首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Lead-free ceramics, SrBi2Nb2O9xBi2O3 (SBN–xBi), with different Bi contents of which the molar ratio, n(Sr) : n(Bi) : n(Nb), is 1 : 2(1 + x/2) : 2 (x = −0.05, 0.0, 0.05, 0.10), were prepared by conventional solid-state reaction method. The effect of excess bismuth on the crystal structure, microstructure and electrical properties of the ceramics were investigated. A layered perovskite structure without any detectable secondary phase and plate-like morphologies of the grains were clearly observed in all samples. The value of the activation energy suggested that the defects in samples could be related to oxygen vacancies. Excellent electrical properties (e.g., d33 = 18 pC N−1, 2Pr = 17.8 μC cm−2, ρrd = 96.4% and Tc = 420 °C) were simultaneously obtained in the ceramic where x = 0.05. Thermal annealing studies indicated the SBN–xBi ceramics system possessed stable piezoelectric properties, demonstrating that the samples could be promising candidates for high-temperature applications.

Lead-free ceramics, SrBi2Nb2O9xBi2O3 (SBN–xBi), with different Bi contents of which the molar ratio, n(Sr) : n(Bi) : n(Nb), is 1 : 2(1 + x/2) : 2 (x = −0.05, 0.0, 0.05, 0.10), were prepared by conventional solid-state reaction method.  相似文献   

2.
Highly active and thermally stable Cu–Re bimetallic catalysts supported on TiO2 with 2.0 wt% loading of Cu were prepared via an incipient wetness impregnation method and were applied for liquid phase selective hydrogenation of levulinic acid (LA) to γ-valerolactone (GVL) in H2. The effect of the molar ratios of Cu : Re on the physico-chemical properties and the catalytic performance of the Cu–Re/TiO2 catalysts was investigated. Moreover, the influence of various reaction parameters on the hydrogenation of LA to GVL was studied. The results showed that the Cu–Re/TiO2 catalyst with a 1 : 1 molar ratio of Cu to Re (Cu–Re(1 : 1)/TiO2) exhibited the highest performance for the reaction. Complete conversion of LA with a 100% yield of GVL was achieved in 1,4-dioxane solvent under the reaction conditions of 180 °C, 4.0 MPa H2 for 4 h, and the catalyst could be reused at least 6 times with only a slight loss of activity. Combined with the characterization results, the high performance of the catalyst was mainly attributed to the well-dispersed Cu–Re nanoparticles with a very fine average size (ca. 0.69 nm) and the co-presence of Cu–Re bimetal and ReOx on the catalyst surface.

Herein, we report a highly efficient and recyclable Cu–Re(1 : 1)/TiO2 bimetallic catalyst for liquid phase hydrogenation of levulinic acid to γ-valerolactone.  相似文献   

3.
Mn-Based catalysts supported on γ-Al2O3, TiO2 and MCM-41 synthesized by an impregnation method were compared to evaluate their NO catalytic oxidation performance with low ratio O3/NO at low temperature (80–200 °C). Activity tests showed that the participation of O3 remarkably promoted the NO oxidation. The catalytic oxidation performance of the three catalysts decreased in the following order: Mn/γ-Al2O3 > Mn/TiO2 > Mn/MCM-41, indicating that Mn/γ-Al2O3 exhibited the best catalytic activity. In addition, there was a clear synergistic effect between Mn/γ-Al2O3 and O3, followed by Mn/TiO2 and O3. The characterization results of XRD, EDS mapping, BET, H2-TPR, XPS and TG showed that Mn/γ-Al2O3 had good manganese dispersion, excellent redox properties, appropriate amounts of coexisting Mn3+ and Mn4+ and abundant chemically adsorbed oxygen, which ensured its good performance. In situ DRIFTS demonstrated the NO adsorption performance on the catalyst surface. As revealed by in situ DRIFTS experiments, the chemically adsorbed oxygen, mainly from the decomposition of O3, greatly promoted the NO adsorption and the formation of nitrates. The Mn-based catalysts showed stronger adsorption strength than the corresponding pure supports. Due to the abundant adsorption sites provided by pure γ-Al2O3, under the interaction of Mn and γ-Al2O3, the Mn/γ-Al2O3 catalyst exhibited the strongest NO adsorption performance among the three catalysts and produced lots of monodentate nitrates (–O–NO2) and bidentate nitrates (–O2NO), which were the vital intermediate species for NO2 formation. Moreover, the NO–TPD studies also demonstrated that Mn/γ-Al2O3 showed the best NO desorption performance among the three catalysts. The good NO adsorption and desorption characteristics of Mn/γ-Al2O3 improved its high catalytic activity. In addition, the activity test results also suggested that Mn/γ-Al2O3 exhibited good SO2 tolerance.

The Mn/γ-Al2O3 catalyst exhibited excellent performance for NO conversion in the presence of a low ratio of O3/NO, which was due to the coexistence of Mn3+ and Mn4+ and abundant chemically adsorbed oxygen.  相似文献   

4.
A series of cerium and tin oxides anchored on reduced graphene oxide (CeO2–SnOx/rGO) catalysts are synthesized using a hydrothermal method and their catalytic activities are investigated by selective catalytic reduction (SCR) of NO with NH3 in the temperature range of 120–280 °C. The results indicate that the CeO2–SnOx/rGO catalyst shows high SCR activity and high selectivity to N2 in the temperature range of 120–280 °C. The catalyst with a mass ratio of (Ce + Sn)/GO = 3.9 exhibits NO conversion of about 86% at 160 °C, above 97% NO conversion at temperatures of 200–280 °C and higher than 95% N2 selectivity at 120–280 °C. In addition, the catalyst presents a certain SO2 resistance. It is found that the highly dispersed CeO2 nanoparticles are deposited on the surface of rGO nanosheets, because of the incorporation of Sn4+ into the lattice of CeO2. The mesoporous structures of the CeO2–SnOx/rGO catalyst provides a large specific surface area and more active sites for facilitating the adsorption of reactant species, leading to high SCR activity. More importantly, the synergistic interaction between cerium and tin oxides is responsible for the excellent SCR activity, which results in a higher ratio of Ce3+/(Ce3+ + Ce4+), higher concentrations of surface chemisorbed oxygen and oxygen vacancies, more strong acid sites and stronger acid strength on the surface of the CeSn(3.9)/rGO catalyst.

