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 15 毫秒
1.
When the perovskites are calcined at 750 °C, the incorporation of Pd into LaMnO3 enhances the activity of the catalyst in methane combustion at temperatures below 750 °C upon substitution of 0.1 mol La with Pd, and at temperatures below 600 °C when Pd is substituted for 0.1–0.15 mol Mn. Monolith catalysts based on La1−xPdxMnO3 (x = 0.1, 0.15) display a higher activity in methane combustion than do LaMn1−xPdxO3-based catalysts, which is due to the higher Pd/(Pd + Mn + La) ratio. The activities of the two perovskite types increase when calcination temperature is raised from 650 to 800 °C. With the increase in calcination temperature, an increase in the Pd content and a decrease in the La content is observed on the surfaces (X-ray photoelectron spectroscopy (XPS)). The rise in the temperature of perovskite calcination to 850 °C produces sintering which leads to the lowering in both the Pd content on the surfaces and the specific surface areas (SSAs) of the perovskites and, consequently, decreases catalytic activity.  相似文献   

2.
Two series of Sr- or Ce-doped La1−xMxCrO3 (x = 0.0, 0.1, 0.2 and 0.3) catalysts were prepared by thermal decomposition of amorphous citrate precursors followed by annealing at 800 °C in air atmosphere. The effect of Ce and Sr on the morphological/structural properties of LaCrO3 was investigated by means of thermogravimetric/differential thermal analysis (TG/DTA) of the precursors decomposition under air, X-ray diffraction (XRD), electron paramagnetic resonance (EPR), transmission electron microscopy–X-ray energy dispersive spectroscopy (TEM–XEDS), SBET determination, scanning electron microscopy (SEM) and X-ray photoelectron spectroscopy (XPS) techniques. The characterization results are employed to explain catalytic activity results for C3H6 combustion. It is shown that the lanthanum chromite perovskite structure is obtained upon thermal treatment of the sol–gel derived precursors at T > ca. 800 °C. The presence of the dopant generally induces the formation of segregated oxide phases in the samples calcined at 800 °C although some introduction of the Sr in the perovskite structure is inferred from EPR measurements. The oxidation activity becomes maximised upon formation of such doped perovskite structure.  相似文献   

3.
A number of nano-gold catalysts were prepared by depositing gold on different metal oxides (viz. Fe2O3, Al2O3, Co3O4, MnO2, CeO2, MgO, Ga2O3 and TiO2), using the homogeneous deposition precipitation (HDP) technique. The catalysts were evaluated for their performance in the combustion of methane (1 mol% in air) at different temperatures (300–600 °C) for a GHSV of 51,000 h−1. The supported nano-gold catalysts have been characterized for their gold loading (by ICP) and gold particle size (by TEM/HRTEM or XRD peak broadening). Among these nano-gold catalysts, the Au/Fe2O3 (Au loading = 6.1% and Au particle size = 8.5 nm) showed excellent performance. For this catalyst, temperature required for half the methane combustion was 387 °C, which is lower than that required for Pd(1%)/Al2O3 (400 °C) and Pt(1%)/Al2O3 (500 °C) under identical conditions. A detailed investigation on the influence of space velocity (GHSV = 10,000–100,000 cm3 g−1 h−1) at different temperatures (200–600 °C) on the oxidative destruction of methane over the Au/Fe2O3 catalyst has also been carried out. The Au/Fe2O3 catalyst prepared by the HDP method showed much higher methane combustion activity than that prepared by the conventional deposition precipitation (DP) method. The XPS analysis showed the presence of Au in the different oxidation states (Au0, Au1+ and Au3+) in the catalyst.  相似文献   

4.
In situ FT-IR was employed to investigate CO or/and NO interaction with CuO supported on Ce0.67Zr0.33O2 (hereafter denoted as CZ) catalysts. The physicochemical properties of CuO–CZ were also studied by combination of XRD, TPR and NO + CO activity tests. The results indicated that the dispersed CuO species were the main active components for this reaction. The catalysts showed different activities and selectivities at low and high temperatures, which should be resulted from the reduction of dispersed copper oxide species. This reaction went through different mechanisms at low and high temperatures due to the change of active species. FT-IR results suggested: (1) CO was activated by oxygen originating from CZ support, which led to surface carbonates formation, and partial dispersed CuO was reduced to Cu+ species above 150 °C; (2) NO interacted with the dispersed CuO and formed several types of nitrite/nitrate species, whereas crystalline CuO made little contribution to the formation of new NO adsorbates; (3) NO was preferentially adsorbed on CuO–CZ catalysts compared with CO in the reactants mixture. These adsorbed nitrite/nitrate species exhibited different thermal stability and reacted with CO at 250 °C. As a result, a possible mechanism was tentatively proposed to approach NO reduction by CO over CuO–CZ catalyst.  相似文献   

