Gao, M. PolyhedralMaghemite Nanocrystals Prepared by a Flame Synthetic Method:Preparations, Characterizations, and Catalytic Properties. ACS Nano2009, 3, 1775−1780.(56) Yen, F. S.; Chen, W. C.; Yang, J. M.; Hong, C. T. Crystallite SizeVariations of Nanosized Fe2O3 Powders during γ- to α-PhaseTransformation. Nano Lett. 2002, 2, 245−252.(57) Azadmanjiri, J.; Simon, G. P.; Suzuki, K.; Selomulya, C.; Cashion,J. D. Phase reduction of coated maghemite (γ-Fe2O3) nanoparticlesunder microwave-induced plasma heating for rapid heat treatment. J.Mater. Chem. 2012, 22, 617−625.(58) Lai, J.; Shafi, K. V. P. M.; Loos, K.; Ulman, A.; Lee, Y.; Vogt, T.;Estournes, C. Doping γ-Fe2O3 Nanoparticles with Mn(III) Suppressesthe Transition to the α-Fe2O3 Structure. J. Am. Chem. Soc. 2003, 125,11470−11471.(59) Gallagher, K. J.; Feitknecht, W.; Mannweiler, U. Mechanism ofOxidation of Magnetite to γ-Fe2O3. Nature 1968, 217, 1118−1121.(60) Jiao, F.; Jumas, J.-C.; Womes, M.; Chadwick, A. V.; Harrison, A.;Bruce, P. G. Synthesis of ordered mesoporous Fe3O4 and gamma-Fe2O3 with crystalline walls using post-template reduction/oxidation. J.Am. Chem. Soc. 2006, 128, 12905−12909.(61) Yun, L.; Li, Y.; Zhou, C.; Lan, L.; Zeng, M.; Mao, M.; Liu, H.;Zhao, X. The formation of CuO/OMS-2 nanocomposite leads to asignificant improvement in catalytic performance for NO reduction byCO. Appl. Catal., A 2017, 530, 1−11.

ACS Applied Materials & Interfaces Research Article

DOI: 10.1021/acsami.8b13220ACS Appl. Mater. Interfaces 2018, 10, 40509−40522

40522