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.
The catalytic decomposition of dichlorodifluoromethane (CFC‐12) in the presence of water vapor on a series of SO42?‐promoted solid adds was investigated. CFC‐12 was decomposed completely on SO42?/ZrO2, SO42?/TiO2, SO42?/SnO2, SO42?/ Fe2O3 and SO42–/Al2O3 at 265°C, 270°C, 325°C, 350°C and 325°C, respectively, and the selectivity to by‐products was neglectable. Obvious deactivation was found on SO42?/ZrO2 and SO42?/Al2O3, during several hours on stream, while the catalytic activity was maintained on SO42?/TiO2, SO42?/SnO2 and SO42?/Fe2O3 for 240 h on stream.  相似文献   

2.
The three‐component Mannich reaction of among dimethyl malonate, aromatic primary amine, and aromatic aldehyde was made successfully in the presence of solid acidic catalyst SO42/TiO2, with excellent catalytic activity, as compared to SO42/γ‐Al2O3 and SO42/ZnO. To the best of our knowledge, SO42/TiO2 prepared at varied calcination temperatures can perform different intensities of Lewis and Brønsted acidities. Because of this point, under the optimum conditions, the effect of SO42/TiO2 (prepared at 200°C) was much more than that of SO42/TiO2 (prepared at 300°C or 400°C) in the three‐component Mannich reaction. In observing ionization activation mode of the three‐component Mannich reaction, it disclosed that the plausible mechanism possibly undergoes formation of aldimines and transformation of aldimines into β‐amino esters by applying solid acidic catalyst SO42/MxOy.  相似文献   

3.
The solid state reactions between TiO2 and Na2S2O8 or K2S2O8 have been investigated using TG, DTG, DTA, IR, and X-ray diffraction studies in the range of 20 to 1000°C.It has been shown that TiO2 reacts stoichiometrically (1 : 1) with Na2S2O8 in the range of 160 and 220°C forming the complex sodium monoperoxodisulfato—titanium(IV) as characterized by IR and X-ray analysis. The new complex then decomposes into the reactants above 190°C.An exothermic reaction has been observed between TiO2 and molten K2S2O7 at mole ratio 1:2 respectively and higher, in the range of 280 and 350°C. The IR and X-ray analyses have shown the formation of a complex namely, potassium tetrasulfato titanium(IV) for which the formula and structure have been proposed. This complex decomposes at higher temperatures into K2SO4 and a mixed sulfate of potassium and titanium. The mixed sulfate melts at 620°C and decomposes into K2SO4, TiO2, and the gaseous SO3.On the other hand, Na2S2O8 decomposes in a special mode producing a polymeric product of Na10S9O32. Decomposition of this species occurs after melting at 560°C into Na2SO4 and sulfur oxides. The decomposition reaction has been proved to be catalysed by TiO2 itself.  相似文献   

4.
A series of composite photocatalysts based on titanium dioxide deposited on the surface of a zirconium phosphate support were synthesized under different synthesis and heat-treatment conditions. The study of the photodestruction kinetics of Rhodamine C showed that the synthesized composites possess high photocatalytic activity that is competitive with the activity of a commercial Hombikat UV100 photocatalyst. The composites based on zirconium phosphate treated with isopropanol at the precipitation stage whereupon heated at 550°C exhibit the highest photocatalytic activity after heating at 750°C. It was found that such zirconium phosphate support has the largest specific surface area (270 m2/g). After heating at 550°C, the surface becomes more stable to the subsequent heating to 750°C, which is necessary for the most complete crystallization of TiO2 ensuring its high photocatalytic characteristics.  相似文献   

5.
The kinetics and equilibria of SO2 sorption in Kapton polyimide film have been studied at temperatures from 25 to 55°C and equilibrium sorption pressures up to 0.76 atm. The data are described well by the dual-mode model of sorption and transport in glassy polymers. The assumption of “partial immobilization” is required to correlate the transport data: the mobility of the Langmuir component of the sorbed population relative to the Henry's-law component is close to zero at 25°C, and increases to roughly 5% of the Henry's-law component mobility at 55°C. The heat of sorption is anomalously low, suggesting the presence of residual solvent in the film, a suggestion confirmed by annealing studies. However, the study also shows that partial removal of the residual has a relatively minor effect on the measured SO2 sorption level and transport rates at 55°C.  相似文献   

