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 328 毫秒
1.
Homogeneous physical mixtures containing a commercial Cu/ZnO/Al2O3 catalyst and a solid–acid catalyst were used to examine the acidity effects on dimethyl ether hydrolysis and their subsequent effects on dimethyl ether steam reforming (DME-SR). The acid catalysts used were zeolites Y [Si/Al = 2.5 and 15: denoted Y(Si/Al)], ZSM-5 [Si/Al = 15, 25, 40, and 140: denoted Z(Si/Al)] and other conventional catalyst supports (ZrO2, and γ-Al2O3). The homogeneous physical mixtures contained equal amounts, by volume, of the solid–acid catalyst and the commercial Cu/ZnO/Al2O3 catalyst (BASF K3-110, denoted as K3). The steam reforming of dimethyl ether was carried out in an isothermal packed-bed reactor at ambient pressure.

The most promising physical mixtures for the low-temperature production of hydrogen from DME contained ZSM-5 as the solid–acid catalyst, with hydrogen yields exceeding 90% (T = 275 °C, S/C = 1.5, τ = 1.0 s and P = 0.78 atm) and hydrogen selectivities exceeding 94%, comparable to those observed for methanol steam reforming (MeOH-SR) over BASF K3-110, with values equaling 95% and 99%, respectively (T = 225 °C, S/C = 1.0, τ = 1.0 s and P = 0.78 atm). Large production rates of hydrogen were directly related to the type of acid catalyst used. The hydrogen production activity trend as a function of physical mixture was

  相似文献   

2.
Steam reforming(SR) of dimethyl ether(DME) was investigated for the production of hydrogen for fuel cells.The activity of a series of solid acids for DME hydrolysis was investigated.The solid acid catalysts were ZSM-5 [Si/Al=25,38 and 50:denoted Z(Si/Al)] and acidic alumina(γ-Al2O3) with an acid strength order that was Z(25)>Z(38)>Z(50)>γ-Al2O3.Stronger acidity gave higher DME hydrolysis conversion.Physical mixtures containing a CuO-ZnO-Al2O3-ZrO2 catalyst and solid acid catalyst to couple DME hydrolysis and methanol SR were used to examine the acidity effects on DME SR.DME SR activity strongly depended on the activity for DME hydrolysis.Z(25) was the best solid acid catalyst for DME SR and gave a DME conversion>90% [T=24℃,n(H2O)/n(DME)=3.5,space velocity=1179 ml·(g cat)-1·h-1,and P=0.1MPa].The influences of the reaction temperature,space velocity and feed molar ratio were studied.Hydrogen production significantly depended on temperature and space velocity.A bifunctional catalyst of CuO-ZnO-Al2O3-ZrO2 catalyst and ZSM-5 gave a high H2 production rate and CO2 selectivity.  相似文献   

