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.
A novel catalyst based on copper-silver was developed to solve the contradiction between the high conversion temperature of Cu-based catalyst and low N2 selectivity of Ag-based catalyst during selective oxidation of ammonium gas. The Cu-Ag-based catalyst (Cu 5 wt.%-Ag 5 wt.%/Al2O3) displayed a relatively low complete conversion temperature (<320 °C) with a high N2 selectivity (>95%). Increasing loading of Cu and Ag decreases N2 selectivity. The low N2 selectivity of Ag-based catalyst is possibly related to the formation of Ag2O crystals. Improvement of N2 selectivity of Ag-based catalyst was obtained by doping Cu to decrease crystallized Ag2O phase. The temperature programmed reaction (TPR) data show that N2O is the main byproduct of oxidation of ammonia at temperature lower than 200 °C. Two bands of nitrate species at 1541 and 1302 cm−1 were observed on Ag 10 wt.%/Al2O3 at the temperature higher than 250 °C, which indicates the formation of NOx during the selective catalytic oxidation of ammonia. No nitrate species was observed on Cu 10 wt.%/Al2O3 and Cu 5 wt.%-Ag 5 wt.%/Al2O3, while only one nitrate species (1543 cm−1) existed on Cu 10 wt.%-Ag 10 wt.%/Al2O3. We proposed that mixing Ag with Cu inhibited the formation of NOx during the selective catalytic oxidation of ammonia over Cu-Ag/Al2O3.  相似文献   

2.
Diffusion of ammonia and ammonium ions in sulphonic acid cation exchangers (gel Purolite SGC 100 × 10 MBH and macroporous Purolite C 160 MBH) from the solutions, representing the composition of “caustic condensate” (waste of nitrogen fertilizers production) is affected by pH of initial solution and structure of the matrix of cation exchanger. In gel matrix the effective intraparticle diffusivity (Def) depends greatly on the solution pH because of shrinkage in alkaline and swelling in acidic medium: on decreasing the initial concentration of ammonia from 0.214 to 0.003 and increasing that of ammonium nitrate from 0 to 0.214 mol l−1 instead, the effect of ion exchange leads to a decrease in pH, resulting in swelling and increase in Def from 0.1 to 0.34 × 10−10 for gel Purolite SGC 100 × 10 MBH and variation of 0.18–0.11 × 10−10 m2 s−1 for macroporous Purolite C 160 MBH (resistant to shrinkage and swelling).

In Purolite C 160 MBH both macropore diffusivity (0.07–0.29 × 10−10 m2 s−1) and gel (solid phase) diffusivity (0.06–0.19 × 10−10 m2 s−1) are higher than micropore diffusivity (0.28–0.56 × 10−18 m2 s−1).

With respect to the effective intraparticle diffusivity, resistance to nitric acid, used for the regeneration, and high concentration of ammonium nitrate in eluate (up to 110 g l−1), Purolite C 160 MBH has been installed for the conversion of ammonia and ammonium ions to ammonium nitrate reusable in the fertilizers production. This allows minimizing the economic loss and preventing the environmental contamination.  相似文献   


3.
Keun Woo Cho  Hyuk Sang Kwon   《Catalysis Today》2007,120(3-4):298-304
Co and Co–P catalysts electroplated on Cu in sulfate based solution without or with an addition of H2PO2 ions were developed for hydrogen generation from alkaline NaBH4 solution. The microstructures of the Co and Co–P catalysts and their hydrogen generation properties were analyzed as a function of cathodic current density and plating time during the electrodeposition. An amorphous Co–P electrodeposit with micro-cracks was formed by electroplating in the sulfate based solution containing H2PO2 ions. It was found that the amorphous Co–P catalyst formed at 0.01 A/cm2 exhibited 18 times higher catalytic activity for hydrolysis of NaBH4 than did the polycrystalline Co catalyst. The catalytic activity of the electrodeposited Co–P catalyst for hydrolysis of NaBH4 was found to be a function of both cathodic current density and plating time, that is, parameters determining the concentration of P in the Co–P catalyst. Especially, Co–13 at.% P catalyst electroplated on Cu in the Co–P bath at a cathodic current density of 0.01 A/cm2 for 1080 s showed the best hydrogen generation rate of 954 ml/min g-catalyst in 1 wt.% NaOH + 10 wt.% NaBH4 solution at 30 °C.  相似文献   