6.
The effects of doping with CeO2 and calcination temperature on the physicochemical properties of the NiO/Al2O3 system have been investigated using DTA, XRD, nitrogen adsorption measurements at −196°C and decomposition of H2O2 at 30–50°C. The pure and variously doped solids were subjected to heat treatment at 300, 400, 700, 900 and 1000°C. The results revealed that the specific surface areas increased with increasing calcination temperature from 300 to 400°C and with doping of the system with CeO2. The pure and variously doped solids calcined at 300 and 400°C consisted of poorly crystalline NiO dispersed on γ-Al2O3. Heating at 700°C resulted in formation of well crystalline NiO and γ-Al2O3 phases beside CeO2 for the doped solids. Crystalline NiAl2O4 phase was formed starting from 900°C. The degree of crystallinity of NiAl2O4 increased with increasing the calcination temperature from 900 to 1000°C. An opposite effect was observed upon doping with CeO2. The NiO/Al2O3 system calcined at 300 and 400°C has catalytic activity higher than individual NiO obtained at the same calcination temperatures. The catalytic activity of NiO/Al2O3 system increased, progressively, with increasing the amount of CeO2 dopant and decreased with increasing the calcination temperature.  相似文献   

7.
Various nickel aluminium mixed hydroxide samples of different compositions were prepared by co-precipitation from their nitrate solutions using dilute NH4OH. Additional samples were prepared by impregnation of hydrated Al2O3, preheated at 600 and 900°C, with nickel nitrate solution in an equimolar ratio. The thermal decomposition of different mixed solids was studied using DTA. The X-ray investigation of thermal products of the mixed solids was also studied.The results obtained revealed that the presence of NiO up to 33.3 mole % with aluminium oxide much enhanced the degree of crystallinity of the γ-Al2O3 phase. In contrast, the presence of Al2O3 much retarded the crystallization process of the NiO phase. With the exception of samples containing 20 mole% NiO, all the mixed hydroxide samples, when heated in air at 900°C, led to the formation of well-crystalline Ni Al2O4 spinel, alone, or together with either NiO or γ-Al2O3, depending on the composition of the mixed oxide samples. The solid containing 20% NiO and heated at 900°C was constituted of amorphous NiO dispersed in γ-Al2O3. Heating the nickel nitrate-impregnated Al2O3 in air at 800–1000°C led to the formation of Ni Al2O4 together with non-reacted NiO and γ-Al2O3. The degree of crystallinity of the spinel was found to increase by increasing the calcination temperature of the impregnated solids from 800 to 1000°C and by increasing the preheating temperature of the hydrated Al2O3 employed from 600 to 900°C.  相似文献   

8.
Manganese dioxides were prepared electrolytically in the temperature range 10–95°C. The oxides prepared at 25°C or lower exhibited a high ion exchange capacity equivalent to one proton per two Mn atoms. Those prepared at higher temperatures exhibited correspondingly lower capacities and the 95°C preparation showed only surface exchange. The Li+ exchanged solid formed a spinel LiMn2O4 on heating and was converted to HMn2O4 on treatment with dilute acid. The protonated solid retained the spinel structure and showed a high specificity for Li+. A corresponding sodium ion exchanged phase yielded a birnessite-like phase on heating followed by treatment in boiling water. The sodium ion was exchanged with acid to yield a layered compound of composition H4Mn9O18 · 7H2O. It is suggested that λ-MnO2 and Mn7O13 are deprotonated versions of the hydrogen ion exchanged species. It is further suggested that similar exchange reactions with insertion type compounds should lead to new types of inorganic ion exchangers.  相似文献   

9.
The polycrystalline solids TiO2Fe2O3, with iron contents in the range 0–10 at.%, prepared by coprecipitation and by impregnation, and treated in air at temperatures in the range 500–1000°C, have been studied by X-ray, ESR, and Mössbauer methods. The TiO2 in the samples treated at 800 and 1000°C always forms the rutile phase and the Fe3+ has a rather low solubility in it (~0.1 at.%). The Fe3+ in excess forms the antiferromagnetic pseudobrookite phase (Fe2TiO5). The samples treated at 500 and 650°C show a dependence on the preparation method. Those prepared by coprecipitation give at 500°C the pure anatase phase in which the Fe3+ has a higher solubility (≥ 1%); those prepared by impregnation give the anatase phase accompanied by a variable amount of rutile. The treatment at 650°C provokes the partial transformation of anatase to rutile and the complete development of the Fe2TiO5 phase. The relevance of these results to the photocatalytic properties shown by these solids for the photoreduction of dinitrogen to ammonia is discussed.  相似文献   