5.
Oxygen storage capacity (OSC) of CeO2–ZrO2 solid solution, CexZr(1−x)O4, is one of the most contributing factors to control the performance of an automotive catalyst. To improve the OSC, heat treatments were employed on a nanoscaled composite of Al2O3 and CeZrO4 (ACZ). Reductive treatments from 700 to 1000 °C significantly improved the complete oxygen storage capacity (OSC-c) of ACZ. In particular, the OSC-c measured at 300 °C reached the theoretical maximum with a sufficient specific surface area (SSA) (35 m2/g) after reductive treatment at 1000 °C. The introduced Al2O3 facilitated the regular rearrangement of Ce and Zr ions in CeZrO4 as well as helped in maintaining the sufficient SSA. Reductive treatments also enhanced the oxygen release rate (OSC-r); however, the OSC-r variation against the evaluation temperature and the reduction temperature differed from that of OSC-c. OSC-r measured below 200 °C reached its maximum against the reduction temperature at 800 °C, while those evaluated at 300 °C increased with the reduction temperature in the same manner as OSC-c.  相似文献   

6.
Alumina supported cobalt catalysts were prepared by atomic layer deposition (ALD) of cobalt acetylacetonate precursors (Co(acac)2 and Co(acac)3). The main modes of interaction between the acetylacetonate precursors and the support were found to be the exchange reaction between the alumina OH-groups and the acac-ligands of the precursor and dissociative adsorption on coordinatively unsaturated Al3+ sites. The amount of precursor that could adsorb on the support was determined by steric hindrance. Samples were prepared using 1–5 reaction cycles, i.e. subsequent precursor addition (Co(acac)2) and calcination, resulting in catalysts containing ca. 3–10 wt.% Co. Samples were also prepared where the last calcination step was omitted, i.e. uncalcined catalysts. Calcination at 450 °C decreased the reducibility of the Co(acac)2/Al2O3 catalysts due to formation of a cobalt oxide phase strongly interacting with the support and aluminate type surface species. The reducibility increased with metal loading on both calcined and uncalcined catalysts; however the reducibility of the calcined catalysts remained lower than of the uncalcined ones. The dispersion was found to be lower on the calcined catalysts. The cobalt particle sizes on the calcined samples was ca. 8 nm and on the uncalcined 4–5 nm, for cobalt loadings of ca. 6–10 wt.%. Catalytic activity was tested by gas phase hydrogenation of toluene in temperature programmed mode (30–150 °C).  相似文献   

7.
This study provides insight into the effect of Pt dispersion on the overall rate and product distribution during NOx storage and reduction. The storage and reduction performance of Pt/BaO/A2O3 monoliths with varied Pt dispersion (3%, 8%, and 50%) and fixed Pt (2.48 wt.%) and BaO (13.0 wt.%) loadings is reported. At low temperature (<200 °C), the differences in storage and reduction activity were the largest between the three catalysts. The amount of NOx stored increased with increased dispersion, as did the amount of stored NOx that was reduced. These trends are attributed to larger Pt surface area and Pt–BaO interfacial perimeter, the latter of which enhances the spillover of surface species between the precious metal and storage components. At high temperature (370 °C), the stored NOx was almost completely regenerated for the three catalysts. However, the regeneration of the 3% dispersion catalyst was much slower, suggesting a rate limitation involving the reverse spillover of stored NOx to Pt and/or of adsorbed hydrogen from Pt to BaO. The results indicate that the catalyst dispersion and operating conditions may be tuned to achieve the desired ammonia selectivity. For the aerobic regeneration feed, the most (net) NH3 was generated by the 50% dispersion catalyst at the lowest temperature (125 °C), by the 3% dispersion catalyst at the highest temperature (340 °C), and by the 8% dispersion catalyst at the intermediate temperatures (170–290 °C). Similar trends were observed for the net production of NH3 with an anaerobic regeneration feed. A phenomenological picture is proposed that describes the effects of Pt dispersion consistent with the established spatio-temporal behavior of the lean NOx trap.  相似文献   

