In situ DRIFTS study of the mechanism of low temperature selective catalytic reduction over manganese iron oxides

Similar documents
Effect of lengthening alkyl spacer on hydroformylation performance of tethered phosphine modified Rh/SiO2 catalyst

Mechanistic Study of Selective Catalytic Reduction of NOx with C2H5OH and CH3OCH3 over Ag/Al2O3 by in Situ DRIFTS

Ho modified Mn Ce/TiO2 for low temperature SCR of NOx with NH3: Evaluation and characterization

Resistance to SO2 poisoning of V2O5/TiO2 PILC catalyst for the selective catalytic reduction of NO by NH3

NO removal: influences of acidity and reducibility

Catalysis Science & Technology

In situ preparation of mesoporous Fe/TiO2 catalyst using Pluronic F127 assisted sol gel process for mid temperature NH3 selective

Tailored temperature window of MnOx CeO2 SCR catalyst by addition of acidic metal oxides

Preparation of mesoporous Fe-Cu mixed metal oxide nanopowder as active and stable catalyst for low-temperature CO oxidation

Synthesis of anisole by vapor phase methylation of phenol with methanol over catalysts supported on activated alumina

Surface reactions of CuCl2 and HY zeolite during the preparation of CuY catalyst for the oxidative carbonylation of methanol

Strategic use of CuAlO 2 as a sustained release catalyst for production of hydrogen from methanol steam reforming

Silver catalyzed three component reaction of phenyldiazoacetate with arylamine and imine

Fabrication of ultrafine Pd nanoparticles on 3D ordered macroporous TiO2 for enhanced catalytic activity during diesel soot combustion

A new approach to inducing Ti 3+ in anatase TiO2 for efficient photocatalytic hydrogen production

Ni based catalysts derived from a metal organic framework for selective oxidation of alkanes

Supporting Information

Species surface concentrations on a SAPO 34 catalyst exposed to a gas mixture

Integrating non-precious-metal cocatalyst Ni3N with g-c3n4 for enhanced photocatalytic H2 production in water under visible-light irradiation

The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA 报告人 : 沈胤

Growth of Cu/SSZ 13 on SiC for selective catalytic reduction of NO

Enhancement of the activity and durability in CO oxidation over silica supported Au nanoparticle catalyst via CeOx modification

Effect of promoters on the selective hydrogenolysis of glycerol over Pt/W containing catalysts

Investigation of the promotion effect of WO 3 on the decomposition and reactivity of NH 4 HSO 4 with NO on V 2 O 5 -WO 3 /TiO 2 SCR catalysts

Catalysis Today 153 (2010) Contents lists available at ScienceDirect. Catalysis Today. journal homepage:

A highly efficient flower-like cobalt catalyst for electroreduction of carbon dioxide

Effects of composite oxide supports on catalytic performance of Ni-based catalysts for CO methanation

Effects of Au nanoparticle size and metal support interaction on plasmon induced photocatalytic water oxidation

Photo induced self formation of dual cocatalysts on semiconductor surface

A Tunable Process: Catalytic Transformation of Renewable Furfural with. Aliphatic Alcohols in the Presence of Molecular Oxygen. Supporting Information

XPS and TPD study of NO interaction with Cu(111): Role of different oxygen species

available at journal homepage:

Selective oxidation of toluene using surface modified vanadium oxide nanobelts

Activity and selectivity of propane oxidative dehydrogenation over VO3/CeO2(111) catalysts: A density functional theory study

Selective reduction of carbon dioxide to carbon monoxide over Au/CeO2 catalyst and identification of reaction intermediate

Synthesis of PdS Au nanorods with asymmetric tips with improved H2 production efficiency in water splitting and increased photostability

Studies on Mo/HZSM-5 Complex catalyst for Methane Aromatization

Single-atom catalysis: Bridging the homo- and heterogeneous catalysis

Influence of nickel(ii) oxide surface magnetism on molecule adsorption: A first principles study

Catalytic Decomposition of Formaldehyde on Nanometer Manganese Dioxide

Preparation of LaMnO3 for catalytic combustion of vinyl chloride

Effect of acidic promoters on the titania nanotubes supported V2O5 catalysts for the selective oxidation of methanol to dimethoxymethane

Selection of oxide supports to anchor desirable bimetallic structures for ethanol reforming and 1,3 butadiene hydrogenation

Novel structured Mo Cu Fe O composite for catalytic air oxidation of dye containing wastewater under ambient temperature and pressure

Facile Synthesis and Catalytic Properties of CeO 2 with Tunable Morphologies from Thermal Transformation of Cerium Benzendicarboxylate Complexes

Zinc doped g C3N4/BiVO4 as a Z scheme photocatalyst system for water splitting under visible light

Influence of surface strain on activity and selectivity of Pd based catalysts for the hydrogenation of acetylene: A DFT study

Effect of the degree of dispersion of Pt over MgAl2O4 on the catalytic hydrogenation of benzaldehyde

Supplementary Text and Figures

Using probe molecule FTIR spectroscopy to identify and characterize Pt group metal based single atom catalysts

RKCL5155 PREPARATION AND EVALUATION OF AMMONIA DECOMPOSITION CATALYSTS BY HIGH-THROUGHPUT TECHNIQUE

Mesoporous polyoxometalate based ionic hybrid as a highly effective heterogeneous catalyst for direct hydroxylation of benzene to phenol

High performance ORR electrocatalysts prepared via one step pyrolysis of riboflavin

Synthesis of nano-sized anatase TiO 2 with reactive {001} facets using lamellar protonated titanate as precursor

Synthesis of Ag/AgCl/Fe S plasmonic catalyst for bisphenol A degradation in heterogeneous photo Fenton system under visible light irradiation

Clean synthesis of propylene carbonate from urea and 1,2-propylene glycol over zinc iron double oxide catalyst

