In-situ observation of nitrogen monoxide adsorption on perovskite-type MTiO 3 (M =Sr, Ba)

Similar documents
A Highly efficient Iron doped BaTiO 3 nanocatalyst for the catalytic reduction of nitrobenzene to azoxybenzene

INVESTIGATION OF SURFACE CHEMISTRY PROPERTIES OF Ga 2 O 3 /Al 2 O 3 CATALYSTS BY FT-IR SPECTROSCOPY

The Effect of Simultaneous Homo- and Heterogeneous Doping on Transport Properties of Ba 2 In 2 O 5

NO removal: influences of acidity and reducibility

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

CuH-ZSM-5 as Hydrocarbon Trap under cold. start conditions

Supplementary Information for

Supporting Information

Supporting Information

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

Supplementary Text and Figures

photo-mineralization of 2-propanol under visible light irradiation

ph-depending Enhancement of Electron Transfer by {001} Facet-Dominating TiO 2 Nanoparticles for Photocatalytic H 2 Evolution under Visible Irradiation

Room Temperature Hydrogen Generation from Hydrous Hydrazine for Chemical Hydrogen Storage

Supporting Information

efficient wide-visible-light photocatalysts to convert CO 2 and mechanism insights

Sulfur-bubble template-mediated synthesis of uniform porous g-c 3 N 4 with superior photocatalytic performance

Electronic Supplementary Information

Supporting Information

Supporting Information for. Size-Dependent Oxidation State and CO Oxidation Activity of Tin

Structural Analysis and Dielectric Properties of Cobalt Incorporated Barium Titanate

Insights into Interfacial Synergistic Catalysis over Catalyst toward Water-Gas Shift Reaction

Pioneering In Situ Recrystallization during Bead Milling: A Top-down Approach to Prepare Zeolite A Nanocrystals

Supporting Information

Electronic Supplementary Information (ESI)

Supporting Information. CdS/mesoporous ZnS core/shell particles for efficient and stable photocatalytic hydrogen evolution under visible light

Supporting Information

Supplementary Figure 1 Result from XRD measurements. Synchrotron radiation XRD patterns of the as-prepared gold-ceria samples.

Supporting Information. High Selectivity of Supported Ru Catalysts in the Selective. CO Methanation - Water Makes the Difference

Determination of Metal Dispersion of Pt/CeO2 Catalyst by CO-pulse Method

Evidence for structure sensitivity in the high pressure CO NO reaction over Pd(111) and Pd(100)

Supporting Information

Electronic Supporting Information (ESI) Porous Carbon Materials with Controllable Surface Area Synthsized from Metal-Organic Frameworks

Relaxor characteristics of ferroelectric BaZr 0.2 Ti 0.8 O 3 ceramics

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

Electronic Supplementary Information (ESI) From metal-organic framework to hierarchical high surface-area hollow octahedral carbon cages

Dry-gel conversion synthesis of Cr-MIL-101 aided by grinding: High surface area high yield synthesis with minimum purification

Supporting Information

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

Supporting Information

Supporting Information

Effects of Crystal Structure on Microwave Dielectric Properties of Ceramics

Supporting Information

DEVELOPMENT OF CATALYSTS FOR ETHANE EPOXIDATION REACTION. Keywords: Ethylene oxide, Partial oxidation, Ethane epoxidation, Second metal.

Supporting Information. Nanoscale Kirkendall Growth of Silicalite-1 Zeolite Mesocrystals with. Controlled Mesoporosity and Size

dissolved into methanol (20 ml) to form a solution. 2-methylimidazole (263 mg) was dissolved in

Microstructure evolution during BaTiO 3 formation by solid-state reactions on rutile single crystal surfaces

Supporting Information

Two Dimensional Graphene/SnS 2 Hybrids with Superior Rate Capability for Lithium ion Storage

High compressibility of a flexible Metal-Organic-Framework

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

Supporting Information

Catalytic Decomposition of Formaldehyde on Nanometer Manganese Dioxide

Role of Re and Ru in Re Ru/C Bimetallic Catalysts for the

Electronic Supplementary Information. Precursor Salt Assisted Syntheses of High-Index Faceted Concave Hexagon and Nanorod like Polyoxometalates