A series of cerium and tin oxides anchored on graphene oxide (CeO2–SnOx/rGO) catalysts are synthesized for selective catalytic reduction of NO with NH3 in the temperature range of 120–280 °C.  相似文献   

5.
Ce modified MnOx/SAPO-34 was prepared and investigated for low-temperature selective catalytic reduction of NOx with ammonia (NH3-SCR). The 0.3Ce–Mn/SAPO-34 catalyst had nearly 95% NO conversion at 200–350 °C at a space velocity of 10 000 h−1. Microporous SAPO-34 as the support provided the catalyst with increased hydrothermal stability. XPS and H2-TPR results proved that the Mn4+ and Oα content increased after incorporation of Ce, this promoted the conversion of NO at low temperature via a ‘fast SCR’ route. NH3-TPD measurements combined oxidation experiments of NO, NH3 indicated the reduction of both the surface acidity and the amount of acid sites, which effectively decreased the NH3 oxditaion to NO or N2O at elevated temperature and promoted the catalytic selectivity for nitrogen. A redox cycle between manganese oxide and Ce was assumed for the active oxygen transfer and facilitated the catalyst durability.

Ce modified MnOx/SAPO-34 was prepared and investigated for low-temperature selective catalytic reduction of NOx with ammonia (NH3-SCR).  相似文献   

6.
Molybdenum oxide-modified ruthenium on titanium oxide (Ru–(y)MoOx/TiO2; y is the loading amount of Mo) catalysts show high activity for the hydroconversion of carboxylic acids to the corresponding alcohols (fatty alcohols) and aliphatic alkanes (biofuels) in 2-propanol/water (4.0/1.0 v/v) solvent in a batch reactor under mild reaction conditions. Among the Ru–(y)MoOx/TiO2 catalysts tested, the Ru–(0.026)MoOx/TiO2 (Mo loading amount of 0.026 mmol g−1) catalyst shows the highest yield of aliphatic n-alkanes from hydroconversion of coconut oil derived lauric acid and various aliphatic fatty acid C6–C18 precursors at 170–230 °C, 30–40 bar for 7–20 h. Over Ru–(0.026)MoOx/TiO2, as the best catalyst, the hydroconversion of lauric acid at lower reaction temperatures (130 ≥ T ≤ 150 °C) produced dodecane-1-ol and dodecyl dodecanoate as the result of further esterification of lauric acid and the corresponding alcohols. An increase in reaction temperature up to 230 °C significantly enhanced the degree of hydrodeoxygenation of lauric acid and produced n-dodecane with maximum yield (up to 80%) at 230 °C, H2 40 bar for 7 h. Notably, the reusability of the Ru–(0.026)MoOx/TiO2 catalyst is slightly limited by the aggregation of Ru nanoparticles and the collapse of the catalyst structure.

Ru–(y)MoOx/TiO2 catalysed the hydroconversion of lauric acid to allow a remarkable yield of n-dodecane (up to 80%) under mild reaction conditions.  相似文献   

7.
A MnOx@PrOx catalyst with a hollow urchin-like core–shell structure was prepared using a sacrificial templating method and was used for the low-temperature selective catalytic reduction of NO with NH3. The structural properties of the catalyst were characterized by FE-SEM, TEM, XRD, BET, XPS, H2-TPR and NH3-TPD analyses, and the performance of the low-temperature NH3-SCR was also tested. The results show that the catalyst with a molar ratio of Pr/Mn = 0.3 exhibited the highest NO conversion at nearly 99% at 120 °C and NO conversion greater than 90% over the temperature range of 100–240 °C. Also, the MnOx@PrOx catalyst presented desirable SO2 and H2O resistance in 100 ppm SO2 and 10 vol% H2O at the space velocity of 40 000 h−1 and a testing time of 3 h test at 160 °C. The excellent low-temperature catalytic activity of the catalyst could ultimately be attributed to high concentrations of Mn4+ and adsorbed oxygen species on the catalyst surface, suitable Lewis acidic surface properties, and good reducing ability. Additionally, the enhanced SO2 and H2O resistance of the catalyst was primarily ascribed to its unique core–shell structure which prevented the MnOx core from being sulfated.

A MnOx@PrOx catalyst with a hollow urchin-like core–shell structure was prepared using a sacrificial templating method and was used for the low-temperature selective catalytic reduction of NO with NH3.  相似文献   

8.
In this experiment, a TiO2–Ce0.9Zr0.1O2 support with core–shell structure was successfully prepared by a precipitation method and VOX/TiO2–Ce0.9Zr0.1O2 catalyst was prepared by an impregnation method, and the catalyst was used to catalyze the NH3-SCR of NO. Based on the results of HRTEM, XRD, BET, H2-TPR, NH3-TPD, XPS, Py-IR, it was speculated that due to the interaction between TiO2 and Ce0.9Zr0.1O2, more oxygen vacancies and Ce3+ are generated, which are beneficial to the existence of low-valence V by electron transfer between high valence state V and Ce3+and increase the acidic sites on the catalyst surface. The catalytic activity (>97%) of the VOX/TiO2–Ce0.9Zr0.1O2 catalyst is superior to the current commercial catalyst (V2O5–WO3/TiO2) and has a higher N2 selectivity (>97.5%) at 40 000 h−1 GHSV and 250–400 °C.