4.
H-AITS-1 zeolite with Si/Ti = 50 and Si/Al = 50 was employed in preparing catalyst samples by ion-exchange and impregnation with a copper nitrate solution to obtain 0.24–1.15 wt.% and 1.5, 2 and 2.5 wt.% Cu loading, respectively. The catalytic properties for the NO decomposition were compared with that of Cu-ZSM-5 (Si/Al = 25 with 2 wt.% Cu loading) and similarity was found between the AITS-1 based samples and Cu-ZSM-5. Due to the higher acidity, the activity at 500°C per total copper atoms (an apparent turnover frequency, TOF) was significantly higher over Cu based AITS-1 samples being 2–3 × 10−3 s−1 as compared to 1 × 10−3 s−1 measured on Cu-ZSM-5. For the ion-exchanged Cu-AITS-1 there was an increase in TOF with increasing copper content, whereas on the impregnated samples a decrease in TOF was found. On all catalysts there was a maximum in the NO conversion at 500–550°C. The amount of NO per copper atom measured by temperature programmed desorption (TPD) was about the same as that on Cu-ZSM-5 and the features of the TPD were also similar. At the first contact of the catalyst at 500°C with the 2 vol% NO/Ar gas a transient N2O formation and a considerable delay in the O2 formation was observed. This could, however, be reproduced only on fresh catalyst, while all further transients showed different but reproducible features using the same sample.  相似文献   

5.
Pd-only three-way catalysts prepared by the sol–gel method were investigated by the three-way catalytic performance test with a simulated exhaust gas in a continuous U-tube quartz reactor at a gas hourly space velocity of 72 000 h−1. The catalysts were characterized with XRD, XPS, BET surface area and pore volume. The activity and thermal stability of the Pd–Al2O3 catalyst prepared at pH 10 were superior to those at pH 4 during hydrolysis and condensation, which could be explained by the anchoring effect. Zr and V were found to be good promoters for the enhancement of the thermal stability and SO2 resistance, respectively. Optimally formulated catalyst, Pd(1)–V(2)–Zr(10)–Al2O3, was thermally stable up to 900oC and showed a much more improved low-temperature activity and excellent SO2 resistance.  相似文献   

6.
Catalytic methane combustion was studied over the palladium single crystals Pd(1 1 1), Pd(1 0 0) and Pd(1 1 0). Under lean reaction conditions at 600 K (O2:CH4 = 10:1), stoichiometric palladium oxide was formed with an increase in surface area by a factor of approximately two. The oxide phase formed a “cauliflower-like” surface structure composed of approximately 4 nm sized semispherical oxide agglomerates. This oxide structure was independent of the original metal single crystal orientation. The turnover rates over the oxide structure starting with metal single crystals were 0.7 s−1 on Pd(1 1 1), 0.9 s−1 on Pd(1 0 0) and 0.9 s−1 on Pd(1 1 0) at 600 K, 160 Torr O2, 16 Torr CH4, 1 Torr H2O and N2 balance to 800 Torr, suggesting that the methane combustion reaction is independent of the initial structure of the catalyst. Methane combustion on palladium single crystals experienced an activation period in which the initial turnover rates based on the initial Pd surface area were about 1/8–1/4 of the steady-state rates determined based on the oxide surface area. This activation period was caused by the slow oxidation of palladium single crystals and concomitant surface area increase during reaction. The increase in surface area happened mostly in the first 10 min of reaction. Carbon dissolution into the crystal was not found during methane combustion under reaction conditions in excess oxygen.  相似文献   