One-pass Selective Conversion of Syngas to para-xylene

Supporting Information. Solution-Based Growth of Monodisperse Cube-Like BaTiO 3 Colloidal Nanocrystals

Supporting Information. Integration of accessible secondary metal sites into MOFs for H 2 S removal

Methane production from CO2 over Ni-Hydrotalcite derived catalysts

Shigeki Yonezawa*, Yuji Muraoka and Zenji Hiroi Institute for Solid State Physics, University of Tokyo, Kashiwa, Chiba

High-Flux CO Reduction Enabled by Three-Dimensional Nanostructured. Copper Electrodes

Precious Metal-free Electrode Catalyst for Methanol Oxidations

Atmospheric Analysis Gases. Sampling and analysis of gaseous compounds

Supplementary Information. Large Scale Graphene Production by RF-cCVD Method

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

EXECUTIVE SUMMARY. especially in last 50 years. Industries, especially power industry, are the large anthropogenic

Supporting Information for: Three-Dimensional Cuprous Oxide Microtube Lattices with High Catalytic

Improvement of CO 2 Absorption Properties of Limestone Ore by the Addition of Reagent Grade-SiO 2 and Natural Diatomite

Co-Ni/Al 2 O 3 catalysts for CO 2 methanation at atmospheric pressure

Please do not adjust margins. Flower stamen-like porous boron carbon nitride nanoscrolls for water cleaning

Supplementary Information. ZIF-8 Immobilized Ni(0) Nanoparticles: Highly Effective Catalysts for Hydrogen Generation from Hydrolysis of Ammonia Borane

Supporting Information. Re-Investigation of the Alleged Formation of CoSi Nanoparticles on Silica. Van An Du, Silvia Gross and Ulrich Schubert

Selective aerobic oxidation of biomass-derived HMF to 2,5- diformylfuran using a MOF-derived magnetic hollow Fe-Co

Cu 2 graphene oxide composite for removal of contaminants from water and supercapacitor

MOHAMED R. BERBER Department of Chemistry, Faculty of Science, Tanta University, Tanta 31527, Egypt.

SUPPORTING INFORMATION

Panadda Sittiketkorn, Sarawut Thountom and Theerachai Bongkarn*

Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS) for In-Situ Analysis of Solid Oxide Electrolysis Cells

KMUTNB Int J Appl Sci Technol, Vol. 9, No. 4, pp , 2016

Synthesis of Mesoporous ZSM-5 Zeolite Crystals by Conventional Hydrothermal Treatment

Investigation on microstructure and dielectric behaviour of (Ba x Gd Cr x )TiO 3 ceramics

Electronic Supplementary Information (ESI) Green synthesis of shape-defined anatase TiO 2 nanocrystals wholly exposed with {001} and {100} facets

A high-efficient monoclinic BiVO 4 adsorbent for selective capture toxic selenite

Supporting Information

Supporting Information

Electronic Supplementary Information

EFFECT OF MORPHOLOGY OF NANOSTRUCTURED CERIA-BASED CATALYSTS OVER CO, SOOT AND NO OXIDATIONS

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

Controlling Interfacial Contact and Exposed Facets for. Enhancing Photocatalysis via 2D-2D Heterostructure

Engineering electronic structure of Two-Dimensional Subnanopore. nanosheet by Molecular Titanium-oxide Incorporation for Enhanced

Experimental Section

Shape Assisted Fabrication of Fluorescent Cages of Squarate based Metal-Organic Coordination Frameworks

Keywords: superconductivity, Fe-based superconductivity, FeTe, alcohol, wine

A vanadium(iv) pyrazolate metal organic polyhedron with permanent porosity and adsorption selectivity

and their Maneuverable Application in Water Treatment

Large-Scale Synthesis of Transition-metal Doped TiO 2 Nanowires. with Controllable Overpotential

Adsorption of Methylene Blue on Mesoporous SBA 15 in Ethanol water Solution with Different Proportions