3.
The electro-oxidation of dimethyl ether (DME) on PtMe/Cs (Me = Ru, Sn, Mo, Cr, Ni, Co, and W) and Pt/C electro-catalysts were investigated in an aqueous half-cell, and compared to the methanol oxidation. The addition of a second metal enhanced the tolerance of Pt to the poisonous species during the DME oxidation reaction (DOR). The PtRu/C electro-catalyst showed the best electro-catalytic activity and the highest tolerance to the poisonous species in the low over-potential range (<0.55 V, 50 °C) among the binary electro-catalysts and the Pt/C, but at the higher potential (>ca. 0.55 V, 50 °C), the Pt/C behaved better than PtRu/C. The apparent activation energy for the DOR decreased in the order: PtRu/C (57 kJ mol−1) > Pt3Sn/C (48 kJ mol−1) ≈ Pt/C (46 kJ mol−1). On the other hand, the activation energy for the MOR showed a different turn, decreased in the following order: Pt/C (43 kJ mol−1) > Pt3Sn/C (35 kJ mol−1) ≈ PtRu/C (34 kJ mol−1). The temperature dependence of the DOR was greater than that of the oxidation of methanol (MOR) on the PtRu/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.
A series of 1 wt.%Pt/xBa/Support (Support = Al2O3, SiO2, Al2O3-5.5 wt.%SiO2 and Ce0.7Zr0.3O2, x = 5–30 wt.% BaO) catalysts was investigated regarding the influence of the support oxide on Ba properties for the rapid NOx trapping (100 s). Catalysts were treated at 700 °C under wet oxidizing atmosphere. The nature of the support oxide and the Ba loading influenced the Pt–Ba proximity, the Ba dispersion and then the surface basicity of the catalysts estimated by CO2-TPD. At high temperature (400 °C) in the absence of CO2 and H2O, the NOx storage capacity increased with the catalyst basicity: Pt/20Ba/Si < Pt/20Ba/Al5.5Si < Pt/10Ba/Al < Pt/5Ba/CeZr < Pt/30Ba/Al5.5Si < Pt/20Ba/Al < Pt/10BaCeZr. Addition of CO2 decreased catalyst performances. The inhibiting effect of CO2 on the NOx uptake increased generally with both the catalyst basicity and the storage temperature. Water negatively affected the NOx storage capacity, this effect being higher on alumina containing catalysts than on ceria–zirconia samples. When both CO2 and H2O were present in the inlet gas, a cumulative effect was observed at low temperatures (200 °C and 300 °C) whereas mainly CO2 was responsible for the loss of NOx storage capacity at 400 °C. Finally, under realistic conditions (H2O and CO2) the Pt/20Ba/Al5.5Si catalyst showed the best performances for the rapid NOx uptake in the 200–400 °C temperature range. It resulted mainly from: (i) enhanced dispersions of platinum and barium on the alumina–silica support, (ii) a high Pt–Ba proximity and (iii) a low basicity of the catalyst which limits the CO2 competition for the storage sites.  相似文献   

6.
A series of Al-HMS with different Si/Al ratio was used as a solid acid catalyst for methanol dehydration to dimethyl ether (DME). The effect of temperature, feed composition, space velocity, and the catalyst Si/Al ratio on the catalytic dehydration of methanol was investigated. By decreasing Si/Al, the temperature required to reach equilibrium conversion of methanol decreased due to the increased number of acidic sites. Compared to commercial γ-Al2O3, Al-HMS-5 and Al-HMS-10, catalysts exhibited a high yield of DME. Among all Al-HMS catalysts, Al-HMS-10 exhibited an optimum yield of 89% with 100% selectivity and excellent stability for methanol dehydration to DME.  相似文献   

7.
An extensive series of 30 Cu exchanged zeolites and Cu impregnated silicas and aluminas have been tested in their capacities to stabilize the bis(μ-oxo)dicopper core. This core shows a remarkably activity towards methane, as it selectively hydroxylates methane into methanol at the low temperature of 125 °C. UV–vis spectroscopy is an easy approach to detect the presence of this bis(μ-oxo)dicopper core since it is characterized by an intense charge transfer band at 22 700 cm−1. In this way it was found that after calcination, only the Cu exchanged zeolites ZSM-5 and MOR are capable of stabilizing this core. In addition, an optimum in the Si/Al ratio and in the calcination temperature were observed, indicating that this core requires a rather specific coordination environment. For ZSM-5, the optimal Si/Al ratio for bis(μ-oxo) dicopper core formation is between 12 and 30 and the amount of this core increases with increasing copper loading above Cu/Al = 0.2. Calcination in O2 should be done at temperatures higher than 280 °C and lower than 700 °C. After reaction with methane at low temperature (150 °C), it was found that only Cu-ZSM-5 and Cu-MOR yielded methanol, whereas all the other Cu based materials yielded almost no methanol. At higher temperatures (200 °C) however, Cu-FER and Cu-BEA showed comparable methanol yields as Cu-ZSM-5 and also the methanol yield of Cu-MOR increased at this higher reaction temperature, indicating that a second not yet identified Cu-oxygen species is activated in the FER, BEA and MOR zeolites at higher temperatures.  相似文献   