7.
Pd nanoparticles (2 nm) stabilized in the micelle core of poly(ethylene oxide)-block-poly-2-vinylpyridine were studied in 2-butyne-1,4-diol partial hydrogenation. Both unsupported micelles (0.6 kgPd/m3) and supported ones on γ-Al2O3 (0.042 wt.% Pd) showed nearly 100% selectivity to 2-butene-1,4-diol up to 94% conversion. The only side product observed was 2-butane-1,4-diol. The catalysis was ascribed to Pd nanoparticles’ surface modified by pyridine units of micelles and alkali reaction medium (pH of 13.4). TOFs over the unsupported and supported catalysts were found to be 0.56 and 0.91 s−1 (at 323 K, 0.6 MPa H2 pressure, solvent 2-propanol/water = 7:3), respectively. Reaction kinetics fit the Langmuir–Hinshelwood model assuming weak hydrogen adsorption. The experiments on the catalyst reuse showed that Pd nanoparticles remain inside the micelle core, but the micelles slightly desorbed (less then 5%) during the catalytic run.  相似文献   

8.
The selective catalytic reduction (SCR) of NO by hydrocarbon is an efficient way to remove NO emission from lean-burn gasoline and diesel exhaust. In this paper, a thermally and hydrothermally stable Al–Ce-pillared clay (Al–Ce-PILC) was synthesized and then modified by SO42−, whose surface area and average pore diameter calcined at 773 K were 161 m2/g and 12.15 nm, respectively. Copper-impregnated Al–Ce-pillared clay catalyst (Cu/SO42−/Al–Ce-PILC) was applied for the SCR of NO by C3H6 in the presence of oxygen. The catalyst 2 wt% Cu/SO42−/Al–Ce-PILC showed good performance over a broad range of temperature, its maximum conversion of NO was 56% at 623 K and remained as high as 22% at 973 K. Furthermore, the presence of 10% water slightly decreased its activity, and this effect was reversible following the removal of water from the feed. Py-IR results showed SO42− modification greatly enhanced the number and strength of Brönsted acidity on the surface of Cu/SO42−/Al–Ce-PILC, which played a vital role in the improvement of NO conversion. TPR and XPS results indicated that both Cu+ and isolated Cu2+ species existed on the optimal catalyst, mainly Cu+, as Cu content increased to 5 wt%, another species CuO aggregates which facilitated the combustion of C3H6 were formed.  相似文献   

9.
The electrodeposition of metallic rhodium on pyrolytic graphite from 10 mM Na3RhCl6 + 0.5 M NaCl aqueous solution was studied by potentiostatic method with the use of a double-pulse technique involving nucleation and growth pulses. Physico-chemical properties of Rh deposits were investigated by electrochemical methods and scanning electron microscopy. The activity of Rh-modified graphite electrodes towards nitrate reduction in neutral medium was demonstrated, the activation energy of nitrate reduction and NO3 Langmuir adsorption constant on Rh deposits were determined.

The use of double-pulse technique resulted in enhanced surface coverage in comparison with usual potentiostatic deposition and in decreasing the mean particle size down to 30 nm, while the specific catalyst surface area attains 32 m2 g−1. The increase in the nucleation pulse duration from 20 to 100 ms enhances the mass catalytic activity towards NO3 reduction, which reaches 175 A g−1 for the best samples. Irrespectively of electrodeposition parameters, only NH3 and NO2 were detected as nitrate reduction products. The rate of NO3 destruction was equal to which is much higher than that of most of Pd/Cu-based nitrate hydrogenation systems and Ag/TiO2 photocatalysts.  相似文献   