Table S1. Structural parameters of shell-by-shell fitting of the EXAFS spectrum for reduced and oxidized samples at room temperature (RT)

Supplementary information

Supplementary Information:

Transcription:

Full paper In-situ observation of nitrogen monoxide adsorption on perovskite-type MTiO 3 (M =Sr, Ba) Kenjiro FUJIMOTO, *, **,³ Yuto ISHIDUKA * and Yuki YAMAGUCHI *, ** *Department of Pure and Applied Chemistry, Faculty of Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278 8510, Japan **Research Institute for Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278 8510, Japan Strontium titanate (SrTiO 3 ; STO) and barium titanate (BaTiO 3 ; BTO) powders were prepared using the solid-state reaction method. To study the tendency for nitrogen monoxide (NO) to adsorb on STO and BTO surfaces, we observed the morphology of the powders, calculated the specific surface areas, and made adsorption/desorption observations using scanning electron microscopy (SEM), nitrogen gas adsorption/desorption isotherms under 77 K, temperature programmed desorption (TPD), and diffuse reflectance fourier transformed spectroscopy (DRIFTs). The average STO and BTO powder particle size was 0.2 m. The specific surface areas of the non-porous STO and BTO powders were 3 and 4 m 2 g ¹1, respectively. TPD and DRIFTs measurements revealed the NO adsorption properties of the STO and BTO powders in the presence of oxygen gas. Additionally, it was considered that the adsorbed NO was formed as nitrates [M(NO 3 ) 2, M = Sr, Ba] on surface of each powder. 2016 The Ceramic Society of Japan. All rights reserved. Key-words : Perovskite, Nitrogen monoxide, Adsorption, DRIFTs, TPD [Received November 13, 2015; Accepted March 28, 2016] 1. Introduction Nitrogen oxides (NO x ; NO and NO 2 ) are known air pollutants. A three-way catalyst, 1) an intelligent catalyst 2) and a NO x storage reduction catalyst 3) have been studied in the past for reducing automobile NO x gas emissions. Watanabe et al. reported that hollandite-type oxides had a capacity for NO selective catalytic reduction (NO-SCR). 4) The structure of hollandite-type compounds are tetragonal system and contain one-dimensional tunnels extending along a unique axis with a lattice period of about 0.3 nm. The hollandite structure framework consists of double chains of metal-oxygen octahedral edges that are adjacently shared. Some hollandite-type compounds include alkaline or alkaline-earth metal ions in the onedimensional tunnels that provide a charge balance. After discovering the catalytic property, Fujimoto et al. predicted that the active point of NO gas adsorption was alkaline metal ions located at the end of the one-dimensional tunnels. 5) Recently, we also demonstrated the NO adsorption for lepidocrocite-type oxides which had layered structure with including alkaline metals in an interlayer. And, it was found that NO adsorption mechanism also corresponded exactly with a hollandite-type oxide. 6) Therefore, we focused on the ability of oxide-containing alkaline-earth metals to also adsorb NO gas. In this study, we studied the presence of NO adsorption properties in strontium titanate (SrTiO 3 ; STO) and barium titanate (BaTiO 3 ; BTO) powders with perovskite-type structures. Figure 1 shows crystal structures, and Table 1 shows lattice ³ Corresponding author: K. Fujimoto; E-mail: fujimoto_kenjiro@rs. tus.ac.jp This article was selected from the presentations at STAC-9 (The 9th International Conference on the Science and Technology for Advanced Ceramics) held in Tsukuba, Japan (2015), through the regular reviewing procedure. 2016 The Ceramic Society of Japan DOI http://dx.doi.org/10.2109/jcersj2.15281 parameters of STO and BTO, respectively. 7),8) STO and BTO are known for their dielectric properties and are widely used as electronic component material for memory, sensor, and, energy devices. 9) 11) Past studies demonstrated the catalytic property of Pt-loaded BaTiO 3, but no mention has been made of direct NO x adsorption/desorption on/from BaTiO 3. 2. Experimental STO and BTO powder were prepared by the solid-state reaction method. Strontium carbonate (SrCO 3 ; High Purity Chemicals Co. Ltd., 99.9%), barium carbonate (BaCO 3 ; Wako Pure Chemical Industry Ltd. special grade) and titanium dioxide (rutile-type TiO 2 ; High Purity Chemicals Co. Ltd., 99.9%) were used as starting materials. These starting materials were mixed with a stoichiometric composition. To obtain the STO and BTO powders, the mixtures were respectively heat-treated at 1000 C and 1100 C for 5 h each. The obtained powders were identified using a powder X-ray diffractometer (XRD; Rigaku Corporation s Miniflex with Cu-K radiation) and observed with a scanning electron microscope (SEM; JSM 5500 by JEOL Ltd.). The average particle sizes of the powders were calculated from 100 randomly selected particles in the SEM images. The specific surface areas were calculated from BET plots obtained by N 2 adsorption and desorption isotherms at 77K measured by a BELSORP-max instrument (MicrotracBEL Corp.). Each powder was then pretreated in a heating process at 150 C for 5 h. Phase identification and particle property observation were followed by in-situ NO adsorption property observation with diffuse reflectance infrared fourier transform spectroscopy (DRIFTs). The DRIFTs equipment comprised an FT-IR spectrometer (Cary 660 by Agilent Technologies) and a mounted furnace Environmental Chamber (Specac Ltd.) with ZnSe windows. After placing the STO or BTO powder in the Environmental Chamber, the powder was pretreated by heating at 650 C 579