Electronic Supplementary Information

Highly selective hydrogenation of furfural to tetrahydrofurfuryl alcohol over MIL 101(Cr) NH2 supported Pd catalyst at low temperature

Proton gradient transfer acid complexes and their catalytic performance for the synthesis of geranyl acetate

Promotional effects of Er incorporation in CeO2(ZrO2)/TiO2 for selective catalytic reduction of NO by NH3

Characterization and activity of V2O5-CeO2/TiO2-ZrO2 catalysts for NH3-selective catalytic reduction of NOx

Influence of preparation methods on the physicochemical properties and catalytic performance of MnOx CeO2 catalysts for NH3 SCR at low temperature

Deactivation of V 2 O 5 /Sulfated TiO 2 Catalyst Used in Diesel Engine for NO X Reduction with Urea

Effect of Gd0.2Ce0.8O1.9 nanoparticles on the oxygen evolution reaction of La0.6Sr0.4Co0.2Fe0.8O3 δ anode in solid oxide electrolysis cell

Photocatalytic hydrogen evolution activity over MoS2/ZnIn2S4 microspheres

SiO2 supported Au Ni bimetallic catalyst for the selective hydrogenation of acetylene

Synergetic effect between non thermal plasma and photocatalytic oxidation on the degradation of gas phase toluene: Role of ozone

Cobalt nanoparticles encapsulated in nitrogen doped carbon for room temperature selective hydrogenation of nitroarenes

available at journal homepage:

Low cost and efficient visible light driven microspheres fabricated via an ion exchange route

Growth of silver nanocrystals on graphene by simultaneous reduction of graphene oxide and silver ions with a rapid and efficient one-step approach

available at journal homepage:

A STUDY ON PRODUCTION OF OXIDANT BY DECOMPOSITION OF H 2

High-Performance Flexible Asymmetric Supercapacitors Based on 3D. Electrodes

沙强 / 讲师 随时欢迎对有机化学感兴趣的同学与我交流! 理学院化学系 从事专业 有机化学. 办公室 逸夫楼 6072 实验室 逸夫楼 6081 毕业院校 南京理工大学 电子邮箱 研 究 方 向 催化不对称合成 杂环骨架构建 卡宾化学 生物活性分子设计

room temperature: Effects of Mn loading and water content

能源化学工程专业培养方案. Undergraduate Program for Specialty in Energy Chemical Engineering 专业负责人 : 何平分管院长 : 廖其龙院学术委员会主任 : 李玉香

Molecular-Level Insight into Selective Catalytic Reduction of NO x with NH 3 to N 2

Supporting Information

Supplementary Information for

Unsupported nanoporous palladium catalyzed chemoselective hydrogenation of quinolines: Heterolytic cleavage of H2 molecule

Magnetic Co/Al2O3 catalyst derived from hydrotalcite for hydrogenation of levulinic acid to γ-valerolactone

Preparation of a fullerene[60]-iron oxide complex for the photo-fenton degradation of organic contaminants under visible-light irradiation

Very low temperature CO oxidation over colloidally deposited gold nanoparticles on Mg(OH) 2 and MgO

Pore structure effects on the kinetics of methanol oxidation over nanocast mesoporous perovskites

Sacrifical Template-Free Strategy

Supporting Information:

Wheat flour derived N doped mesoporous carbon extrudes as an efficient support for Au catalyst in acetylene hydrochlorination

An efficient and stable Cu/SiO2 catalyst for the syntheses of ethylene glycol and methanol via chemoselective hydrogenation of ethylene carbonate

Highly Sensitive, Temperature-Independent Oxygen Gas Sensor based on Anatase TiO 2 Nanoparticles-grafted, 2D Mixed Valent VO x Nanoflakelets

Preparation of Cu nanoparticles with NaBH 4 by aqueous reduction method

Enhancement of catalytic activity by homo-dispersing S2O8 2 -Fe2O3 nanoparticles on SBA-15 through ultrasonic adsorption

Pd-P nanoalloys supported on porous carbon frame as efficient catalyst for benzyl alcohol oxidation

Effect of KCl on selective catalytic reduction of NO with NH 3 over a V 2 O 5 /AC catalyst

Highly effective electrochemical water oxidation by copper oxide film generated in situ from Cu(II) tricine complex

Facile preparation of composites for the visible light degradation of organic dyes

In plasma catalytic degradation of toluene over different MnO2 polymorphs and study of reaction mechanism

Surface acidity over vanadia/titania catalyst in the selective catalytic reduction for NO removal in situ DRIFTS study

Coating Pd/Al2O3 catalysts with FeOx enhances both activity and selectivity in 1,3 butadiene hydrogenation

Plasma driven ammonia decomposition on Fe-catalyst: eliminating surface nitrogen poisoning

Transcription:

Chinese Journal of Catalysis 35 (2014) 294 301 催化学报 2014 年第 35 卷第 3 期 www.chxb.cn available at www.sciencedirect.com journal homepage: www.elsevier.com/locate/chnjc Article In situ DRIFTS study of the mechanism of low temperature selective catalytic reduction over manganese iron oxides Ting Chen, Bin Guan, He Lin *, Lin Zhu Key Laboratory for Power Machinery and Engineering of Ministry of Education, Shanghai Jiao Tong University, Shanghai 200240, China A R T I C L E I N F O A B S T R A C T Article history: Received 21 August 2013 Accepted 15 October 2013 Published 20 March 2014 Keywords: Nitrogen Oxide Ammonia Manganese iron catalysts Low temperature selective catalytic reduction In situ diffuse reflectance infrared Fourier transform spectroscopy To investigate the mechanism of selective catalytic reduction (SCR) of NOx with NH3, Ti0.9Mn0.05Fe0.05O2 δ catalyst was prepared by self propagating high temperature synthesis (SHS) method and evaluated at 25 450 C. The catalyst was characterized by X ray diffraction (XRD) and transmission electron microscopy (TEM). The possible SCR mechanism over Ti0.9Mn0.05Fe0.05O2 δ was studied by in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS). Ti0.9Mn0.05Fe0.05O2 δ showed both high SCR activity and N2 selectivity over a broad temperature window of 100 350 C. The XRD and TEM results indicated that the active components of Mn and Fe were in a highly dispersed state and in an amorphous form on TiO2. The DRIFTS results revealed that Brönsted acid sites were the active centers for NO removal and monodentate nitrates were the key intermediate in the SCR reaction. At 150 C, both Langmuir Hinshelwood and Eley Rideal mechanisms are involved in the SCR reaction, while the former one mechanism dominates the catalytic activity of Ti0.9Mn0.05Fe0.05O2 δ. Additionally, the presence of O2 significantly affects NO oxidation and coordinated NH3 activation. 2014, Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved. 1. Introduction Nitrogen oxides (NOx) are a major component of air pollution. Selective catalytic reduction (SCR) of NOx with NH3 is a well known promising technique for removing NOx [1,2]. The industrial catalysts for NH3 SCR usually comprise V2O5/TiO2 in combination with either WO3 or MoO3, however, there are also some disadvantages, such as the narrow operating temperature window (300 400 C) [3,4] and the environmental toxicity of vanadium species at high temperatures [5]. Therefore, novel high performance catalysts with transition metal alternatives to vanadium have been extensively investigated [6 11]. Mn/TiO2 anatase catalyst reported by Ettireddy et al. [4] shows high activity for low temperature SCR of NO with NH3 and shows stable NO conversion and N2 selectivity, even after 10 days of time on stream in the presence of 11 vol% water vapor at 175 C. However, the low temperature SCR performance of Mn based catalysts is reduced by poisoning from sulfur residuals in fuels and engine oils. A series of catalysts composed of manganese oxide and iron manganese oxide supported on TiO2 were studied for low temperature SCR in the presence of excess oxygen. It was found that the addition of iron oxide not only increased the NO conversion and N2 selectivity but also increased catalyst resistance to H2O and SO2 [12]. Roy et al. [8] have prepared serial Ti0.9M0.1O2 δ (M = Cr, Mn, Fe, Co, and Cu) catalysts using a self propagating high temperature synthesis (SHS) method. It was found that the optimum Ti0.9Mn0.05Fe0.05O2 δ catalyst showed good activity in the * Corresponding author. Tel: +86 21 34207774; E mail: linhe@sjtu.edu.cn This work was supported by the National Natural Science Foundation of China (51176118, 51306115) and the China Postdoctoral Science Foundation (2012M520894, 2013T60445). DOI: 10.1016/S1872 2067(12)60730 X http://www.sciencedirect.com/science/journal/18722067 Chin. J. Catal., Vol. 35, No. 3, March 2014

Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 295 150 500 C temperature range and more than 80% N2 selectivity even at 450 C. SCR catalysts prepared by the SHS method exhibited reasonably high specific surface areas, appropriate pore volumes, average pore diameters, and typical nanoparticle sizes [10], which significantly enhanced the catalyst performance. The Mn Fe based catalysts exhibit high activities at low temperature, which is possibly caused by the strong interactions of Fe and Mn. However, the SCR mechanism for this catalyst is still uncertain. For the Fe Ce Mn/ZSM 5 catalyst [13], two possible reaction pathways were proposed. One was that NO2 could react with NH4 + on Brönsted acid sites and the formed NO2[NH4 + ]2 would react with NO, producing N2 and H2O. Another pathway involves NH3 adsorption and subsequent reaction with NO or HNO2. Possible intermediates NH4NO2 and NH2NO were unstable and would decompose into N2 and H2O. Additionally, the SCR reaction over (Fe2.5Mn0.5)1 δo4 mainly follows an Eley Rideal (E R) mechanism [14]. Adsorbed ammonia species are activated to form amide species ( NH2) by Mn 4+ and Fe 3+ on the surface. Then, gaseous NO is reduced by NH2 on the surface to form N2 and H2O. For the case of the Ti0.9Mn0.05Fe0.05O2 δ prepared by Roy et al. [8], according to a Langmuir Hinshelwood (L H) mechanism, the adsorbed NH3 or NH4 + can react with adsorbed NO to form an NH2NO adduct, which can then dissociate into N2 and H2O. Overall, the process of adsorption and activation of reactant NH3 or NO is important because it is related to the intermediate species and determines the reaction pathways. The purpose of this study is to obtain information on the reaction mechanism of manganese and iron oxide catalysts at low temperatures through the respective adsorption of NH3 and NO, as well as the reaction between NH3 and NO. SHS is an effective, low cost method for production of various industrially useful materials [15]. In this paper, an SCR catalyst, Ti0.9Mn0.05Fe0.05O2 δ was also prepared by the SHS method [16]. The catalyst sample was exposed to simulated exhaust gas mixtures (NO/O2/NH3/N2) to evaluate the SCR performance and was characterized by in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) to investigate the possible mechanism of NH3 SCR. 2. Experimental 2.1. Catalyst preparation Catalyst Ti0.9Mn0.05Fe0.05O2 δ was prepared by the SHS method. First, the required amount of absolute alcohol was added dropwise into the precursor tetrabutyl titanate (Ti(OCH3(CH2)3)4), and deionized water was likewise added to the product of metatitanic acid (TiO(OH)2). Next, in order to get titanyl nitrate (TiO(NO3)2), the right amount of HNO3 solution was added. At the same time, a solution of metal nitrate (Mn(NO3)2 4H2O and Fe(NO3)3 9H2O) in deionized water was prepared in stoichiometric amounts. Then, the nitrate solution was combined with the titanyl nitrate. Glycine (CH2NH2COOH) was sequentially added to fuel the reaction. After stirring at 50 C for 1 h, the aqueous solution was placed in a muffle furnace at 350 C in air. During the first few minutes, the solution boiled with froth and combusted with smoke. The resulting product was finally calcined at the ignition temperature for 1 h. The synthesized catalyst was allowed to cool and was sieved to 40 80 mesh for further analysis. 2.2. Catalytic activity tests Catalytic activity tests were carried out using the temperature programmed reaction (TPR) method. The powder sample (0.5 ml) was used in a fixed bed quartz reactor (i.d. 9 mm), and the reaction temperature was raised from room temperature to 450 C at a ramp rate of 5 C/min. A simulated exhaust gas mixture (0.06% NO, 0.06% NH3, 5% O2, and N2 for balance) was fed to the catalyst sample through a set of mass flow controllers. The total flow rate was set to 100 ml/min, corresponding to a gas hourly space velocity (GHSV) of 12000 h 1. The outlet gas composition containing NO, NO2, N2O, and NH3 was measured on line with an FTIR gas analyzer (Thermo Nicolet 6700). Before the experiment, concentrations of the gases were quantified based on the classic least squares methodology using the Omnic Quantpad software. A calibration run of each gas consisting of four spectral absorption standards was performed and used to generate each standard curve within Quantpad. 2.3. Catalyst characterization X ray diffraction (XRD) was used to reveal the catalyst crystal phases. XRD continuous measurements were performed on a Rigaku D/max 2200/PC system. Diffraction patterns were obtained from 20 to 90 (2θ) with a scanning rate of 5 /min, and a step size of 0.02 at ambient temperature. The microscopic morphology of catalyst was interrogated by transmission electron microscopy (TEM) on a JEOL JEM 2010 analytical electron microscope operating at an accelerating voltage of 200 kv. Before TEM characterization, the catalyst powder was ultrasonically suspended in alcohol for 10 min, after which the obtained suspension was deposited on copper grid supported amorphous carbon films. The impregnated mesh was allowed to dry in air for TEM analysis. 2.4. In situ DRIFTS studies In situ DRIFTS spectra were measured by an FTIR gas analyzer (Thermo Nicolet 6700) equipped with a diffuse reflectance optics accessory (HVC DRP 4). The inlet gas concentrations were identical to those used in the catalytic activity test (0.06% NO, 0.06% NH3, 5% O2, and N2 for balance). A 60 mg of powder sample was used, and the total flow rate was 50 ml/min. To clean the catalyst surface, the sample was purged in N2 at 450 C for 2 h prior to acquisition of DRIFTS spectra. Thereafter, the background spectra were recorded from 50 to 450 C, with 50 C increments. In this experiment, the IR range was from 4000 to 650 cm 1, and the resolution was set to 4 cm 1. 3. Results and discussion