VOX/TiO2–Ce0.9Zr0.1O2 catalyst exhibits high activity and selectivity in a wide temperature window.  相似文献   

9.
In this work, a series of mesoporous NixMn6−xCe ternary oxides were prepared to investigate their NO catalytic oxidation ability. The sample Ni2Mn4Ce4 showed a 95% NO conversion at 210 °C (GHSV, ∼80 000 h−1). Characterization results showed the good catalytic performance of Ni2Mn4Ce4 was due to its high specific surface area, more surface oxygen and high valance manganese species, which can be ascribed to the incorporation of three elements. Based on the results of XRD, H2-TPR, O2-TPD and XPS, we confirmed the existence of Ni3+ + Mn3+ → Ni2+ + Mn4+, Ce4+ + Ni2+ → Ce3+ + Ni3+ in Ni2Mn4Ce4, and the oxidation–reduction cycles were proved to be helpful for NO oxidation. The results from an in situ DRIFTS study indicated the presence of bidentate nitrate and monodentate nitrate species on the catalyst''s surface. The nitrate species were proved to be intermediates for NO oxidation to NO2. A nitrogen circle mechanism was proposed to explain the possible route for NO oxidation. Nickel introduction was also helpful to improve the SO2 resistance of the NO oxidation reaction. The activity drop of Ni2Mn4Ce4 was 13.15% in the presence of SO2, better than Mn6Ce4 (25.29%).

In this work, a series of mesoporous NixMn6−xCe ternary oxides were prepared to investigate their NO catalytic oxidation ability.  相似文献   

10.
A series of MOx–Cr2O3–La2O3/TiO2–N (M = Cu, Fe, Ce) catalysts with nitrogen doping were prepared via the impregnation method. Comparing the low-temperature NH3-SCR activity of the catalysts, CeCrLa/Ti–N (xCeO2yCr2O3zLa2O3/TiO2–N) exhibited the best catalytic performance (NO conversion approaching 100% at 220–460 °C). The physico-chemical properties of the catalysts were characterized by XRD, BET, SEM, XPS, H2-TPR, NH3-TPD and in situ DRIFTS. From the XRD and SEM results, N doping affects the crystalline growth of anatase TiO2 and MOx (M = Cu, Fe, Ce, Cr, La) which were well dispersed over the support. Moreover, the doping of N promotes the increase of the Cr6+/Cr ratio and Ce3+/Ce ratio, and the surface chemical adsorption oxygen content, which suggested the improvement of the redox properties of the catalyst. And the surface acid content of the catalyst increased with the doping of N, which is related to CeCrLa/TiO2–N having the best catalytic activity at high temperature. Therefore, the CeCrLa/TiO2–N catalyst exhibited the best NH3-SCR performance and the redox performance of the catalysts is the main factor affecting their activity. Furthermore, in situ DRIFTS analysis indicates that Lewis-acid sites are the main adsorption sites for ammonia onto CeCrLa/TiO2–N and the catalyst mainly follows the L–H mechanism.

A series of MOx–Cr2O3–La2O3/TiO2–N (M = Cu, Fe, Ce) catalysts with nitrogen doping were prepared via the impregnation method.  相似文献   

11.
A series of manganese-based catalysts supported by 5–10 nm, 10–25 nm, 40 nm and 60 nm anatase TiO2 particles was synthesized via an impregnation method to investigate the effect of the initial support particle size on the selective catalytic reduction (SCR) of NO with NH3. All catalysts were characterized by transmission electron microscopy (TEM), N2 physisorption/desorption, X-ray diffraction (XRD), temperature programmed techniques, X-ray photoelectron spectroscopy (XPS) and in situ diffuse reflectance infrared transform spectroscopy (DRIFTS). TEM results indicated that the particle sizes of the MnOx/TiO2 catalysts were similar after the calcination process, although the initial TiO2 support particle sizes were different. However, the initial TiO2 support particle sizes were found to have a significant influence on the SCR catalytic performance. XPS and NH3-TPD results of the MnOx/TiO2 catalysts illustrated that the surface Mn4+/Mn molar ratio and acid amount could be influenced by the initial TiO2 support particle sizes. The order of surface Mn4+/Mn molar ratio and acid amount over the MnOx/TiO2 catalysts was as follows: MnOx/TiO2(10–25) > MnOx/TiO2(40) > MnOx/TiO2(60) > MnOx/TiO2(5–10), which agreed well with the order of SCR performance. In situ DRIFTS results revealed that the NH3-SCR reactions over MnOx/TiO2 at low temperature occurred via a Langmuir–Hinshelwood mechanism. More importantly, it was found that the bridge and bidentate nitrates were the main active substances for the low-temperature SCR reaction, and bridge nitrate adsorbed on Mn4+ showed superior SCR activity among all the adsorbed NOx species. The variation of the initial TiO2 support particle size over MnOx/TiO2 could change the surface Mn4+/Mn molar ratio, which could influence the adsorption of NOx species, thus bringing about the diversity of the SCR catalytic performance.

Support particle size could influence the surface Mn4+/Mn ratio of catalysts, promoting the reactivity of bridge nitrate, therefore enhancing SCR performance.  相似文献   

12.
Reaction induced PdxBiy/SiC catalysts exhibit excellent catalytic activity (92% conversion of benzyl alcohol and 98% selectivity of benzyl aldehyde) and stability (time on stream of 200 h) in the gas phase oxidation of alcohols at a low temperature of 240 °C due to the formation of Pd0–Bi2O3 species. TEM indicates that the agglomeration of the 5.8 nm nanoparticles is inhibited under the reaction conditions. The transformation from inactive PdO–Bi2O3 to active Pd0–Bi2O3 under the reaction conditions is confirmed elaborately by XRD and XPS.