JCS-Japan Fujimoto et al.: In-situ observation of nitrogen monoxide adsorption on perovskite-type MTiO 3 (M Sr, Ba) Fig. 1. Crystal structure of perovskite-type of SrTiO 3 (STO) and BaTiO 3 (BTO). Table 1. STO and BTO crystal structure information 7) SrTiO 3 8) BaTiO 3 Crystal system Cubic Tetragonal Space group Pm-3m P4mm Cell parameters a = 0.399 nm a = 0.3905 nm c = 0.404 nm = = = 90 = = = 90 Table 2. STO and BTO lattice parameters Experimental Ref. SrTiO 3 a/nm 0.3885(3) 0.3905 7) BaTiO 3 a/nm 0.39994(8) 0.3994 8) c/nm 0.4020(2) 0.4038 8) Fig. 2. Powder X-ray diffraction patterns of STO (a) and BTO (b) heat-treated at 1000 and 1100 C. for 30 min to clean the sample surface. Single beam spectra of 650, 600, 500, 400, 300, 200, 100 C, and room temperature (RT) were measured as background data. The sample was then exposed to NO gas (c.a. 8000 ppm, Ar balanced) in the Environmental Chamber for 10 min followed by replacement of the atmosphere in the chamber with the carrier gas. DRIFT spectra of the NO-exposed powder were measured in-situ at RT, 100, 200, 300, 400, 500, 600 and 650 C by subtraction of the background data at each temperature. The carrier gases for this study were N 2 gas (99.9999%) and a mixture of N 2 and O 2 (99.5% or better) gas, adjusted to (N 2 ) 80% (O 2 ) 20%. The flow rates of the carrier gas and NO gas were controlled at 50 ml/min. The behavior and amount of NO desorption from the STO and BTO powders were examined using a temperature programmed desorption apparatus (TPD; BELCAT-B by MicrotracBEL Corp.). Approximately 100 mg of powder was placed in a quartz glass tube and heated at 650 C for 30 min. The pretreated powder was cooled to room temperature and exposed to NO gas (c.a. 8000 ppm, He balanced) at 50 ml/min for 1 h at RT. The internal glass tube was purged by He gas with the same flow rate to remove the NO gas. The NO-absorbing sample was continuously heated at 5 C/min up to 700 C with a He gas flow. Desorption gases were analyzed by a quadrupole mass spectrometer (Qmass). The signal intensity corresponding to NO and O 2 was measured respectively at 30 and 32 in m/z. 3. Results and discussions Figures 2(a) and 2(b) show the X-ray diffraction patterns of the STO and BTO powders, respectively. These patterns were identified as single-phase perovskite-type structures. Table 2 shows the lattice parameters calculated with the Appleman-Evans program. 12) Figure 3 is the morphology of the STO and BTO powders. Based on the SrTiO 3 and BaTiO 3 SEM images, the mean particle size for both powders were estimated to be 0.2 m, respectively. Figures 4 and 5 are the N 2 gas adsorption and desorption isotherms under 77K and their calculated BET plots. The STO and BTO powders exhibit Type II isotherms characteristic of nonporous powders. The STO and BTO specific surface areas were approximately 3 and 4 m 2 g ¹1, respectively. Table 3 shows the particle sizes calculated from the specific surface areas and SEM images. We used the following equation to calculate the particle size from the specific surface area: r ¼ 3=ða s µþ ð1þ where r [m] is the radius of the particle, a s [m 2 g ¹1 ] is the specific surface area, and µ [g cm ¹3 ] is the density. In this study, we could not apply the Scherrer method because the average particle size was larger than 0.1 m. Based on the morphology and calculated values, the produced particles were not agglomerates but primary particles. Figure 6 is the DRIFT spectra for STO powder with adsorbed NO gas from R.T. to 650 C. Figures 6(a) and 6(b) are the respective results for the carrier gases N 2 and (N 2 ) 80% (O 2 ) 20%. After exposure to NO gas in the Environmental Chamber, the STO powder exhibited a distinguishable difference in absorption bands and desorption temperature. The absorption band near 1200 cm ¹1 in Fig. 6(a) is considered to be for physisorption of the NO gas 580