10.
The synthesis of SiO2 core-TiO2 shell composites from a titanium dioxide sol and a suspension of microspherical silicon dioxide is described. The main factors ensuring the formation of a composite with a preset morphology are the size and charge of the TiO2 sol particles (10–45 nm) and silicon dioxide core particles (300–700 nm), the pH values of the suspensions of the starting components and the resulting composite, and the proportions and way of mixing of the siliconand titanium-containing components. The SiO2 core-TiO2 shell composites show high photocatalytic activity in the degradation of Rhodamine FL-BM dye (rate constant of k = 0.0813 min−1) and are much more active than precipitated TiO2 powder (k = 0.0022 min−1). The activity of the composite is determined by the calcination temperature (700–800°C), by the proportion and accessibility of the active component (TiO2), and by the presence of a dopant (P2O5).  相似文献   

11.
Hierarchical flowerlike β‐Ni(OH)2 superstructures composed of intermeshed nanoflakes are synthesized by hydrothermal treatment with a mixed solution of C2H4(NH2)2, NaOH, and Ni(NO3)2. The as‐prepared β‐Ni(OH)2 superstructures could be easily changed into NiO superstructures without great morphology change by calcination at 400 °C for 5 h. Furthermore, the TiO2 nanoparticles can be homogeneously deposited on the surface of NiO superstructures by dispersing β‐Ni(OH)2 powders in Ti(OC4H9)4–C2H5OH mixed solution and then vaporizing to remove the ethanol at 100 °C, and finally calcination at 400 °C for 5 h. The prepared NiO/TiO2 p–n junction superstructures show much higher photocatalytic activity for photocatalytic degradation of p‐chlorophenol aqueous solution than conventional TiO2 powders and NiO superstructures prepared under the same experimental conditions. An obvious enhancement in the photocatalytic activity can be related to several factors, including formation of hierarchical porous structures, dispersion of TiO2 particles on the surface of NiO superstructures, and production of a pn junction. Further results show that NiO/TiO2 composite superstructures can be more readily separated from the slurry system by filtration or sedimentation after photocatalytic reaction and re‐used, compared with conventional powder photocatalysts. After many recycling experiments for the photodegradation of p‐chlorophenol, the NiO/TiO2 composite sample does not exhibit any great activity loss, confirming that NiO/TiO2 sample is stable and not photocorroded.  相似文献   

12.
Mesoporous WO3–TiO2 support was synthesized by hydrothermal method, mesoporous V2O5/WO3–TiO2 catalyst was synthesized by impregnation method and used for selective catalytic reduction (SCR) of NOx with a excellent NOx conversion at a wider operating temperature ranging from 200 to 460?°C. In the range of 260–440?°C, NOx conversion reached to 98.6%, and nearly a complete conversion. Even with the existence of 300 ppm SO2, NOx conversion was only a little decline. The catalyst was characterized by a series of techniques, such as XRD, BET, XPS, TEM, Raman and H2-TPR. It was concluded that V2O5/WO3–TiO2 catalyst was ascribe to antase TiO2, and also the high crystallinity of anatase TiO2 could improve the SCR performance. More interested, V2O5/WO3–TiO2 catalyst exhibited the typical mesoporous structure according to the BET results. In addition, the TEM results indicated that the active components of V and W were well-dispersed on the surface of TiO2, while the enhancement of dispersion could improve the activity of catalysts. More importantly, the concentration ratio of V4+/(V5+?+?V4+?+?V3+) performed the key role in improving the activity of V2O5/WO3–TiO2 catalyst.  相似文献   

13.
New potassium-conducting solid electrolytes in the mixed gallate-ferrite systems (1 − x)Ga2O3 · xFe2O3 · 0.25TiO2 · K2O and 1.5[(1 − x)Ga2O3 · xFe2O3] · TiO2 · 2K2O are synthesized and studied. The electrolytes exhibit high ionic conductivity in the test temperature range of 300 to 750°C (above 10−2 S/cm at 300°C and above 10−1 S/cm at 700°C). An increase in the conductivity with increasing concentration of iron in the specimens is a general tendency. Possible reasons for the effect of Ga/Fe ratio in the structure of solid electrolytes on their transport properties are discussed.  相似文献   