Reaction induced PdxBiy/SiC catalysts exhibit excellent catalytic activity and stability in the gas phase oxidation of monopolistic alcohols at a low temperature of 240 °C due to the formation of Pd0–Bi2O3 species.  相似文献   

13.
A series of CeO2–WO3/SiO2–TiO2 (CeWxTiSiy) catalysts with different loading amounts of WO3 were synthesized by wet co-impregnation of ammonium metatungstate and cerium nitrate on a SiO2–TiO2 support, and were employed for the selective catalytic reduction (SCR) of NO by NH3. The catalytic activity of the CeO2/SiO2–TiO2 (CeSiTi) catalyst was enhanced by the addition of WO3, and the W-containing catalysts showed higher hydrothermal stability especially between 550 and 600 °C. The introduction of WO3 to the CeSiTi catalyst could produce more chemisorbed oxygen species, reducible subsurface oxygen species, acid sites and ad-NOx species. Moreover, the modification of CeO2–WO3/TiO2 (CeWTi) by SiO2 could enhance the specific surface area, especially the aged specific surface area, thus improving the hydrothermal stability of the catalyst.

A series of CeO2–WO3/TiO2–SiO2 catalysts were prepared for SCR of NOx with NH3, and the effect of WO3 and SiO2 doping on the activity and stability of the catalysts were discussed.  相似文献   

14.
Modulating the active sites for controllable tuning of the catalytic activity has been the goal of much research, however, this remains challenging. The O vacancy is well known as an active site in reducible oxides. To modify the activity of O vacancies in praseodymia, we synthesized a series of praseodymia–titania mixed oxides. Varying the Pr : Ti mole ratio (2 : 1, 1 : 2, 1 : 1, 1 : 4) allows us to control the electronic interactions between Au, Pr and Ti cations and the local chemical environment of the O vacancies. These effects have been studied study by X-ray photoelectron spectroscopy (XPS), CO diffuse reflectance Fourier transform infrared spectroscopy (CO-DRIFTS) and temperature-programmed reduction (CO-TPR, H2-TPR). The water gas shift reaction (WGSR) was used as a benchmark reaction to test the catalytic performance of different praseodymia–titania supported Au. Among them, Au/Pr1Ti2Ox was identified to exhibit the highest activity, with a CO conversion of 75% at 300 °C, which is about 3.7 times that of Au/TiO2 and Au/PrOx. The Au/Pr1Ti2Ox also exhibited excellent stability, with the conversion after 40 h time-on-stream at 300 °C still being 67%. An optimal ratio of Pr content (Pr : Ti 1 : 2) is necessary for improving the surface oxygen mobility and oxygen exchange capability, a higher Pr content leads to more O vacancies, however with lower activity. This study presents a new route for modulating the active defect sites in mixed oxides which could also be extended to other heterogeneous catalysis systems.

Schematic illustration of H2O activation on the Pr-TiOx support and the following reaction with CO in the Au–oxide interface.  相似文献   

15.
Goethite (α-FeOOH) possesses excellent catalytic activity, high selectivity and good stability as a catalyst for NO oxidation through the catalytic decomposition of gaseous H2O2. as the primary reactive oxygen species is involved in the NO oxidation process together with ·OH, and N2O5 is found for the first time in the products of NO oxidation.

A catalyst α-FeOOH possesses excellent catalytic activity for NO oxidation, and N2O5 is first found in NO oxidation products.