10.
Well-known, yet undefined, changes in the conditions and activity of palladized zerovalent iron (Fe/Pd) over an extended period of time hindered a careful study of dechlorination kinetics in long-term experiments. A short-term experimental method was, therefore, developed to study the effects of temperature and solvent on the dechlorination of monochlorobiphenyls (MCBs), 2-chlorobiphenyl (2-ClBP), in particular by Fe/Pd. The experiments started with specified initial conditions and lasted only for 10 min. The average value (k) of the first-order rate constant for the dechlorination of 2-ClBP was 0.13 ± 0.03 L m−2 h−1, not significantly different from the average values for 3-chlorobiphenyl and 4-chlorobiphenyl. The apparent activation energy was 20 ± 4 kJ mol−1 and 17 ± 7 kJ mol−1, in a temperature range between 4 °C and 60 °C, for the dechlorination of 2-ClBP using two batches of Fe/Pd catalyst. The k values decreased significantly in mixtures with a methanol concentration higher than 10%. The values of the rate constant were slightly influenced by the initial concentrations in the experiments at a low temperature and in a solution with a high methanol concentration. The concentration dependence was described with a Langmuir equation, based on the Langmuir–Hinshelwood mechanism that includes an adsorption step of a single species preceding a rate-determining catalytic reaction.  相似文献   

11.
Pdn+/Cen+/Na+/γ-Al2O3-type materials used as FCC additives for CO/NOx control were characterized by extended X-ray absorption fine structure (EXAFS) spectroscopy and in situ FTIR. The EXAFS data indicate that in freshly prepared samples palladium is present in the form of highly dispersed PdO species. Reduction with H2 at 500 °C leads to the formation of small Pd clusters incorporating on average approximately six to eight metal atoms at a Pd−Pd bond distance of 2.76 Å. All components of these materials can interact with NO and promote the formation of nitrate/nitrite species, essentially “trapping” NOx species on the catalyst surface. However, the Na+ species dominate the surface chemistry and readily form sodium nitrates with a characteristic IR band at 1370–1385 cm−1. Finally, hydroxyls from the support are also actively participating in the formation of HNOx type compounds with characteristic stretching vibrations in the 3500–3572 cm−1 region.  相似文献   

12.
A silica composite of a perfluorocarbonsulfonic acid resin, Aciplex, has been used as a solid acid catalyst for a variety of reactions concerning water. The Aciplex–SiO2 composite containing 20 wt% Aciplex has a surface area of 1.3 m2 g−1 and possesses an ion-exchanged capacity of 0.46 meq. g−1 after pretreatment at 423 K, which is higher than that of 13 wt% Nafion–SiO2 (0.12 meq. g−1). The acid strengths estimated from an initial heat of adsorption of NH3 were similar for these polymer resin composites. It was found that the Aciplex–SiO2 was more active than typical solid acids such as Cs2.5H0.5PW12O40, H-ZSM-5, and SO42−/ZrO2 for hydrolysis of ethyl acetate in excess water and esterification of acrylic acid with 1-butanol, while it was less active than Cs2.5H0.5PW12O40 for N-alkylation of acrylonitrile with 1-adamantanol and solid–solid hydrolysis of 2-naphthyl acetate. The Aciplex–SiO2 was superior in activity to Nafion–SiO2 for all the above reactions and in thermal stability. These results indicate that Aciplex–SiO2 is a promising solid acid catalyst for reactions involving liquid phase water.  相似文献   

13.
Niobia and alumina supported palladium catalyst promoted by copper were investigated in the reaction of nitrate catalytic reduction in water and characterized by temperature programmed reduction, physisorption, H2 chemisorption and X-ray diffraction. Niobia supported Pd–Cu catalysts were as active and selective as an alumina supported catalyst. All catalysts had similar turnover frequencies independent of the support. The control of pH and the interaction between Pd and Cu were critical to improving the selectivity and activity of Pd–Cu/Nb2O5 catalysts.  相似文献   