14.
Anatase TiO2 nanosheets (TiO2 NS) with dominant (001) facets and TiO2 nanoparticles (TiO2 NP) with dominant (101) facets are fabricated by hydrothermal hydrolysis of Ti(OC4H9)4 in the presence and absence of hydrogen fluoride (HF), respectively. Adsorption of N719 onto the as‐prepared samples from ethanol solutions is investigated and discussed. The adsorption kinetic data are modeled using the pseudo‐first‐order, pseudo‐second‐order, and intraparticle diffusion kinetics equations, and indicate that the pseudo‐second‐order kinetic equation and intraparticle diffusion model can better describe the adsorption kinetics. Furthermore, adsorption equilibrium data of N719 on the as‐prepared samples are analyzed by Langmuir and Freundlich models; this suggests that the Langmuir model provides a better correlation of the experimental data. The adsorption capacities (qmax) of N719 on TiO2 NS at various temperatures, determined using the Langmuir equation, are 65.2 (30 °C), 68.2 (40 °C), and 76.6 (50 °C) mg g−1, which are smaller than those on TiO2 NP, 92.4 (30 °C), 100.0 (40 °C), and 108.2 (50 °C) mg g−1, respectively. The larger adsorption capacities of N719 for TiO2 NP versus NS are attributed to its higher specific surface areas. However, the specific adsorption capacities (qmax/SBET) at various temperatures are 1.5 (30 °C), 1.6 (40 °C), and 1.7 (50 °C) mg m−2 for TiO2 NS, which are otherwise higher than those for NP, 0.9 (30 °C), 1.0 (40 °C), and 1.1 (50 °C) mg m−2, respectively. The larger specific adsorption capacities of N719 for TiO2 NS versus NP are because the (001) surface is more reactive for dissociative adsorption of reactant molecules compared with (101) facets. Notably, the qmax and qmax/SBET for both TiO2 samples increase with increasing temperature, suggesting that adsorption of N719 on the TiO2 surface is an endothermic process, which is further confirmed by the calculated thermodynamic parameters including free energy, enthalpy, and entropy of adsorption process. The present work will provide a new understanding on the adsorption process and mechanism of N719 molecules onto TiO2 NS and NP, and this should be of great importance for enhancing the performance of dye‐sensitized solar cells.  相似文献   

15.
Fine‐powdered, heterostructured, nanoporous Bi2O3–TiO2 (BTO) was synthesized by a green approach using ultrasonication, with the mole ratio Bi/Ti of 1:1 and calcined at different temperatures. The physical and optical properties of the mixed oxides were investigated. The phase structure, as identified by X‐ray diffraction (XRD), showed the appearance of new phases as a function of the calcination temperature. Morphological examinations indicated the formation of a nanoporous structure with a drastic change in morphology at the calcination temperature of 850°C from a globule to a rod‐shaped structure, which further got transformed to a rocky appearance at 1200°C. Doping with Bi2O3 led to the lowering of the bandgap of TiO2 from 3.25 to 2.5 eV. A BTO nanocatalyst calcined at 450°C exhibited promising photocatalytic activity for the degradation of quinalphos (QP) (92%) after a time interval of 100 min under visible light and at the optimum pH 8. The kinetics of degradation of QP showed that it follows a pseudo‐first‐order path with a rate constant 0.01267 min?1. The synthesized BTO mixed oxide showed profound improvement in photocatalytic activity in the visible region as compared to TiO2.  相似文献   

16.
This paper presents a study regarding the obtaining of NiCr2O4 by two new unconventional synthesis methods: (i) the first method is based on the formation of Cr(III) and Ni(II) carboxylate-type precursors in the redox reaction between the nitrate ion and 1,3-propanediol. The thermal decomposition of these complex combinations, at ~300 °C, leads to an oxide mixture of Cr2O3+x and NiO, with advanced homogeneity, small particles and high reactivity. On heating this mixture at 500 °C, Cr2O3 reacts with NiO to form NiCr2O4, which was evidenced by FT-IR and X-ray diffractometry (XRD) analysis; (ii) the second method starts from a mechanical mixture of (NH4)2Cr2O7 and Ni(NO3)2·6H2O. On heating this mixture, a violent decomposition at 240 °C with formation of an oxides mixture (Cr2O3 + CrO3) and NiO takes place. On thermal treatment up to 500 °C, an intermediary phase NiCrO4 is formed, which by decomposition at ~700 °C leads to NiCr2O4, evidenced by FT-IR and XRD analysis. NiCr2O4 is formed, in both cases, starting with a temperature higher than 400 °C, when the non-stoichiometric chromium oxide (Cr2O3+x ) loses the oxygen excess and turns to stoichiometric chromium oxide (Cr2O3), which further reacts with NiO.  相似文献   