Sulphur dioxide (SO2) and nitrogen oxides (NOx) yielded by fossil fuels and ore combustion are the major air pollutants produced by traditional chemical industries, including iron and steel, coking, and boilers, and these pollutants cause acid rain, fog, and haze. Wet flue gas desulphurisation (WFGD) and selective catalytic reduction (SCR) are efficient methods used in power plants for desulphurisation and denitration, respectively.1 However, in traditional chemical industries, for example, coke oven flue gas yielded by coke oven gas combustion has a flue gas temperature of about 200–230 °C and the composition is complex,2 which limits the utilisation of SCR.3Recently, we proposed a promising process of gas-phase oxidation combined with wet scrubbing using steel slag slurry to treat NOx from low-temperature flue gas.4,5 In this process, the sparingly soluble NO could be oxidised into soluble NO2, HNO3 or N2O5 through the gas-phase oxidation process, and then the oxidation products could be absorbed together with SO2 in a wet scrubbing device. This method can achieve the simultaneous removal of SO2 and NOx using the oxidisers. Some oxidisers, such as O3,6,7 H2O2,8–12 NaClO2,13 NaClO,14 persulphate salt (S2O82−)15 and ferrate (Fe[vi]),16 can be used as the gas-phase oxidiser for NO oxidation after being gasified if needed.H2O2 is a green and low-cost oxidizer, which can be used as the gas-phase oxidizer for NO oxidation after liquid H2O2 is gasified at low temperature. Furthermore, the oxidation potential of H2O2 (1.77 eV) is lower compared with that of O3 (2.08 eV) and ·OH (2.80 eV) and thus its efficiency in oxidising NO is low.12,17 Thermal decomposition of gaseous H2O2 can decompose H2O2 into radicals (·OH or ) with high oxidation potential for NO oxidation. However, this technology requires high H2O2 consumption (H2O2/NO = 80) and excessive residence time (34 s) and results in low NO oxidation efficiency (∼60%).18 Introducing catalysts into the H2O2 decomposition process can effectively decompose H2O2 into radicals and considerably reduce the consumption of H2O2.11 Iron-based materials, such as hematite (α-Fe2O3),8 nanoscale zero-valent iron,9 Fe3O4,10 γ-Fe2O3@Fe3O4,11 Fe2(MoO4)319 and Fe2(SO4)3,12 have been used as catalysts to decompose gaseous H2O2 for NO oxidation. These catalysts have high removal efficiencies as heterogeneous catalysts for the simultaneous removal of NO and SO2 in an integrated catalytic oxidation/wet scrubbing process. However, the use of these catalysts results in relatively high H2O2 consumption8,9,12,20 and relatively low gas hourly space velocity (GHSV)19 and catalytic stability.9,12 Therefore, a catalyst with low H2O2 consumption and high GHSV and catalytic stability for NO oxidation through catalytic decomposition of gaseous H2O2 should be developed.Goethite (α-FeOOH) is a ubiquitous natural mineral in soils and sediments at the Earth''s surface that is widely used as a heterogeneous Fenton catalyst for wastewater treatment due to its abundance, availability, relative stability and low cost.21 He et al. explored the catalytic performance of α-FeOOH and found that it can be used to catalyse H2O2 vapour for NO oxidation under low-temperature (<160 °C) flue gas;22 however, coke oven flue gas has a higher flue gas temperature (200–230 °C), and the reaction products and catalytic mechanism may be different under different temperature regions. Therefore, the performance of α-FeOOH for NO oxidation through catalysing gaseous H2O2 under high flue gas temperature should be investigated, and the SO2 oxidation efficiency of this process should also be evaluated. Herein, NO and SO2 conversions and NO2 yield using α-FeOOH with gaseous H2O2 were performed under the wide temperature range of 100–350 °C, and the catalytic stability and reaction mechanism were determined.Fresh α-FeOOH catalyst was prepared via precipitation reaction of Fe3+ followed by crystal transformation according to the method of Böhm.23 All chemicals used in the experiments were of analytical grade, and deionised water was used. NO, NO2, and SO2 concentrations in simulated flue gas were analysed using a UV differential optical absorption spectroscopy (DOAS) flue gas comprehensive analyser (Laoying3023, Qingdao Laoying Environmental Technology Co., Ltd.). The presence of HNO3 and N2O5 in the flue gas were detected using a Fourier transform infrared (FTIR) spectrometer (Tensor 27, Bruker Optik, Inc.), which was equipped with a 2.4 m gas cell. Catalysts were characterised via X-ray diffraction (XRD) spectroscopy (Empyrean, PANalytical Instruments) and FTIR spectrometry. Hydroxyl radical (·OH) and superoxide anion radical decomposed by H2O2 were detected via electron paramagnetic resonance (EPR) spectroscopy (A300-10/12, Bruker Optik Inc.), and 5,5-dimethyl-1-pyrroline N-oxide (DMPO) was used for capturing ·OH and . Specifically, DMPO–H2O solvent was used to capture the ·OH generated after H2O2 was catalysed, whereas DMPO–CH3OH was used to capture The experimental equipment for evaluating the catalytic performance of α-FeOOH was established (Fig. S1, ESI). In brief, the bypass of N2 carried the gaseous H2O2 generated by the evaporation of H2O2 solution and mixed with the simulated flue gas, and then the mixture gas contacted with the catalyst, in which gaseous H2O2 was decomposed into radicals over the catalyst and oxidised NO and SO2. At the outlet of the catalytic reactor, the simulated flue gas after being oxidised was detected using a UV DOAS flue gas comprehensive analyser and an FTIR spectrometer equipped with a gas cell. Each experiment was conducted for 20 min after the temperature was stabilised.According to Christensen et al., α-FeOOH is transformed to hematite (α-Fe2O3) within the temperature range of 171–311 °C, and α-FeOOH is totally converted to α-Fe2O3 at 350 °C.24 The fresh catalyst and the catalyst calcinated at 350 °C were characterised via XRD and FTIR spectrometry. The FTIR and XRD spectra (Fig. S2 and S3, ESI) indicated that the fresh catalyst was α-FeOOH and the catalyst calcinated at 350 °C was α-Fe2O3 in an N2 atmosphere. The colour of the fresh catalyst (yellow) and the calcinated catalyst (red) also proved the above results (Fig. S4, ESI). An FTIR spectrometer equipped with a gas cell was used to analyse the water vapour and further investigate the transformation temperature of α-FeOOH. Results (Fig. S5, ESI) showed that α-FeOOH began to decompose into α-Fe2O3 and H2O when the temperature was above 200 °C in an N2 atmosphere. However, when H2O(g) and H2O2(g) were existed in N2 atmosphere, they could improve the thermal stability of α-FeOOH. The reason may be that (1) multiple types of surface hydroxyls (–OHs) generated by the adsorption of water on α-FeOOH surface prevented the dehydration of –OHs;25,26 and (2) H2O(g) from the injected H2O2 solution were adsorbed on the reduced Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) and generated Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH viaeqn (1) and (2), and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH was converted to Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH viaeqn (3).12,19,22,27,28 The catalyst α-FeOOH was still stable when temperature was up to 225 °C, and little α-FeOOH close to the wall of the reactor was converted to α-Fe2O3 (the color of catalyst was changed from yellow to red) when temperature was up to 350 °C. Therefore, the catalyst α-FeOOH possesses great thermal stability under the simulated flue gas condition (e.g., the experimental condition in Fig. 1).1 Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) + Vo + H2O → Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH2+ Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH + H+2 Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH + H2O2 Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH + ·OH3Open in a separate windowFig. 1Effects of temperature on NO and SO2 conversions and NO2 yield (H2O2/NO, 2.0; H2O2 solution feeding rate, 148.9 μL min−1; catalyst dosage, 0.5 g; GHSV, 137 747 h−1; NO concentration, 200 ppm; SO2 concentration, 660 ppm; O2 concentration, 6%).As shown in Fig. 1, NO conversion and NO2 yield were both lower than SO2 conversion when gaseous H2O2 only was used for NO oxidation. This result indicated that SO2 was more easily oxidised by gaseous H2O2 than NO, which consumed a large amount of H2O2. NO conversion and NO2 yield were higher, but SO2 conversion was considerably lower than using gaseous H2O2 only when α-FeOOH was used to catalyse gaseous H2O2 for NO oxidation. Therefore, α-FeOOH performed excellent catalytic activity and high selectivity for NO oxidation. Specifically, NO conversion achieved 98.8% and NO2 yield reached 77.0% at H2O2/NO of 2.0, reaction temperature of 225 °C, catalyst dosage of 0.5 g and GHSV of 137 747 h−1. The catalyst of α-FeOOH achieved a higher NO oxidation efficiency under low H2O2 consumption and high GHSV compared with those reported by previous studies (Table S1, ESI).8,9,12,17,19,20 Fig. 1 shows that NO conversion was higher than NO2 yield within the temperature range of 100–225 °C, and NO conversion was closer to NO2 yield when the temperature further increased from 250 °C to 350 °C. This trend indicated that the products of NO oxidation was not just NO2 within the temperature range of 100–225 °C, and the product of NO oxidation might only be NO2 within the temperature range of 250–350 °C. The products of NO oxidation were determined via FTIR spectroscopy. As shown in Fig. 2, the peaks for NO2 (1599 cm−1), HNO3 (886 cm−1) and N2O5 (1719 cm−1) were observed in the FTIR spectra within the temperature range of 100–200 °C.29,30 However, no obvious peaks for NO2, HNO3, and N2O5 were detected in the FTIR spectra when H2O2 was not added. This result indicated that NO2, HNO3 and N2O5 were the main products of NO oxidation through the catalytic decomposition of gaseous H2O2 over α-FeOOH in low-temperature (100–200 °C) region. Furthermore, the peaks for HNO3 and N2O5 in the FTIR spectra disappeared when the temperature increased from 225 °C to 350 °C. The reason is that HNO3 and N2O5 decomposed in the high-temperature region (225–350 °C) (eqn (4) and (5)).29 In summary, NO2, HNO3 and N2O5 were the main products of NO oxidation in the low-temperature region (100–200 °C), whereas NO2 was the main product in the high-temperature region (225–350 °C).N2O5 → NO + NO2 + O242HNO3 → NO + NO2 + O2 + H2O5Open in a separate windowFig. 