296 Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 3.1. Ti0.9Mn0.05Fe0.05O2 δ catalyst activity TPR was performed to evaluate the SCR performance of the Ti0.9Mn0.05Fe0.05O2 δ catalyst, as shown in Fig. 1. The calculation methods to determine NOx conversion, N2O selectivity, and N2 selectivity were described in our earlier paper [10]. More than 80% of NO is reduced by NH3 on the Ti0.9Mn0.05Fe0.05O2 δ catalyst in a wide temperature window of 100 to 350 C (Fig. 1(a)). Ti0.9Mn0.05Fe0.05O2 δ reaches the maximum NO conversion of 100% at an especially low temperature of 120 C, and this NO conversion rate remains at its peak until 350 C. NO conversion subsequently decreases sharply above 350 C. The main reason for this is unselective NH3 oxidation by O2. The N2 yield initially reaches over 95% at ambient temperature to 100 C, but then decreases to a lower level of around 70% at 250 C (Fig. 1(b)). With further rise in temperature, the N2 yield again increases. From 350 C, N2 selectivity decreases monotonically due to oxidation of NH3. This can be seen from the NO2 concentration curve in Fig. 1(b). At the same time, under 250 C, N2O selectivity shows the reverse trend as N2 selectivity. A fast drop in selectivity happens between 250 and 300 C. N2O selectivity then decreases slightly with increasing temperature. Overall, in a broad temperature window of 100 350 C, Ti0.9Mn0.05Fe0.05O2 δ has high SCR activity, which is comparable to that reported by Roy et al. [8]. 3.2. Catalyst characterization The power XRD pattern of the Ti0.9Mn0.05Fe0.05O2 δ catalyst sample is shown in Fig. 2, with peaks indicated as identified in the ICDD (anatase TiO2 phase: PDF #21 1272). The XRD peaks of the catalyst closely conform to anatase phase TiO2, with no evidence for the rutile phase. No diffraction peaks for FeOx, MnOx, or any other mixed oxides are seen, which indicates that Intensity Anatase TiO 2 20 30 40 50 60 70 80 90 2 /( o ) Fig. 2. Power XRD pattern of Ti0.9Mn0.05Fe0.05O2 δ catalyst. the Mn and Fe are either in a highly dispersed state and assumes an amorphous form on TiO2, or that the formed crystallites cannot be detected by XRD. This is due to strong Fe Mn interactions, which leads to enhanced dispersion and lower crystallization for both [12]. TEM was used to examine the morphology and size of the Ti0.9Mn0.05Fe0.05O2 δ. As presented in Fig. 3, there are many nanoparticles dispersed on the catalyst, and no obvious particle agglomeration is apparent. The TEM image shows that the typical nanoparticles are well dispersed with sizes on the order of 10 nm. At higher magnification, a crystal plane spacing of 0.346 nm is detected, which is slightly smaller than the spacing of the anatase TiO2 (101) plane. Any morphology of the manganese and iron oxides are not observed, which is consistent with the XRD result. NO conversion (%) Selectivity (%) 100 90 (a) 80 70 60 50 40 30 20 10 1000 0.010 90 (b) 0.009 80 0.008 70 0.007 60 N 2 selectivity 0.006 50 N 2O selectivity 0.005 40 NO 2 concentration 0.004 30 0.003 20 0.002 10 0.001 0 0 100 200 300 400 0.000 500 Temperature ( o C) NO2 concentration (%) Fig. 1. NO conversion (a) and N2 selectivity, N2O selectivity, and NO2 concentration (b) over Ti0.9Mn0.05Fe0.05O2 δ. Reaction conditions: 0.06% NO, 0.06% NH3, 5% O2, N2 as balance gas, total gas flow rate 100 ml/min, and GHSV = 12000 h 1. Fig. 3. TEM images of Ti0.9Mn0.05Fe0.05O2 δ sample.

Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 297 3.3. In situ DRIFTS 3.3.1. Adsorption of NH3 NH3 adsorption measurements were performed in a feed of 0.06% NH3 in N2 at 50 C for 30 min. Catalyst Ti0.9Mn0.05Fe0.05O2 δ was then purged with N2 for 30 min. Surface adsorption species with IR bands at 1188 and 1598 cm 1 are formed (Fig. 4), which are assigned to coordinated NH3 adsorbed on Lewis acid sites [10]. In the NH vibration region, IR bands at 3160, 3253, and 3334 cm 1 are seen and are also assigned to coordinated NH3 bound to Lewis acid sites. The band at 1306 cm 1 is probably due to NH3 adsorbed on different Lewis sites [17]. At the same time, IR bands observed at 1458 and 1680 cm 1 are from asymmetric and symmetric deformations of NH4 + on Brönsted acid sites. It seems that there are less Brönsted acid sites than Lewis acid sites on the Ti0.9Mn0.05Fe0.05O2 δ catalyst. It can also be seen that the intensity of the bands at 1188 and 1598 cm 1 are far stronger than those at 1458 and 1680 cm 1. Additionally, the two stronger bands reach their peak intensities after NH3 is saturated on the catalyst at 50 C. All surface adsorbed NH3 species are quite stable during subsequent purging with N2 at 50 C. The spectra of these surface species at different temperatures are also presented in Fig. 4. With increasing temperature, the intensity of the bands at 1188 and 1598 cm 1 drops off gradually. At 450 C, these bands still obviously exist. The intensities of the bands in the 3400 3100 cm 1 region also decrease slightly. Meanwhile, the intensities of the bands at 1458 and 1680 cm 1 decline drastically, and the band at 1458 cm 1 can no longer be observed above 200 C. These results demonstrate that NH3 desorbs at increased temperatures. Moreover, Brönsted acid sites have lower thermal stability than Lewis acid sites. 3.3.2. Co adsorption of NH3 + O2 In this experiment, the Ti0.9Mn0.05Fe0.05O2 δ catalyst is treated with NH3 + O2 in the same way. Both NH4 + ions and coordinated 3253 3334 3160 1188 1680 1598 1458 1350 450 o C 400 o C 350 o C 300 o C 250 o C 200 o C 150 o C 100 o C 50 o C 3500 3000 2500 2000 1500 1000 Fig. 4. In situ DRIFTS spectra of Ti0.9Mn0.05Fe0.05O2 δ during NH3 adsorption at different temperatures. Reaction conditions: 0.06% NH3, N2 as balance, total gas flow rate 50 ml/min. NH3 are formed (Fig. 5). The intensities of the coordinated NH3 bands at 1188 and 1598 cm 1 decrease with increasing temperature as NH3 desorbs. However, when compared to the spectra of NH3 adsorbed on Lewis acid sites in Fig. 4, the variations between spectra at different temperatures are larger. The intensities of the bands at 1188 and 1598 cm 1 initially decrease gradually beginning at 150 C. When reaching a temperature of 450 C, the coordinated NH3 almost vanishes, without any obvious intensity. According to the literature [18], this may be caused by oxidation of NH3 to NH2 at a relatively low temperature range and oxidation of NH3 to N2 at a relatively high temperature. Additionally, a new band begins to appear at 1245 cm 1 at 400 C, which was attributed to a bridging nitrate. This nitrate could be due to side reactions taking place, yielding a certain amount of NO. 3.3.3. Adsorption of NO NO adsorption measurements were performed in a feed of 0.06% NO in N2 for 30 min. The catalyst was then purged with N2 for 30 min. Surface nitrate species are observed at 1245, 1276, 1470, 1576, and 1612 cm 1 (Fig. 6). The IR band at 1245 cm 1 can be assigned to a bridging nitrate, and that at 1276 cm 1 and 1470 cm 1 are assigned to monodentate nitrate, while bands at 1576 cm 1 are attributed to bidentate nitrate. According to results of Long et al. [17], the band at 1612 cm 1 is not easy to assign. They assigned it to an NO2 adspecies (nitro or adsorbed NO2 molecule) on the catalyst, because this species had a different thermal stability from that of other nitrate species. It can be also seen that the species appeared at 1612 cm 1 exhibits a similar thermal stability as the bridging nitrate. Thus, we assign it to the bridging nitrate. IR bands in these regions decrease with increasing temperature (Fig. 6), indicating desorption of NO. IR bands at 1276 and 1470 cm 1 show the most rapid disappearance, occurring at 200 C. The intensities of the bands at 1245 and 1612 cm 1 3253 3334 3160 sample with NH 3 + O2 (50 o C) 450 o C 400 o C 350 o C 300 o C 250 o C 200 o C 150 o C 100 o C 50 o C 1598 12451188 1680 1350 1458 3500 3000 2500 2000 1500 1000 Fig. 5. In situ DRIFTS spectra of Ti0.9Mn0.05Fe0.05O2 δ during NH3 + O2 adsorption at different temperatures. Reaction conditions: 0.06% NH3, 5% O2, N2 as balance, and total gas flow rate 50 ml/min.