17.
The present study deals with preparation and characterization of spinel mixed oxide systems NiM 2 III O4, where MIII?=?FeIII, CrIII. In order to obtain 50% NiFe2O4/50% SiO2 and 50% NiCr2O4/50% SiO2 nanocomposite, we have used a versatile route based on the thermal decomposition inside the SiO2 matrix, of some particular precursors, coordination compounds of the involved MII and MIII cations with dicarboxylate ligands. The ligands form in the redox reaction between metal nitrates mixture and 1,3-propanediol at the heating around 140?°C of the gels (tetraethylorthosilicate?Cmetal nitrates?C1,3-propanediol?Cwater). The as-obtained precursors, embedded in silica gels, have been characterized by FT-IR spectrometry and thermal analysis. Both precursors thermally decompose up to 350?°C leading to the formation of the corresponding metal oxides inside the silica matrix. X-ray diffraction of the annealed powders have evidenced the formation of NiFe2O4 starting with 600?°C, and NiCr2O4 starting with 400?°C. This behavior can be explained by the fact that, by thermal decomposition of the Fe(III) carboxylate at 300?°C, the spinelic phase ??-Fe2O3 is formed, which interacts with the NiO, forming the ferrite nuclei. By thermal decomposition of chromium carboxylate, a nonstoichiometric chromium oxide (Cr2O3+x ) is formed. In the range 380?C400?°C, Cr2O3+x turns into Cr2O3 which immediately interacts with NiO leading to the formation of nickel chromites nuclei inside the pores of silica matrix. Both spinels have been obtained as nanocrystalites homogenously dispersed as resulted from XRD and TEM data.  相似文献   

18.
The reduction of H2SO4 to SO2 occurs with a relatively good efficiency only at high temperatures, in the presence of catalysts. Some experimental results, regarding conversion of sulfuric acid (96 wt.%) to sulfur dioxide and oxygen, are reported. The reduction has been performed at 800 ?C 900°C and atmospheric pressure, in a tubular quartz reactor. The following commercial catalysts were tested: Pd/Al2O3 (5 wt.% and 0.5 wt.% Pd), Pt/Al2O3 (0.1 wt.% Pt) and ??-Fe2O3. The fresh and spent catalysts were characterized by X-Ray diffraction and BET method. The highest catalytic activity was determined for 5 wt.% Pd/Al2O3, a conversion of 80% being obtained for 5 hours time on stream, at 9 mL h?1 flow rate of 96 wt.% H2SO4. A conversion of 64% was determined for 0.5 wt.% Pd/Al2O3 and 0.1 wt.% Pt/Al2O3. For ??-Fe2O3, a less expensive catalyst, a conversion of 61% for about 60 hours was obtained.   相似文献   

19.
TiO2 nanopowders doped by Si and Zr were prepared by sol–gel method. The effects of Si and Zr doping on the structural, optical, and photo-catalytic properties of titania nanopowders have been studied by X-ray diffraction (XRD), scanning electron microscopy, transmission electron microscopy, and UV–Vis absorption spectroscopy. XRD results suggest that adding impurities has a significant effect on anatase phase stability, crystallinity, and particle size of TiO2. Titania rutile phase formation in ternary system (Ti–Si–Zr) was inhibited by Zr4+ and Si4+ co-doped TiO2 in high temperatures (500–900 °C) and 36 mol% anatase composition is retained even after calcination at 1,000 °C. The photocatalyst activity was evaluated by photocatalytic degradation kinetics of aqueous methylen orange under visible radiation. The results show that the photocatalytic activity of the 20 %Si and 15 %Zr co-doped TiO2 nanopowders have a larger degradation efficiency than pure TiO2 under visible light.  相似文献   

20.
A series of tungsten‐doped Titania photocatalysts were synthesized using a low‐temperature method. The effects of dopant concentration and annealing temperature on the phase transitions, crystallinity, electronic, optical, and photocatalytic properties of the resulting material were studied. The X‐ray patterns revealed that the doping delays the transition of anatase to rutile to a high temperature. A new phase WyTi1‐yO2 appeared for 5.00 wt% W‐TiO2 annealed at 900 °C. Raman and diffuse reflectance UV–Vis spectroscopy showed that band gap values decreased slightly up to 700 °C. X‐ray photoelectron spectroscopy showed that surface species viz. Ti3+, Ti4+, O2?, oxygen‐vacancies, and adsorbed OH groups vary depending on the preparation conditions. The photocatalytic activity was evaluated via the degradation of methylene blue using LED white light. The degradation rate was affected by the percentage of dopants. The best photocatalytic activity was achieved with the sample labeled 5.00 wt% W‐TiO2 annealed at 700 °C.  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