14.
Niobium nitride was synthesized on a Si(400) substrate and a γ-alumina pellet using a CVD method with a stream of NbCl5/Ar, NH3, and H2 gases at 723–973 K under reduced pressure. The composition and surface properties of the deposited niobium nitride were analyzed using XRD and XPS measurements. The activity of alumina-supported niobium nitrides for the hydrodesulfurization (HDS) of thiophene at 673 K and atmospheric pressure was determined. The alumina had a surface area of 177 m2 g−1 and the alumina-supported niobium nitride catalyst had surface areas of 179–190 m2 g−1. Although the catalysts had low activity in the initial stages, the activity increased after 200–300 min started to about three times the initial activities. XPS analysis indicated that the activity of the niobium nitride catalysts was decreased by sulfur accumulation on the surface and nitrogen released from niobium nitride. The relationship between the surface properties of the niobium nitride catalysts and the activities for thiophene HDS is discussed.  相似文献   

15.
The reaction mechanism of the reduction of NO by propene over Pd-based catalysts was studied by FTIR spectroscopy. It was observed that the reaction between NO and propene most probably goes via isocyanate (2256–2230 cm−1), nitrate (1310–1250 cm−1) and acetate (1560 and 1460 cm−1) intermediates formation. Other possible intermediates such as partially oxidized hydrocarbons, NO2, and formates were also detected. The reaction between nitrates and acetates or carbonates reduced nitrates to N2 and oxidized carbon compounds to CO2. In situ DRIFT provides quick and rather easily elucidated data from adsorbed compounds and reaction intermediates on the catalyst surface. The activity experiments were carried out to find out the possible reaction mechanism and furthermore the kinetic equation for NO reduction by propene.  相似文献   

16.
In the preparation of 1% Au/TiO2 catalysts supported on either Degussa P-25 or anatase (90 m2 g−1) by deposition–precipitation, the gold content passes through a maximum at about the isoelectric point (pH 6), but maximum specific rates occur at pH 8–9 because the Au particle size becomes smaller as the pH is further increased. The gold uptake increases with the surface area of the support (anatase, rutile, P-25) and is complete above 200 m2 g−1; adsorption of the gold precursor at pH 9 is shown to be equilibrium-limited. Highest activities are found with supports of 50 m2 g−1. Catalysts made with high-area anatase (240 or 305 m2 g−1) are least active but show least deactivation.With Au/SnO2 catalysts, gold uptake does not depend on the area of the support, and is highest at pH 7–8; very active catalysts (T50 = 230–238 K) are obtained using SnO2 of 47 m2 g−1. Storing a catalyst at 258 K for 1 week dramatically improves its stability. Results for Au/CeO2 and Au/ZrO2 catalysts confirm that moderate support areas give the most active catalysts, and suggest that surface area is often more important than chemical composition.  相似文献   

17.
The oxidation of benzene to phenol has been successfully carried out in air over Cu-ZSM-5 at moderate temperatures. Several parameters such as Cu loading, calcination temperature and co-exchanged metal ions influence the nature of the catalyst. At low Cu loadings, the catalyst is more selective to phenol while at high Cu loadings CO2 is the major product. In situ H2-TPR XAFS studies reveal that at low Cu loadings, Cu exists as isolated pentacoordinated ions, with 4 equatorial oxygens at 1.94 Å and a more distant axial oxygen at 2.34 Å. At higher loadings, monomeric as well as dimeric Cu species exist, with a Cu–Cu distance of 2.92 Å. This suggests that the isolated Cu sites are the active sites responsible for phenol formation. When the catalyst was calcined at 450 °C, the activity peaked in the first hour and then slowly deactivated, but when the calcination temperature was increased to 850 °C, the activity slowly increased until it reached a plateau. Analysis of the XANES region during in situ H2-TPR shows that at lower calcination temperatures two reduction peaks are present, corresponding to Cu2+ → Cu+ and Cu+ → Cu0. At high calcination temperatures, only a small fraction of the Cu undergoes the two-step reduction and most of the Cu remains in the +2 state. Slow deactivation of the catalyst calcined at 450 °C is due to migration of the Cu ions to inaccessible sites in the zeolite; at high calcination temperatures the Cu is tightly bound to the framework and is unable to migrate. EXAFS analysis of the sample calcined at 850 °C reveals two Cu–Si(Al) scattering paths at 2.83 Å. Doping the catalyst with other metals, in particular Ag and Pd, further improves the activity and selectivity of the reaction. The addition of water to the reaction increases the selectivity of the reaction by displacing the product from the active site.  相似文献   