8.
This study deals with emission quenching of zeolite encapsulated trisbipyridyl ruthenium (II) (Ru(bpy)32+) by oxygen. Oxygen saturated solutions of Ru(bpy)32+ typically show about 70% quenching (I0/I=3.3), where I0 and I are the peak intensities of the emission in N2 and O2, respectively. However, an aqueous suspension of Ru(bpy)32+-zeolite Na–Y (Si/Al = 2.5) (abbreviated as Ru–Na–Y) showed no quenching at all. This observation motivated us to analyze how the transport of O2 is occurring in the zeolite. Upon exposure of solid Ru–Na–Y (99% of intrazeolitic water) to N2/O2 dry gases, quenching in oxygen was found to be 5% (I0/I=1.07). Partial dehydration at room temperature with loss of 33% of the water molecules from the zeolite led to 66% (I0/I=2.96) quenching. Dehydration of Ru–Na–Y at 250 °C under vacuum overnight led to complete loss of intrazeolitic water and increased quenching to 90% (I0/I=10.7). Nanocrystalline Ru(bpy)32+-zeolite Y upon vacuum dehydration lost 55% of the intrazeolitic water and showed 96% (I0/I=25.3) quenching. The extent of quenching of Ru(bpy)32+ in zeolites by O2 is by far the largest as compared to previously studied matrices, and is being attributed to confinement of O2 in the supercages, which leads to increase in number of collisions with Ru(bpy)32+ and enhanced quenching. However, these samples showed complete lack of sensitivity (I0/I=1) to oxygen upon exposure to water saturated gas or dissolved gas. Dealumination of zeolite framework by treatment with (NH4)2SiF6 produced a framework of Si/Al = 9.5, and with SiCl4 a framework of Si/Al > 100. With increasing dealumination, the extent of quenching by dissolved O2 increased.  相似文献   

9.
The influence of various concentrations of NaClO4, as a pitting corrosion agent, on the corrosion behaviour of pure Al, and two Al–Cu alloys, namely (Al + 2.5 wt% Cu) and (Al + 7 wt% Cu) alloys in 1.0 M Na2SO4 solution was investigated by potentiodynamic polarization and potentiostatic techniques at 25 °C. Measurements were conducted under the influence of various experimental conditions, complemented by ex situ energy dispersive X-ray (EDX) and scanning electron microscopy (SEM) examinations of the electrode surface. In free perchlorate sulphate solutions, for the three Al samples, the anodic polarization exhibits an active/passive transition. The active dissolution region involves an anodic peak (peak A) which is assigned to the formation of Al2O3 passive film on the electrode surface. The passive region extends up to 1500 mV with almost constant current density (jpass) without exhibiting a critical breakdown potential or showing any evidence of pitting attack. For the three Al samples, addition of ClO4 ions to the sulphate solution stimulates their active anodic dissolution and tends to induce pitting corrosion within the oxide passive region. Pitting corrosion was confirmed by SEM examination of the electrode surface. The pitting potential decreases with increasing ClO4 ion concentration indicating a decrease in pitting corrosion resistance. The susceptibility of the three Al samples towards pitting corrosion decreases in the order: Al > (Al + 2.5 wt% Cu) alloy > (Al + 7 wt% Cu) alloy. Potentiostatic measurements showed that the rate of pitting initiation increases with increasing ClO4 ion concentration and applied step anodic potential, while it decreases with increasing %Cu in the Al samples. The inhibitive effect of SO42− ions was also discussed.  相似文献   