298 Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 50 o C 100 o C 150 o C 200 o C 250 o C 300 o C 350 o C 400 o C 450 o C 1612 1276 1576 1245 1470 3500 3250 3000 1750 1500 1250 1000 Fig. 6. In situ DRIFTS spectra for adsorption of NO over Ti0.9Mn0.05Fe0.05O2 δ at different temperatures. Reaction conditions: 0.06% NO, N2 as balance, and total gas flow rate 50 ml/min. decrease slightly slower, such that they can still be observed at 350 C. The bidentate nitrate at 1576 cm 1 does not disappear until 450 C. The thermal stability of the nitrate species follows the order: bidentate nitrate > bridging nitrate > monodentate nitrate. To investigate the influence of gaseous O2, we carried out the adsorption of NO and NO + O2 over Ti0.9Mn0.05Fe0.05O2 δ at 150 C. First, the catalyst was treated with NO and NO + O2 in N2 for 30 min, and then was purged with N2 for an extended 90 min. The recorded IR spectra are shown in Fig. 7. It is obvious that NO + O2 adsorption is much stronger than NO adsorption because gaseous O2 could enhance the oxidation of NO to NO2, leading to the formation of surface nitrate species. In the NO only treatment case, oxidation of NO is simply caused by lattice oxygen. After purging with N2 for 90 min, only a slight decrease in the intensities of the surface nitrate species bands is seen in the two cases, which implies that surface NO adsorption species are caused by fairly stable chemisorption, regardless of O2. This is consistent with results reported by Centi et al. [19], who proposed that chemisorption of NO on the catalyst copper sites was significant and that chemisorption is enhanced by the presence of oxygen. 3.3.4. Reaction studies under transient conditions 3.3.4.1. Reaction between only NO and adsorbed NH3 To investigate the reaction mechanism of SCR by NH3 over the Ti0.9Mn0.05Fe0.05O2 δ catalyst, the sample was first exposed to NH3 at 150 C until it was saturated. The catalyst was then purged with N2 and subsequently exposed to NO. IR spectra were recorded as a function of time (Fig. 8). Because of exposure to NH3, IR bands assigned to coordinated NH3 or NH4 + appear. However, after 70 min of NO exposure, no apparent decrease in the adsorbed NH3 is seen, and no surface NO adsorption species are seen to accumulate on the catalyst. At the same time, we could not find evidence for the formation of any new adsorption species. This can be explained from several aspects [17,20,21]. First, the absence of gaseous O2 inhibits NO oxidation, suppressing the formation of nitrate species. Additionally, our observations are possibly caused by competition between NO and NH3 for the same adsorption sites on the catalyst surface. Moreover, gaseous NO has a low reaction probability with adsorbed NH3, owing to lack of gaseous O2 in the reaction conditions. The activation of adsorbed NH3 by gaseous O2 to form NH2 is a main step in the catalytic mechanism. 3.3.4.2. Reaction between gas mixtures NO + O2 and adsorbed NH3 For the case of reaction between NO + O2 and adsorbed NH3, the catalyst sample was first exposed to NH3 at 150 C until it was saturated, then purged with N2. After that, NO+O2 was introduced into the cell. The results are shown in Fig. 9. IR bands at 1188, 3160, 1598, and 3253 cm 1, assigned to coordinated NH3 on Lewis acid sites as noted above, decease gradually with time until they disappear. The vibrational signature of NH4 + on Brönsted acid sites at 1680 cm 1 vanishes quickly after exposure to NO for about 25 min, indicating that NO rapidly reacts with adjacent adsorbed NH4 + 25 min later. Thus, it is suggested that both Brönsted and Lewis acid sites are involved in NO reduction. Purged for 90 min Sample with NO+O 2 Purged for 90 min Sample with NO 1612 1276 1576 1245 1486 1492 3500 3000 2500 2000 1500 1000 Fig. 7. In situ DRIFTS spectra for adsorption of NO over Ti0.9Mn0.05Fe0.05O2 δ with and without O2 at 150 C. Reaction conditions: 0.06% NO, 5% O2, N2 as balance, total gas flow rate 50 ml/min. 3334 3253 3160 70 min 50 min 35 min 25 min 15 min 10 min 5 min Sample with NH 3 (30 min) 3500 3000 2500 2000 1500 1000 1603 1680 1458 1350 1188 Fig. 8. In situ DRIFTS spectra of Ti0.9Mn0.05Fe0.05O2 δ during transient NO exposure after NH3 adsorption at 150 C. Reaction conditions: 0.06% NO, N2 as balance, and total gas flow rate 50 ml/min.

Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 299 3334 3253 3160 1680 1598 14581350 1245 1188 45 min 1576 40 min 35 min 30 min 25 min 20 min 15 min 10 min 5 min Sample with NH 3 (30 min) 3400 3200 1800 1600 1400 1200 1000 Fig. 9. In situ DRIFTS spectra of Ti0.9Mn0.05Fe0.05O2 δ during transient NO + O2 exposure after NH3 adsorption at 150 C. Reaction conditions: 0.06% NO, 5% O2, N2 as balance, total gas flow rate 50 ml/min. Some new species appear after 25 min of reaction. The band at 1245 cm 1 is assigned to low frequency vibrations of a bridging nitrate, and the band at 1576 cm 1 is assigned to bidentate nitrate. The intensities of the bands at 1245 and 1576 cm 1 increase with time after their initial appearance. However, signals for the monodentate nitrate species at 1276 and 1470 cm 1 do not emerge, indicating that the monodentate nitrate on the Ti0.9Mn0.05Fe0.05O2 δ surface reacts with NH3 immediately preventing detectable concentrations to arise. The bridging nitrate and bidentate nitrate seem to be inactive for the SCR reaction. Hence, the monodentate nitrate is suggested to be a key intermediate. Furthermore, it can be seen that the rate of SCR reaction becomes faster after 25 min. For the competitive adsorption of NO and NH3 on the catalyst surface, no nitrate species formation is seen in the first 25 min. Coordinated NH3 is first activated by active oxygen from the dehydration between manganese oxides, yielding the amide species from H abstraction [21]. At this point, the intermediate species NH2 further reacts with gaseous NO to form N2 and H2O following an E R mechanism. Then as the adsorbed NH3 species are consumed, sufficient adsorption sites are available for NO. The SCR process then begins to follow a L H mechanism, in which adsorbed NH4 + and coordinated NH3 react with nitrate species. Comparing with the results presented in Section 3.3.4.1., it is clear that oxygen plays an important role in the reaction between NO and NH3 [21], as a result of the oxidation of NH3 and NO. At a low temperature (150 C), it is difficult to form NH2 species because of the high activation energy [20]. Conversely, NO oxidation occurs a little faster by the high oxidative ability of Mn. As more nitrate species adsorb, the reaction rate is accelerated. This is why the catalysts with Mn have outstanding low temperature performance. As a result, at 150 C both L H and E R mechanisms are involved in the SCR reaction, while the L H mechanism is dominant for the Ti0.9Mn0.05Fe0.05O2 δ catalyst. 3.3.4.3. Reaction between NH3 and adsorbed NO + O2 To investigate the reaction between NH3 and adsorbed gas mixture NO + O2, the Ti0.9Mn0.05Fe0.05O2 δ catalyst sample was first treated with gas mixtures NO + O2 at 150 C until it was saturated, followed by N2 purging. When the treated sample is exposed to NH3, the intensities of the bands at 1245, 1276, 1576, and 1612 cm 1, assigned to surface nitrate species, decrease (Fig. 10). The band intensity for the bridging nitrate at 1245 cm 1 decreases quickly, disappearing within 15 min, while the band intensity of the monodentate nitrate at 1276 cm 1 decreases slightly slower, disappearing after about 35 min of NH3 exposure. The bridging nitrate band at 1612 cm 1 can still be observed after 40 min. Exposure to NH3 results in little change to the band intensity of the bidentate nitrate at 1576 cm 1. These intensity changes for the nitrate species indicate that monodentate and bridging nitrates have high reactivities with NH3, while the bidentate nitrate seems to be little reactive towards NH3. This is consistent with the results of reaction between NO + O2 and adsorbed NH3. Meanwhile, IR bands at 1188, 1230, 1295, 1458, 1680, 3160, 3253, and 3334 cm 1 emerge, belonging to a group of surface NH3 adsorption complexes on the Ti0.9Mn0.05Fe0.05O2 δ catalyst. The intensity of all these bands increases quickly within about 10 min, because surface acid sites have a strong adsorption capacity of NH3 due to the effect of the active Fe metal [20]. The intensity of the band at 1188 cm 1 remains strong after 10 min exposure to NH3, suggesting continuous accumulation of coordinated NH3 on the catalyst. However, almost no changes are seen in the ionic NH4 + bands at 1458 and 1680 cm 1 after 10 min. New bands at 1230 and 1295 cm 1 are detected, and they increase with time in NH3 flow. So these bands could not be attributed to surface nitrate species. In contrast to the results of adsorbed NH3 shown in Fig. 4, no adsorbed species are seen at 1230 and 1295 cm 1 when the catalyst is first treated with NH3 and then purged by N2. Therefore, we assigned these bands to 3334 3253 3160 40 min 35 min 30 min 25 min 20 min 15 min 10 min 5 min Sample with NO+O 2 1216 1245 1188 1576 1276 1295 1680 1612 1458 3400 3200 1800 1600 1400 1200 1000 Fig. 10. In situ DRIFTS spectra of NH3 SCR reaction over Ti0.9Mn0.05Fe0.05O2 δ during transient NH3 exposure after NO + O2 co adsorption at 150 C. Reaction conditions: 0.06% NH3 and N2 as balance, total gas flow rate 50 ml/min.