2FTIR spectra of simulated flue gas oxidised by gaseous H2O2 over α-FeOOH under different temperature conditions. (Experimental condition was the same as that in Fig. 1.) Fig. 1 shows that NO conversion sharply increased when the temperature increased from 100 °C to 200 °C and then decreased when the temperature further increased from 225 °C to 350 °C. The increase in NO conversion with temperature was because high temperatures enhance H2O2 decomposition into radicals according to the Arrhenius law,12 which further accelerates NO conversion. The decrease in NO conversion at temperatures above 200 °C could be explained by (1) the thermal decomposition of gaseous H2O2 under high temperature and/or (2) the transformation of α-FeOOH under high temperature. On the one hand, gaseous H2O2 can be decomposed under high temperature. An FTIR spectrometer equipped with a gas cell was used to detect gaseous H2O2 under different temperature conditions (free of catalyst) to investigate the decomposition temperature of gaseous H2O2,31 and the variations in H2O2 content were measured by their homologous infrared absorption characteristic peaks (Fig. S6, ESI). Results (Fig. S7, ESI) indicated that H2O2 almost did not decompose at 100–300 °C, and the thermal decomposition of gaseous H2O2 occurred when the temperature increased further to 300 °C. On the other hand, NO-TPD (Fig. S8, ESI) showed that the absorbed NO began to desorb when the temperature was above 200 °C. Therefore, when the temperature was above 200 °C, NO conversion decreased when the temperature was above 200 °C. Overall, the transformation of α-FeOOH under high temperature (>200 °C) led to the decrease in NO conversion when the temperature was above 200 °C, and the thermal decomposition of gaseous H2O2 also resulted in the decrease in NO conversion when the temperature was above 300 °C.The catalytic stability of α-FeOOH was also investigated. As shown in Fig. 3, NO conversion was maintained at a high level (>97.0%) within the first 15 h and then fluctuated slightly but remained at >90.0% at 15–45 h. SO2 conversion remained stable at about 2.0%. Results indicated that α-FeOOH possesses excellent catalytic activity, good stability and high selectivity for NO oxidation. According to the results in Fig. 1 and and3,3, the catalyst has an ideal temperature window of 175–250 °C; therefore, α-FeOOH can be applied in coke oven flue gas (200–230 °C) treatment.Open in a separate windowFig. 3Catalytic stability of α-FeOOH. (H2O2/NO, 2.0; H2O2 solution feeding rate, 148.9 μL min−1; catalyst dosage, 0.5 g; GHSV, 137 747 h−1; temperature, 225 °C; NO concentration, 200 ppm; SO2 concentration, 660 ppm; O2 concentration, 6%.)Fresh and used (after the 45 h test) α-FeOOH were characterised via FTIR and XRD spectroscopy. The bands at 3363, 3127 and 1628 cm−1 are the O–H stretching mode in α-FeOOH, the stretching mode of surface water and the bending mode of H2O, respectively. The characteristic strong bands at 795 and 891 cm−1 were assigned to the Fe–O–H bending vibrations of α-FeOOH. The band at 638 cm−1 was assigned to the Fe–O stretching vibration of pure α-FeOOH. The characteristic absorption peaks of α-FeOOH in the catalyst after the 45 h test did not change compared with that of the fresh catalyst (Fig. S9, ESI), indicating that the catalyst maintained the structure of α-FeOOH even after the 45 h test. The band at 999 cm−1 was assigned to ν1(SO4) frequency, and the bands at 1076, 1136 and 1229 cm−1 were interpreted as ν3(SO4) frequencies. These vibrational frequencies are attributed to specifically adsorbed SO42− ions on the external and internal surfaces of catalyst particles after the 45 h test.32,33 The XRD patterns showed that the phase of the catalyst before and after the stability test was still α-FeOOH (Fig. S10, ESI). This result also indicated that α-FeOOH possessed good stability in the NO oxidation process.EPR test was conducted to detect the radicals generated by H2O2 decomposition and analyse the oxygen species of the α-FeOOH/H2O2 system. Fig. 4 shows the 4-fold characteristic peak of DMPO–·OH adducts with an intensity ratio of 1 : 2 : 2 : 1 and four notable signals ascribed to adducts. These peaks and signals indicate the formation of ·OH and in the α-FeOOH/H2O2 system.21,34 Furthermore, the intensity of the 4-fold characteristic peak of DMPO–·OH adducts was significantly weaker than that of adducts, indicated that the quantity of produced by the α-FeOOH/H2O2 system was larger than that of ·OH. Therefore, α-FeOOH can decompose H2O2 into ·OH and and may be dominant among ·OH and for NO oxidation. Radical trapping experiments (Fig. S11, ESI) further proved the roles of radicals (·OH and ) in the NO oxidation process. Benzoquinone (BQ, scavenger20) was added into the H2O2 solutions, and H2O2 and BQ in the mixture solution was vapoured together. NO conversion substantially decreased from 92.9% to 6.1% when the molar ratio of BQ to H2O2 increased from 0 to 1.0. The reason was that the addition of BQ captured the generated by gaseous H2O2 decomposition over α-FeOOH, thereby decreasing NO conversion. NO conversion decreased from 92.9% to 8.0% when isopropanol (i-PrOH, ·OH scavenger20) was introduced into the same system as the molar ratio of i-PrOH to H2O2 increased from 0 to 8.0. The reason was that ·OH generated in the α-FeOOH/H2O2 system was captured by i-PrOH instead of oxidised NO. These results indicated that when the addition concentration of the ·OH scavenger (i-PrOH) was eight times that of scavenger (BQ), both systems obtained the similar decrease in NO conversion. Therefore, as the primary reactive oxygen species was involved in the NO oxidation process together with ·OH.Open in a separate windowFig. 4EPR spectra of the α-FeOOH/H2O2 system.In this catalytic process, the reaction between H2O2 as an oxidant and iron ions as a catalyst to produce highly active species (·OH and ).35 The conversion between Fe2+ and Fe3+ was proceed according to the Haber–Weiss mechanism. Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii) are reduced by H2O2 and generate Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) (eqn (6) and (1)), the resulting Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) also can be oxidized by H2O2 and produce Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii) (eqn (3) and (7)).12,19,22,27,28,366 Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) + H2O2 Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii) + ·OH + OH7The proposed mechanism of NO oxidation by H2O2 decomposition over α-FeOOH is presented in Fig. 5 based on the study of the products of NO oxidation and reactive oxygen species ( and ·OH) in the NO oxidation process. (1) Gaseous H2O2 decomposed into ·OH and on the active sites ( Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii), Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH and Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)) of the catalyst (eqn (1)–(3), (6) and (7)) according to the Haber–Weiss mechanism, a result that was proved through EPR analysis and scavenger experiments. (2) NO was oxidised by the generated ·OH and and produced NO2 and HNO3viaeqn (8) and (9). (3) The produced HNO3 reacted with ·OH and produced NO3 or NO2 (eqn (10) and (11)). (4) NO2 reacted with NO3 and produced N2O5viaeqn (12).37 The production of NO2, HNO3 and N2O5 during the NO oxidation process was proved by the FTIR spectra (Fig. 2).NO + 2·OH → NO2 + H2O89HNO3 + ·OH → NO3 + H2O10HNO3 + ·OH → NO2 + H2O211NO2 + NO3 → N2O512Open in a separate windowFig. 5Catalytic mechanism of NO oxidation by H2O2 decomposition over α-FeOOH ( Created by potrace 1.16, written by Peter Selinger 2001-2019 FeIII: Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii), Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(iii)–OH; Created by potrace 1.16, written by Peter Selinger 2001-2019 FeII: Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii), Created by potrace 1.16, written by Peter Selinger 2001-2019 Fe(ii)–OH).In summary, α-FeOOH can be used as an efficient catalyst for coke oven flue gas to enhance NO oxidation efficiency through the catalytic decomposition of gaseous H2O2. Moreover, α-FeOOH showed high selectivity for NO oxidation in the presence of SO2. In this study, NO conversion achieved 98.8% under the following experimental conditions: H2O2/NO of 2.0, reaction temperature of 225 °C, catalyst dosage of 0.5 g and GHSV of 137 747 h−1. The 45 h test indicated that α-FeOOH has good catalytic stability. The EPR test and radical trapping experiments revealed that as the primary reactive oxygen species was involved in the NO oxidation process together with ·OH. Furthermore, NO2, HNO3 and N2O5 were the products of NO oxidation through the catalysis of gaseous H2O2 over α-FeOOH within the temperature range of 100–200 °C, and NO2 was the only oxidation product within the temperature range of 225–350 °C.The catalyst α-FeOOH can be used to decompose gaseous H2O2 into radicals and oxidise NO into high-valence NOx, and then the oxidation products can be absorbed together with SO2 in existing industrial-scale WFGD systems.4,5 This process can achieve the simultaneous removal of SO2 and NOx from coke oven flue gas.  相似文献   