10.
Periodic mesostructured organosilicas (PMO) were first synthesized using 1,2-bis(triethoxysilyl)ethylene (BTENE) under acidic conditions using Pluronic 123 as surfactant. The ethylene bridges were then arylated with benzene using AlCl3 as catalyst. These materials were further treated with sulfuric acid for the sulfonation of the phenyl moieties yielding a new preparation of sulfonic acid functionalized PMO. Ordered hexagonal mesostructures with surface areas up to 440 m2/g and narrow pore size distribution (around 5.3 nm) were obtained. This work thus provides a new example of chemical modification for the conception of functionalized PMO catalysts. Liquid phase self-condensation of heptanal was performed at 75 °C in the presence of these catalysts and the results were compared with those obtained with several other heterogeneous acid catalysts.  相似文献   

11.
A method for the sulfonation of PEEK-WC, a glassy poly(ether ether ketone) with sulphuric acid is presented. Depending on the reaction time, polymers with ion exchange capacity (IEC) from 0.30 to 0.76 meqH+/g are obtained, as determined by titration with NaOH solutions. The thermal properties of the polymers were studied by differential scanning calorimetry, showing that the glass transition temperature increases with increasing degree of sulfonation, from 224 °C for pure PEEK-WC to 246 °C for the polymer having an IEC of 0.76 meqH+/g. The sulfonated polymers were used to prepare proton exchange membranes for possible application in fuel cells. Dense membranes were prepared by solvent evaporation, using DMA as the solvent. The transport properties of the membranes were determined in terms of water uptake and permeability for hydrogen and oxygen. Electrochemical characterization was performed by measuring cell voltage and power density curves as a function of current density at different working temperatures and the results were compared with those of a commercial Nafion membrane. A power density of 284 mW/cm2 was obtained for S-PEEK-WC membrane at 120 °C in H2/air fuel cell, slightly above the corresponding value found for Nafion.  相似文献   

12.
二甲醚蒸气重整催化制氢工艺   总被引:4,自引:2,他引:2  
考察了HZSM-5型分子筛在不同温度下对二甲醚水解的活性与稳定性,分析了分子筛硅铝比的影响。还将HZSM-5-50(硅铝比为50)与自制Cu-Zn-Al(摩尔比为3∶5∶2)催化剂进行物理混合制成复合催化剂,研究了两者质量配比对复合催化剂上二甲醚蒸气重整制氢过程中转化率与氢选择性的影响,并考察了不同原料气体空速下复合催化剂的性能。结果表明,复合催化剂质量比Cu-Zn-Al∶HZSM-5-50为1∶1、温度为275 ℃、气体空速为18000 mL/(gcat·h)的反应条件下性能比较稳定,当空速接近4000 mL/(gcat·h)时,H2收率可获极值,达到54%左右。  相似文献   

13.
This study is focused on the adsorption of a chlorinated volatile organic compound, the tetrachloroethylene (PCE), on dealuminated faujasite type zeolites with framework Si/Al ratio between 5 and 100. PCE dynamic adsorption experiments with and without water vapour (relative humidity of, respectively, 50% and 0%) were carried out in a fixed bed reactor at 50 °C. Breakthrough curves were fitted by a model using the integral of a Gauss distribution. PCE adsorption capacities depend on the adsorbent microporous volume. However, in presence of water vapour, PCE adsorption is favoured on hydrophobic zeolites but also depends on the diffusional limitations inside the porous system. In order to have a better understanding of water molecules adsorption, isotherms were measured using thermogravimetric method at 25 °C. The presence of water vapour generally decreases PCE uptake but its influence decreases as the Si/Al ratio of the adsorbent increases. Experiments with various gases hourly space velocity (GHSV) and inlet PCE concentrations were also performed. PCE complete desorption was obtained on HFAU(Si/Al = 17) at 180 °C. This easy regeneration of the sample permitted adsorption/regeneration cycles maintaining good adsorption properties.  相似文献   