18.
A novel well-ordered super-microporous layered material, silica-pillared niobic acid, was synthesized by a guest-exchange route and structurally characterized by powder X-ray diffraction (XRD), infrared absorption spectroscopy (IR), thermogravimetric and differential thermal analysis (TG/DTA), transmission electron micrographs (TEM), nitrogen adsorption method and ammonia-temperature-programmed desorption (NH3-TPD). The obtained silica-pillared layered niobic acid had a supergallery of 1.78 nm, a large BET surface area of 250 m2 g−1, and a high thermal stability exceeding 973 K.The pillared layered material was also found to be an efficient solid acid catalyst for the vapor-phase Beckmann rearrangement of cyclohexanone oxime. When 1-hexanol was fed with cyclohexanone oxime, this solid catalyst exhibited a 100% conversion of the oxime with a selectivity of ε-caprolactam beyond 85% at a reaction temperature of 613 K and a WHSV of 0.17 h−1 in terms of cyclohexanone oxime, and there was no significant change of the conversion and selectivity within 6 h.  相似文献   

19.
G. Piehl  T. Liese  W. Grünert   《Catalysis Today》1999,54(4):333-406
ZSM-5 zeolite was loaded with vanadyl ions (VO2+) by treatment of Na–ZSM-5 with aqueous VOSO4 solution at pH 1.5–2. The catalytic material was tested for the selective catalytic reduction of NO with ammonia at temperatures between 473 and 823 K and normal pressure using a feed of 1000 ppm NO, 1000 (or 1100) ppm NH3 and 2% O2 in He. The catalyst proved to be highly active, providing, e.g. initial NO conversions of >90% at 620 l g−1 h−1 (≈400 000 h−1) and 723 K, and selective, providing nitrogen yields equal to NO conversion at equimolar feed in a wide temperature range and only minor N2O formation at NH3 excess. Admixture of SO2 (200 ppm) resulted in an upward shift of the useful temperature range, but did not affect the catalytic behaviour at temperatures ≥623 K. No SO2 conversion was noted at T ≤ 723 K and 450 l g−1 h−1. The poisoning effect of water (up to 4.5 vol%) was weak at temperatures between 623 and 773 K. VO-ZSM-5 catalysts are gradually deactivated already under dry conditions, probably by oxidation of the vanadyl ions into pentavalent V species. This deactivation is considerably accelerated in the presence of water.  相似文献   

20.
Incipient wetness impregnation of zeolite Y with copper(II) nitrate solution and inert activation at 650 °C led to active catalysts for the oxidative carbonylation of methanol to dimethyl carbonate in the gas phase. Activities were measured under elevated pressure (0.4–1.6 MPa) with feed compositions of CO/MeOH/O2 = 40/20/6–1.5 vol.% (balanced by N2) over zeolite Y loaded with 10–17 wt.% copper. It could be shown that inert activation at 650 °C enhanced the activity, and that Cu loading of 14–17 wt.% gave the best performance. By combined XRD, TEM, TPR and DRIFT characterization it was found that the inert activation initiated dispersion of crystalline CuO, auto-reduction of Cu2+ to Cu+ and redistribution of copper ions with enrichment inside the supercages of the zeolite. The O2 content of the feed was found to control the selectivity to dimethyl carbonate. Dimethyl carbonate selectivities of 70–75% were achieved within the temperature range of 140–170 °C at an O2 content of 1.5 vol.%. This allowed space-time yields of dimethyl carbonate up to 632 g lcat−1 h−1 at methanol conversions of 5–12%. Formation of the main side product, dimethoxymethane, was surprisingly affected by CO, which is not in line with suggested reaction pathways. A mechanism is proposed including formation of surface carbonate structures as common intermediate.  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