8.
The sintering behaviors and microwave dielectric properties of the 16CaO–9Li2O–12Sm2O3–63TiO2 (abbreviated CLST) ceramics with different amounts of V2O5 addition had been investigated in this paper. The sintering temperature of the CLST ceramic had been efficiently decreased by nearly 100 °C. No secondary phase was observed in the CLST ceramics and complete solid solution of the complex perovskite phase was confirmed. The CLST ceramics with small amounts of V2O5 addition could be well sintered at 1200 °C for 3 h without much degradation in the microwave dielectric properties. Especially, the 0.75 wt.% V2O5-doped ceramics sintered at 1200 °C for 3 h have optimum microwave dielectric properties of Kr = 100.4, Q × f = 5600 GHz, and TCF = 7 ppm/°C. Obviously, V2O5 could be a suitable sintering aid that improves densification and microwave dielectric properties of the CLST ceramics.  相似文献   

9.
Two series of supported Pd catalysts were synthesized on new mesoporous–macroporous supports (ZrO2, TiO2) labelled M (Zr and Ti). The deposition of palladium was carried out by wet impregnation on the calcined TiO2 and ZrO2 supports at 400 °C (Pd/Zr4, Pd/Ti4) and 600 °C (Pd/Zr6, Pd/Ti6) and followed by a calcination at 400 °C for 4 h. The pre-reduced Pd/MX catalysts were investigated for the chlorobenzene total oxidation and their catalytic properties where compared to those of a reference catalyst Pd/Ti-Ref (TiO2 from Huntsman Tioxide recalcined at 500 °C) and of a palladium supported on the fresh mesoporous–macroporous TiO2 (Pd/Ti). Based on the activity determined by T50, the Pd/Ti and Pd/Ti4 catalysts have been found to be more active than the reference one. Moreover activity decreased owing to the sequence: Pd/TiX  Pd/ZrX and in each series when the temperature of calcination of the support was raised. The overall results clearly showed that the activity was dependant on the nature of the support. The better activity of Pd/TiX compared to Pd/ZrX was likely due to a better reducibility of the TiO2 support (Ti4+ into Ti3+) leading to an enhancement of the oxygen mobility. Production of polychlorinated benzenes PhClx (x = 2–6) and of Cl2 was also observed. Nevertheless at 500 °C the selectivity in HCl was higher than 90% for the best catalysts.  相似文献   

10.
Ammonium nitrate is thermally stable below 250 °C and could potentially deactivate low temperature NOx reduction catalysts by blocking active sites. It is shown that NO reduces neat NH4NO3 above its 170 °C melting point, while acidic solids catalyze this reaction even at temperatures below 100 °C. NO2, a product of the reduction, can dimerize and then dissociate in molten NH4NO3 to NO+ + NO3, and may be stabilized within the melt as either an adduct or as HNO2 formed from the hydrolysis of NO+ or N2O4. The other product of reduction, NH4NO2, readily decomposes at ≤100 °C to N2 and H2O, the desired end products of DeNOx catalysis. A mechanism for the acid catalyzed reduction of NH4NO3 by NO is proposed, with HNO3 as an intermediate. These findings indicate that the use of acidic catalysts or promoters in DeNOx systems could help mitigate catalyst deactivation at low operating temperatures (<150 °C).  相似文献   

11.
The effects of B2O3 additives on the sintering behavior, microstructure and dielectric properties of CaSiO3 ceramics have been investigated. The B2O3 addition resulted in the emergence of CaO–B2O3–SiO2 glass phase, which was advantageous to lower the synthesis temperature of CaSiO3 crystal phase, and could effectively lower the densification temperature of CaSiO3 ceramic to as low as 1100 °C. The 6 wt% B2O3-doped CaSiO3 ceramic sintered at 1100 °C possessed good dielectric properties: r = 6.84 and tan δ = 6.9 × 10−4 (1 MHz).  相似文献   

12.
Dibenzothiophene (DBT) hydrodesulphurization (HDS) reaction at 3 MPa and 325–375 °C on Mo/γ-Al2O3 single-bed and Me/γ-Al2O3//SiO2//Mo/γ-Al2O3 (Me = Co or Ni) double-bed catalysts were investigated. Results indicate that ratio cyclohexylbenzene (CHB)/biphenyl (BP) or selectivity is higher when using double-beds rather than a single-bed. Synergy in dibenzothiophene hydrodesulphurization on Co//Mo and Ni//Mo double-beds is also detected. Changes in selectivity and conversion are attributed to the action of spillover hydrogen (Hso) formed in the first bed that reaches the second bed.  相似文献   