300 Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 weak adsorbed NH3 species. On the basis of the above analysis, the L H mechanism is dominant for the Ti0.9Mn0.05Fe0.05O2 δ catalyst at 150 C, and both adsorbed NH4 + and coordinated NH3 participate in the L H reaction path. During the reaction of NH3 and adsorbed NO + O2, surface NH4 + on Brönsted acid sites remain stable, maintaining their intensity, probably because the rate of NH3 adsorption on Brönsted acid sites is equal to the rate of NH4 + reaction with NO. The amount of coordinated NH3 on Lewis acid sites continues to increase due to the relatively slow rate of reaction between NO and coordinated NH3. This suggests that Brönsted acid sites are more active than Lewis acid sites on the Ti0.9Mn0.05Fe0.05O2 δ catalyst. According to the results of in situ DRIFTS, the reaction mechanism can be deduced as following the reaction scheme depicted in equations (1) (4): n+ + n+ + M O NO 2+2NH4 M O NO 2[NH 4 ] 2 (1) n+ n+ M O NO 2+2NH 3(a) M O NO 2[NH 3] 2 (2) n+ + + n+ M O NO 2[NH 4 ] 2+NO 2N 2+3H2O+2H +M O (3) n+ n+ M O NO 2[NH 3] 2+NO N 2+3H2O+M O (4) The formation of monodentate nitrate M n+ O NO2 (M = Fe and Mn) is involved in reaction with adsorbed NH4 + or NH3 on neighboring acid sites, to produce an active intermediate M n+ O NO2[NH4 + ]2 or M n+ O NO2[NH3]2, which finally reacts with gaseous NO to form N2 and H2O [17,20]. 4. Conclusions The Ti0.9Mn0.05Fe0.05O2 δ catalyst shows high SCR performance over a broad temperature window of 100 350 C. The active components of Mn and Fe exist in a highly dispersed state and amorphous form on TiO2. In situ DRIFTS measurements suggest that Brönsted and Lewis acid sites both take part in the SCR reaction, and Brönsted acid sites may play a more important role in determining the activity of Ti0.9Mn0.05Fe0.05O2 δ. The monodentate nitrate species is key active intermediate, in comparison to the less important bridging and bidentate nitrate species. At 150 C, both L H mechanism and E R mechanisms are involved in the SCR reaction, while the L H mechanism dominates catalytic activity of Ti0.9Mn0.05Fe0.05O2 δ. The presence of O2 significantly affects NO oxidation and coordinated NH3 activation. Hence, at a low temperature, NO oxidation is a main step, which depends on the oxidative ability of the active metal. References [1] Balle P, Geiger B, Kureti S. Appl Catal B, 2009, 85: 109 [2] Devadas M, Kröcher O, Elsener M, Wokaun A, Mitrikas G, Söger N, Pfeifer M, Demel Y, Mussmann L. Catal Today, 2007, 119: 137 [3] Balle P, Geiger B, Klukowski D, Pignatelli M, Wohnrau S, Menzel M, Zirkwa I, Brunklaus G, Kureti S. Appl Catal B, 2009, 91: 587 [4] Ettireddy P R, Ettireddy N, Mamedov S, Boolchand P, Smirniotis P G. Appl Catal B, 2007, 76: 123 [5] Domingo J L. Reprod Toxicol, 1996, 10: 175 [6] Liu F D, He H, Ding Y, Zhang C B. Appl Catal B, 2009, 93: 194 [7] Liu F D, He H, Zhang C B, Feng Z C, Zheng L R, Xie Y N, Hu T D. Appl Catal B, 2010, 96: 408 [8] Roy S, Viswanath B, Hegde M S, Madras G. J Phys Chem C, 2008, 112: 6002 [9] Larrubia M A, Ramis G, Busca G. Appl Catal B, 2001, 30: 101 [10] Guan B, Lin H, Zhu L, Huang Z. J Phys Chem C, 2011, 115: 12850 [11] Biabani A, Rezaei M, Fattah Z. J Nat Gas Chem, 2012, 21: 415 [12] Qi G, Yang R T. Appl Catal B, 2003, 44: 217 [13] Zhou G Y, Zhong B C, Wang W H, Guan X J, Huang B C, Ye D Q, Wu H J. Catal Today, 2011, 175: 157 [14] Yang S J, Wang C Z, Li J H, Yan N Q, Ma L, Chang H Z. Appl Catal B, 2011, 110: 71 [15] Aruna S T, Mukasyan A M. Curr Opin Solid State Mater Sci, 2008, 12: 44 [16] Zhu L, He L Guan B, Huang Z. Vehicle Engine ( 朱霖, 林赫, 管斌, 黄 Chin. J. Catal., 2014, 35: 294 301 Graphical Abstract doi: 10.1016/S1872 2067(12)60730 X In situ DRIFTS study of the mechanism of low temperature selective catalytic reduction over manganese iron oxides Ting Chen, Bin Guan, He Lin *, Lin Zhu Shanghai Jiao Tong University Langmuir Hinshelwood mechanism N 2 +H 2 O NO NO+O 2 NO 2 *-O-NO 2 *-O-NO 2 [NH 3 ] 2 NH 3 (ad) NH 3 *:active metal ions *-O-NO 2 [NH 4 ] + 2 NO N 2 +H 2 O NH 4 + Selective catalytic reduction reactions over Ti0.9Mn0.05Fe0.05O2 δ catalysts possibly proceed according to both Langmuir Hinshelwood and Eley Rideal mechanisms, but NO more likely reacts as an adsorbed species.

Ting Chen et al. / Chinese Journal of Catalysis 35 (2014) 294 301 301 震. 车用发动机 ), 2012, (3): 72 [17] Long R Q, Yang R T. J Catal, 2000, 190: 22 [18] Kijlstra W S, Brands D S, Smit H I, Poels E K, Bliek A. J Catal, 1997, 171: 219 [19] Centi G, Perathoner S. J Catal, 1995, 152: 93 [20] Liu F D, He H, Zhang C B, Shan W P, Shi X Y. Catal Today, 2011, 175: 18 [21] Jiang B Q, Li Z G, Lee S C. Chem Eng J, 2013, 225: 52 原位漫反射傅里叶变换红外光谱研究锰铁基催化剂上低温选择性催化还原反应机理 陈婷, 管斌, 林赫 *, 朱霖上海交通大学动力机械及工程教育部重点实验室, 上海 200240 摘要 : 采用自蔓延燃烧法制备了 Ti 0.9 Mn 0.05 Fe 0.05 O 2-δ 催化剂, 运用原位漫反射傅里叶变换红外光谱对该催化剂的 NO 和 NH 3 稳态吸附以及 NO 和 NH 3 瞬态反应进行了详细地分析与讨论. 结果表明, 相比于 Lewis 酸性位, 150 ºC 时 Brönsted 酸性位吸附的 NH 3 更具有 SCR 活性 ; 与双齿硝酸盐和桥式硝酸盐相比, NO 吸附产生的单齿硝酸盐是主要的中间物种 ; 该 SCR 反应遵循 Eley-Rideal 和 Langmuir-Hinshelwood 机理, 但以后者为主. 另外, O 2 的存在有利于 NO 的氧化和配位态 NH 3 的活化. 关键词 : 氮氧化物 ; 氨 ; 锰铁基催化剂 ; 低温选择性催化还原 ; 原位漫反射傅里叶变换红外光谱 收稿日期 : 2013-08-21. 接受日期 : 2013-10-15. 出版日期 : 2014-03-20. * 通讯联系人. 电话 : (021)34207774; 电子信箱 : linhe@sjtu.edu.cn 基金来源 : 国家自然科学基金 (51176118, 51306115); 中国博士后科学基金 (2012M520894, 2013T60445). 本文的英文电子版由 Elsevier 出版社在 ScienceDirect 上出版 (http://www.sciencedirect.com/science/journal/18722067).