16.
CuO–CeO2 nanocatalysts with different amounts of Mn dopping (Mn/Cu molar ratios of 0.5 : 5, 1 : 5 and 1.5 : 5) were synthesized by flame spray pyrolysis (FSP) method and tested in the catalytic oxidation of CO. The physicochemical properties of the synthesised samples were characterized systematically, including using X-ray diffraction (XRD), Raman spectroscopy, field-emission scanning electron microscopy (FESEM), Brunauer–Emmett–Teller (BET), X-ray photoelectron spectroscopy (XPS), oxygen-temperature programmed desorption (O2-TPD), hydrogen-temperature programmed reduction (H2-TPR) and in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS). The results showed that the 1Mn–Cu–Ce sample (Mn/Cu molar ratio of 1 : 5) exhibited superior catalytic activity for CO oxidation, with the temperature of 90% CO oxidation at 131 °C at a high space velocity (SV = 60 000 mL g−1 h−1), which was 56 °C lower than that of the Cu–Ce sample. In addition, the 1Mn–Cu–Ce sample displays excellent stability with prolonged time on CO stream and the resistance to water vapor. The significantly enhanced activity was correlated with strong synergetic effect, leading to fine textual properties, abundant chemically adsorbed oxygen and high lattice oxygen mobility, which further induced more Cu+ species and less formation of carbon intermediates during the CO oxidation process detected by in situ DRIFTS analysis. This work will provide in-depth understanding of the synergetic effect on CO oxidation performances over Mn doped CuO–CeO2 composite catalysts through FSP method.