13.
Perovskite (1 − x)(0.06BiYbO3–0.94Pb(Ti0.5Zr0.5)O3)–xLiNbO3 (BYPTZ-LN) ceramics were synthesized by the conventional ceramic processing. The effect of LiNbO3 on the microstructure and piezoelectric properties was investigated. The perovskite phase and the Yb2Ti2O7 pyrochlore phase are coexisting in the BYPTZ-LN ceramics sintered at 1140 °C. The material with perovskite structure is tetragonal at x ≤ 0.04 and becomes single rhombohedral at x ≥ 0.08. A morphotropic phase boundary between rhombohedral and tetragonal phases is found in the composition range 0.04 ≤ x ≤ 0.08. Analogous to Pb(Zr,Ti)O3, the piezoelectric and electromechanical properties are enhanced for compositions near the morphotropic phase boundary. Piezoelectric constant d33values reach 290–360 pC/N. Electromechanical coefficients kp reach 0.38–0.55. The maximum values of d33, kp and Pr are obstained as x = 0.08, accompanying with the minimum values of Qm and Ec. The Curie temperature Tc and the maximum value of dielectric constant decrease with increasing LiNbO3 content. BYPTZ-LN ceramics with the high d33 value and the high thermal-depoling temperatures of >320 °C are obtained.  相似文献   

14.
A series of MnO2/ZrO2 mixed oxides were prepared in reverse microemulsions for NOx adsorption and abatement. The results show that the amount of NOx adsorbed was increased with increasing MnO2 content in various MnO2/ZrO2 samples. The maximum uptake value of NOx was 27.66 mg NOx/g adsorbent on the 40% Mn–Zr sample at 200 °C with NOx initial adsorption rate as 2.63 mg/(g adsorbent min). TPD results show that the complete desorption of NOx was easily obtained by heating the sample to 450 °C, and the temperature for the complete desorption can be further decreased to 210 °C by adding carbon monoxide into the argon desorption streams. Furthermore, water vapor was found to reduce NOx adsorption capacity because of its stronger competitive adsorption with NOx species. It is noteworthy that a small amount of sulfur dioxide could significantly increase the initial rate of NOx adsorption although it slightly decreased the NOx adsorption capacity.  相似文献   

15.
The employment of mineral SrSO4 crystals and powders for preparing SrTiO3 compound was investigated, with coexistence of Ti(OH)4·4.5H2O gel under hydrothermal conditions, at various temperatures (150–250 °C) for different reaction intervals (0.08–96 h) in KOH solutions with different concentrations. The complete dissolution of the SrSO4 crystal occurred at 250 °C for 96 h in a 5 M KOH solution, resulting in the synthesis of SrTiO3 particles with two different shapes (peanut-like and cubic). In contrast, very fine SrTiO3 pseudospherical particles were crystallized when SrSO4 powders were employed as precursor. Variations on the SrTiO3 particle shape and size were found to be caused by the differences in the dissolution rate of the SrSO4 phase in the alkaline KOH solution. The crystallization of SrTiO3 particles was achieved by a bulk dissolution–precipitation mechanism of the raw precursors, and this mechanism was further accelerated by increasing the reaction temperature and concentration of the alkaline media. Kinetic data depicted that the activation energy required for the formation of SrTiO3 powders from the complete consumption of a SrSO4 single crystal plate under hydrothermal conditions, is 27.9 kJ mol−1. In contrast, when SrSO4 powders were employed (28–38 μm), the formation of SrTiO3 powder proceeded very fast even for a short reaction interval of 3 h at 250 °C in a 5 M KOH solution.  相似文献   

16.
The effect of partial substitution of Co by Pd in LaCoO3 perovskite structure (i.e., LaCo0.95Pd0.05O3) and the reductive diffusion of Pd from the bulk of perovskite to its surface, thus forming Pd nanoparticles, on CO and C3H8 oxidation present in air (simulated exhaust gas) are reported. X-ray powder diffraction (XRD) analyses confirm the perovskite structure for the catalysts. Scanning electron microscopy (SEM) and BET surface area measurements show that partial substitution of Co by Pd decreases the crystallite size of the perovskite and therefore increases its surface area. H2-temperature programmed reduction (TPR) experiments reveal that Pd reduces at 135 °C and facilitates the reduction of Co in the perovskite structure. By partial reduction of the Pd containing catalyst at 180 °C for 30 min, the complete oxidation temperatures of CO and C3H8 decrease by about 70 and 50 °C, respectively.The reduction duration of the Pd containing catalyst strongly affects the T50 and T90 temperatures (temperatures at which 50 and 90% conversion occurs, respectively) and has an optimum, where it decreases by increasing the reduction temperature of the catalyst.  相似文献   