JCS-Japan Fig. 3. Morphologies of STO and BTO powder [(a) STO and (b) BTO]. Fig. 4. N 2 gas adsorption and desorption isotherms under 77K [(a) STO and (b) BTO]. Fig. 5. BET plot of STO and BTO powder [(a) STO and (b) BTO]. Table 3. Comparison of specific surface areas (SSA), calculated particle sizes from SSA and density, and estimated sizes from SEM images Specific surface area /m 2 g ¹1 Density /g cm ¹3 Particle size (Calc.) / m Particle size (SEM images) / m SrTiO 3 2.98 5.12 0.197 0.224 BaTiO 3 3.61 6.02 0.138 0.169 by the STO powder. And, in Fig. 6(b), the absorption bands of 1300 1400 and 1800 cm ¹1 are likely from Sr(NO 3 ) 2 formed by chemisorption; the spectra are very similar to that of commercially available Sr(NO 3 ) 2 (Aldrich Chemical Co., 99.995%). The absorption bands of 1500 1700 cm ¹1 in Fig. 6(b) are also believed to be from physisorption of NO. Fig. 6. DRIFT spectra of STO powder with adsorbed NO gas from R.T. to 650 C [(a) Carrier gas; N 2, (b) (N 2 ) 80% (O 2 ) 20% ]. 581