The synergetic effect is promoted on Mn doped CuO–Ce O2 catalyst to induce less carbon intermediates to enhance CO oxidation performance.  相似文献   

17.
We investigated the hydride reduction of tetragonal BaTiO3 using LiH. The reactions employed molar H : BaTiO3 ratios of 1.2, 3, and 10 and variable temperatures up to 700 °C. The air-stable reduced products were characterized by powder X-ray diffraction (PXRD), scanning electron microscopy, thermogravimetric analysis (TGA), X-ray fluorescence (XRF), and 1H magic-angle spinning (MAS) NMR spectroscopy. Effective reduction, as indicated by the formation of dark blue to black colored, cubic-phased, products was observed at temperatures as low as 300 °C. The product obtained at 300 °C corresponded to oxyhydride BaTiO∼2.9H∼0.1, whereas reduction at higher temperatures resulted in simultaneous O defect formation, BaTiO2.9−xH0.1x, and eventually – at temperatures above 450 °C – to samples void of hydridic H. Concomitantly, the particles of samples reduced at high temperatures (500–600 °C) display substantial surface alteration, which is interpreted as the formation of a TiOx(OH)y shell, and sintering. Diffuse reflectance UV-VIS spectroscopy shows broad absorption in the VIS-NIR region, which is indicative of the presence of n-type free charge carriers. The size of the intrinsic band gap (∼3.2 eV) appears only slightly altered. Mott–Schottky measurements confirm the n-type conductivity and reveal shifts of the conduction band edge in the LiH reduced samples. Thus LiH appears as a versatile reagent to produce various distinct forms of reduced BaTiO3 with tailored electronic properties.

Different forms of reduced BaTiO3, which include oxyhydride BaTiO2.9H0.1 and O-deficient BaTiO2.9−xH0.1x, were obtained from reactions with LiH at various temperatures.  相似文献   

18.
A series of W–Zr-Ox/TiO2 catalysts with hierarchical pore structure were prepared and used for selective catalytic reduction of NO by NH3. Results showed that the 5C-WZ/T had a hierarchical pore structure, and exhibited high catalytic activity and good resistance to water and sulfur poisoning. The activity of the 5C-WZ/T catalyst was close to 100% in the range of 350–500 °C. The hierarchical pore structure not only improved the specific surface area, redox performance, and acid quantity, but also enabled the catalyst to expose more active sites and improved the catalytic performance. Although the reducing atmosphere caused by citric acid monohydrate reduced the chemical adsorption oxygen concentration, it increased the oxygen mobility and Ti3+ ion concentration, which was more conducive to the improvement of catalytic activity. Finally, the NH3-SCR reaction over 5C-WZ/T catalyst followed L-H and E-R mechanisms. The monodentate nitrite, bidentate nitrate, gas-phase NO2, coordinated NH3 and NH4+ were the main reaction intermediates.

A series of W–Zr-Ox/TiO2 catalysts with hierarchical pore structure were prepared and used for selective catalytic reduction of NO by NH3.  相似文献   

19.
NASICON-structured Na3V2O2x(PO4)2F3−2x (0 < x ≤ 1) solid solutions have been prepared using a microwave-assisted hydrothermal (MW-HT) technique. Well-crystallized phases were obtained for x = 1 and 0.4 by reacting V2O5, NH4H2PO4, and NaF precursors at temperatures as low as 180–200 °C for less than 15 min. Various available and inexpensive reducing agents were used to control the vanadium oxidation state and final product morphology. The vanadium oxidation state and O/F ratios were assessed using electron energy loss spectroscopy and infrared spectroscopy. According to electron diffraction and powder X-ray diffraction, the Na3V2O2x(PO4)2F3−2x solid solutions crystallized in a metastable disordered I4/mmm structure (a = 6.38643(4) Å, c = 10.62375(8) Å for Na3V2O2(PO4)2F and a = 6.39455(5) Å, c = 10.6988(2) Å for Na3V2O0.8(PO4)2F2.2). With respect to electrochemical Na+ (de)insertion as positive electrodes (cathodes) for Na-ion batteries, the as-synthesized materials displayed two sloping plateaus upon charge and discharge, centered near 3.5–3.6 V and 4.0–4.1 V vs. Na+/Na, respectively, with a reversible capacity of ∼110 mA h g−1. The application of a conducting carbon coating through the surface polymerization of dopamine with subsequent annealing at 500 °C improved both the rate capability (∼55 mA h g−1 at a discharge rate of 10C) and capacity retention (∼93% after 50 cycles at a discharge rate of C/2).

NASICON-structured Na3V2O2x(PO4)2F3−2x (0 < x ≤ 1) solid solutions have been prepared using a microwave-assisted hydrothermal (MW-HT) technique.  相似文献   

20.
Submicron-sized niobia (Nb2O5) porous spheres with a high specific surface area (300 m2 g−1) and nano concave–convex surfaces were synthesized via a rapid one-pot single-step alcothermal reaction. Prolonged reaction time or high reaction temperatures resulted in a morphology change of Nb2O5 from amorphous sphere to rod crystals with hexagonal crystal phase. A similar alcothermal reaction yielded TiO2–Nb2O5 composite porous spheres, whose Ti : Nb molar ratio was controlled by changing the precursor solution component ratios. A simple thermal treatment of amorphous TiO2–Nb2O5 porous spheres consisting of 1 : 2 (molar ratio) Ti : Nb at 600 °C for 2 h induced crystal phase transfer from amorphous to a monoclinic crystal phase of submicron-sized TiNb2O7 porous spheres with a specific surface area of 50 m2 g−1.

Nb2O5, TiO2–Nb2O5, and TiNb2O7 porous spheres with large surface area were synthesized by alcothermal reaction and calcination.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号