17.
Phosphorous-doped NiMo/Al2O3 hydrodesulfurization (HDS) catalysts (nominal Mo, Ni and P loadings of 12, 3, and 1.6 wt%, respectively) were prepared using ethyleneglycol (EG) as additive. The organic agent was diluted in aqueous impregnating solutions obtained by MoO3 digestion in presence of H3PO4, followed by 2NiCO3·3Ni(OH)2·4H2O addition. EG/Ni molar ratio was varied (1, 2.5 and 7) to determine the influence of this parameter on the surface and structural properties of synthesized materials. As determined by temperature-programmed reduction, ethyleneglycol addition during impregnation resulted in decreased interaction between deposited phases (Mo and Ni) and the alumina carrier. Dispersion and sulfidability (as observed by X-ray photoelectron microscopy) of molybdenum and nickel showed opposite trends when incremental amounts of the organic were added during catalysts preparation. Meanwhile Mo sulfidation was progressively decreased by augmenting EG concentration in the impregnating solution, more dispersed sulfidic nickel was evidenced in materials synthesized at higher EG/Ni ratios. Also, enhanced formation of the so-called “NiMoS phase” was registered by increasing the amount of added ethyleneglycol during simultaneous Ni–Mo–P–EG deposition over the alumina carrier. That fact was reflected in enhanced activity in liquid-phase dibenzothiophene HDS (batch reactor, T = 320 °C, P = 70 kg/cm2) and straight-run gas oil desulfurization (steady-state flow reactor), the latter test carried out at conditions similar to those used in industrial hydrotreaters for the production of ultra-low sulfur diesel (T = 350 °C, P = 70 kg/cm2, LHSV = 1.5 h−1 and H2/oil = 2500 ft3/bbl).  相似文献   

18.
The phase stability of SrCo0.8Fe0.2O3−d perovskite doped with niobium was studied by in situ high-temperature X-ray diffraction in the temperature range of 30–1000 °C and oxygen partial pressure 0.2–10−5 atm. The stability of the cubic perovskite structure in a wide range of oxygen partial pressures is the main advantage of SrCo0.8−xFe0.2NbxO3−d (x = 0.1–0.3) system in comparison with SrCo0.8Fe0.2O3−d. It is suggested that equilibrium of the thermal expansion with changes of the oxygen non-stoichiometry leading to the same lattice parameters in the oxidizing and reducing environments at the catalytic temperatures is a necessary requirement for stable operation of perovskite as an oxygen-conducting membrane. In the case of SrCo0.8−xFe0.2NbxO3−d perovskite this condition is met at x = 0.2. This makes the SrCo0.6Fe0.2Nb0.2O3−d composition promising for application as oxygen-conducting membrane.  相似文献   

19.
We focused on the linear negative thermal expansion of Y2W3O12 in a wide-temperature range and on the chemical stability of ZrSiO4 in the fabrication of the composite material ZrSiO4/Y2W3O12 with a zero-thermal-expansion. The compact composed of Y2W3O12 and ZrSiO4 had a thermal shrinkage rate smaller than that of Y2W3O12 and higher than that of ZrSiO4. SEM–EDX observation clarified that the ZrSiO4/Y2W3O12 sintered body fabricated at 1400 °C for 10 h had a microstructure composed of ZrSiO4 and Y2W3O12 grains, and XRD indicated that only ZrSiO4 and Y2W3O12 phases existed in the sintered body. The relative density of the ZrSiO4/Y2W3O12 sintered body reached 92%, which was larger than that of the ZrSiO4 sintered body because Y2W3O12 grains could be sintered at lower temperatures. The average linear thermal expansion coefficients of the ZrSiO4/Y2W3O12 sintered body were −0.4 × 10−6 and −0.08 × 10−6 °C−1 in the temperature ranges from 25 to 500 °C and from 25 to 1000 °C, respectively, which showed an almost zero-thermal-expansion.  相似文献   

20.
In this paper, the CuO/TiO2 catalysts prepared by the deposition–precipitation (DP) method were extensively investigated for CO oxidation reaction. The structural characters of the CuO/TiO2 catalysts were comparatively investigated by TG-DTA, XRD, and XPS measurements. It was shown that the catalytic behavior of CuO/TiO2 catalysts greatly depended on the TiO2-support calcination temperature, the CuO loading amount and the CuO/TiO2 catalysts calcination temperature. CuO supported on the anatase phase of TiO2-support calcined at 400 °C showed better catalytic activity than those supported on TiO2 calcined at 500 and 700 °C. Among all our investigated catalysts with CuO loading from 2% to 12%, the catalyst with 8 wt% CuO loading exhibited the highest catalytic activity. The optimum calcination temperature of the CuO/TiO2 catalysts was 300 °C. The XRD results indicated that the catalytic activity of the CuO/TiO2 catalysts was related to the crystal phase and particle size of TiO2 support and CuO active component.  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