JCS-Japan Fujimoto et al.: In-situ observation of nitrogen monoxide adsorption on perovskite-type MTiO 3 (M Sr, Ba) Figure 7 is the DRIFT spectra of BTO powder with adsorbed NO gas from R.T. to 650 C. Figures 7(a) and 7(b) are the respective results for the carrier gases N 2 and (N 2 ) 80% (O 2 ) 20%. These spectra exhibited the same tendency as the STO powder. In Fig. 7(b), the absorption bands of 1300 1450 and 1780 cm ¹1 were attributed to Ba(NO 3 ) 2 based on the spectra of commercially available Ba(NO 3 ) 2 (Aldrich Chemical Co., 99.999%). The upper left corner in Figs. 6 and 7 showed absorption band of NO gas in the Environmental Chamber. Based on these results, it can be said that the presence of oxygen in the carrier gas plays a significant role in adsorption to form NO 3 on Sr or Ba-sites. In addition, these powders exhibit NO gas adsorption capability at 650 C in Figs. 6(b) and 7(b). Figure 8 shows the TPD curves for the STO and BTO powders. Figure 9 shows the wave deconvolutions of the TPD curves. Tables 4 and 5 show the amount of desorbed NO and O 2 from the surfaces of the STO and BTO powders. The TPD curves of m/z = 30 (NO) and m/z = 32 (O 2 ) can be separated into 5 peaks and their behavior corresponds closely to the DRIFT spectra. The amount of NO and O 2 desorption between 420 to 550 C Table 4. Amount of desorption NO (m/z = 30) and O 2 (m/z = 32) from surface of STO powder Temperature range / C 1 40 100 C 0.860 2 30 370 C 8.88 m/z = 30 3 220 430 C 9.42 4 370 430 C 4.80 5 420 550 C 13.7 m/z = 32 420 550 C 12.7 Fig. 7. DRIFT spectra of BTO powder with adsorbed NO gas from R.T. to 650 C [(a) Carrier gas; N 2, (b) (N 2 ) 80% (O 2 ) 20% ]. Table 5. Amount of desorption NO (m/z = 30) and O 2 (m/z = 32) from surface of BTO powder Temperature range / C 1 100 330 C 0.970 2 160 330 C 1.83 m/z = 30 3 230 450 C 3.89 4 400 560 C 1.62 5 420 600 C 2.55 m/z = 32 420 600 C 2.54 Fig. 8. TPD curves of STO and BTO powder. Fig. 9. Wave deconvolutions of TPD curves. 582

JCS-Japan Table 6. Amount of desorption NO per unit weight and area Specific surface area a s /m 2 g ¹1 / mol m ¹2 SrTiO 3 2.98 13.7 4.60 BaTiO 3 3.61 4.17 0.866 Table 7. Estimated number of exposed Sr 2+ on STO surface and area per particle Plane direction from the STO powder were respectively calculated to be 14 and 13 mol g ¹1, which corresponds with the NO chemisorption. Similarly for BTO, the amount of NO and O 2 desorption from 400 to 600 C were respectively calculated to be 4.2 and 2.5 mol g ¹1, which corresponds with the NO chemisorption. Therefore, we believe that the active point of NO adsorption on the STO and BTO surfaces was alkaline-earth metal ions and that the presence of oxygen in the carrier gas plays a significant role in NO adsorption. However, we wanted to know why the amounts of NO desorption from STO and BTO are different. Table 6 shows the desorption amount of the chemically adsorbed NO molecule per unit weight and area. This result indicates that the amount of NO desorption from BaTiO 3 was less than that of SrTiO 3. It is considered that the difference in the amount of NO desorption is due to the difference in basicity strength. For a quantitative analysis, we assumed NO molecules being adsorbed to all Sr 2+ sites on the STO powder surface. And this adsorbed amount was treated as theoretical amount. Initially, we assumed the amount based on the exposed plane direction. We hypothesized that the following planes were exposed if one particle is a polyhedron. Table 7 shows the number of exposed Sr 2+ site on surface and area per unit cell which were calculated by the following equations: ð100þ; ð010þ; ð001þ; ð100þ; ð010þ; ð001þ; 6 a 2 ð2þ ð110þ; ð101þ; ð011þ; ð110þ; ð101þ; ð011þ; ð110þ; p ð101þ; ð011þ; ð110þ; ð101þ; ð011þ; 12 ffiffiffi 2 a 2 ð3þ ð111þ; ð111þ; ð111þ; ð111þ; ð111þ; ð111þ; ð111þ; pffiffi ð111þ; 8ð 3 a 2 =2Þ ð4þ where a is the lattice parameter. We can calculate the number of Sr 2+ per unit area from multiplication of the hypothetical area and the number of Sr 2+ on the STO surface. The theoretical amount of NO gas adsorption can be derived from Eq. (5): NO ads ¼ a s N Sr S t 26 N a Number of Sr 2+ on STO surface Area per plane direction/m 2 (100), (010), (001), (100), (010), (001) 0 6 (1.51 10 ¹19 ) 1 12 12 (2.13 10 ¹19 ) (110), (101), (011) (111), (111), (111), (111), (111), (111), (111), (111) 1 8 8 (1.31 10 ¹19 ) Total 0.77 (=N Sr ) 1.73 10 ¹19 (=S t ) ð5þ where: NO ads [ mol g ¹1 ] is the amount of NO adsorption, a s [m 2 g ¹1 ] is the specific surface area, N Sr, S t [m 2 ] is the estimated number of exposed Sr 2+ ions and the total area on one particle, and N a [mol ¹1 ] is Avogadro s number. N Sr and S t were calculated from the hypothesis that the one particle was accumulated 27 unit cells and formed by 26 direction planes as described in Eqs. (2) (4) on the particle surface. Then, N Sr was 0.77 because there was no Sr 2+ on (100), (010), (001), (100), (010) and (001) planes. From the above assumption, NO ads was calculated to be 21.9 mol g ¹1. Therefore, it was considered that 62.4% of the NO molecules desorbed from Sr 2+ and 37.6% of the NO molecules remained on STO surface. The absorption bands at 650 C were attributed to the remaining NO molecules. 4. Conclusion SrTiO 3 (STO) and BaTiO 3 (BTO) with perovskite-type structure were prepared by using the solid-state reaction method. The specific surface areas of STO and BTO were approximately 3 and 4m 2 g ¹1, respectively. In-situ DRIFTs measurements and TPD curves indicated that NO gas adsorbed on surface alkaline-earth metal ions (Sr 2+ and Ba 2+ ) as Sr(NO 3 ) 2 and Ba(NO 3 ) 2. Moreover, the presence of oxygen gas promoted formation of nitrate with Sr 2+ and Ba 2+. It was considered that these powders can be using higher temperature than usually catalyst (The operating temperature: 300 400 C) because these spectra were able to exhibit NO gas adsorption capability at 650 C. NO desorption behavior in the TPD curve correlated well with the DRIFTs measurements. We therefore believe that the active point of NO adsorption on STO and BTO was alkaline-earth metal ions. The reason for the difference in the amount of NO desorption from STO and BTO is the different basicity between Sr 2+ and Ba 2+. Finally, STO and BTO are promising newly catalysts. References 1) S. H. Oh and J. E. Carpenter, J. Catal., 98, 178 190 (1986). 2) Y. Nishihata, J. Mizuki, T. Akao, H. Tanaka, M. Uenishi, M. Kimura, T. Okamoto and N. Hamada, Nature, 418, 164 167 (2002). 3) J. Suzuki and S. Matsumoto, Top. Catal., 28, 171 176 (2004). 4) M. Watanabe, T. Mori, S. Yamauchi and H. Yamamura, Solid State Ionics, 79, 376 381 (1995). 5) K. Fujimoto, J. Suzuki, M. Harada, S. Awatsu, T. Mori and M. Watanabe, Solid State Ionics, 152 153, 769 775 (2002). 6) K. Fujimoto, R. Imaki and S. Ito, J. Jpn. Soc. Powder and Powder Metallurgy, 61[S1], S170 S172 (2014). 7) V. V. Murashov, G. M. Kuz micheva and A. V. Mintin, Russ. J. Inorg. Chem., 39, 1137 1142 (1994). 8) H. Terauchi, M. Ishida, K. Kamigaki, H. Kato and N. Sano, J. Phys. Soc. Jpn., 61, 2194 2197 (1992). 9) Y. Yano, K. Iijima, Y. Daitoh, T. Terashima, Y. Bando, Y. Watanabe, H. Kasatani and H. Terauchi, J. Appl. Phys., 76, 7833 7838 (1994). 10) B. Luo, X. Wang, E. Tian, G. Li and L. Li, J. Mater. Chem. C, 3, 8625 8633 (2015). 11) T. Furuta and K. Miura, Solid State Commun., 150, 2350 2353 (2011). 12) D. E. Appleman and H. T. Evans Jr., Job 9214; Indexing and Least-Squares Refinement of Powder Diffraction Data (Springfield, Virginia: NTIS) Doc PB 216 188 (1973). 583