14.
The single gas H2 and N2 permeability of a 4 μm thick dense fcc-Pd66Cu34 layer has been studied between room temperature and 510 °C and at pressure differences up to 400 kPa. Above 50 °C the H2 flux exhibits an Arrhenius-type temperature dependence with JH2=(5.2±0.3) mol m−2 s−1 exp[(−21.3 ± 0.2) kJ mol−1/(R·T)]. The hydrogen transport rate is controlled by the bulk diffusion although the pressure dependence of the H2 flux deviates slightly from Sieverts’ law. A sudden increase of the H2 flux below 50 °C is attributed to embrittlement.  相似文献   

15.
8-Hydroxyquinoline-5-sulphonate/Al(III) aqueous solutions were studied both by potentiometric titrations and voltammetric measurements, in order to obtain the number, the stoichiometry and the stability constants of the complexes formed at equilibrium, and to evaluate the redox and (electro)kinetic properties of the free ligand and of the metal/ligand complexes. The complexes formed in 0.2 m (Na)Cl aqueous solution (stability log beta values ± standard deviation) are AlL+ (8.95 ± 0.05), AlL2 (17.43 ± 0.03) and AlL33− (24.58 ± 0.05), where “L” denotes the free ligand in the completely deprotonated form (L2−, pKa1 = 3.910 ± 0.008, pKa2 = 8.319 ± 0.004). AlL33− is the predominant Al(III) species in a very wide range of pH, metal and ligand concentrations and metal-to-ligand ratios. The free ligand shows an oxidation wave at 0.62 V versus SCE. The proposed oxidation mechanism includes a first reversible one-electron oxidation of the ligand, followed by a coupling reaction and by a second reversible one-electron oxidation, and finally by a decomposition reaction. The addition of Al(III) lowers the intensity of the oxidation wave due to the formation of the redox-inactive complex AlL33−. A residual low signal was attributed to the free ligand produced by the complex dissociation, AlL33− = AlL2 + L2−. All the kinetic parameters involved in the ligand oxidation and in the complex disruption were calculated on the basis of the agreement between experimental and simulated linear sweep and cyclic voltammetries. Correctness of the mechanisms proposed was further confirmed “a posteriori” by the agreement between potentiometric and linear sweep voltammetric results. The low residual signal observed in the presence of fully formed complex was attributed to the free ligand produced by the complex dissociation, having a kinetic constant estimated 0.2 s−1.  相似文献   

16.
Colored products have been obtained by thermal treatment of erionite mixed with elemental sulfur and alkalis (Na2CO3 or K2CO3). Synthesis at lower temperature (500 °C) resulted in forming colored materials with preserved ERI structure (particularly with potassium cations), whereas the treatment of highly alkaline mixtures at high temperature (800 °C) caused re-crystallization to SOD (when Na was used) or to unknown structures (with K). The ESR spectra of the products with preserved original ERI structure recorded at room temperature show two signals, one with g = 2.030 and another with g tensor value either 2.045 (for sodium containing samples) or 2.005 (for the potassium bearing products). The latter signal could be considered as a reflection of radicals immobilized in small ε-cages, but it is more likely that it presents anisotropic component of the radical spectrum. The spectra measured at 77 K show anisotropy indicating three values of orthorhombic g-tensor, which is typical of radical, although the values vary for particular samples.  相似文献   

17.
A heterogeneous sono-catalytic system with addition of hydrogen peroxide (USH2O2+Cat.) was employed for the degradation of 200 ppm of p-chlorophenol (4-CP) at 25 °C and 100 W of ultrasound power. One thousand and six hundred parts per million of initial hydrogen peroxide (H2O2) concentration and 1 g/L of catalyst loading over three heterogeneous copper catalysts, CuO, Cu/Al2O3 (Cu/Al) and CuO·ZnO/Al2O3 (Cu/Zn) was used. The benefits of ultrasound in a heterogeneous catalytic system were evaluated. A considerable synergistic effect of the USH2O2+Cat. system was only achieved with supported catalysts (Cu/Al and Cu/Zn) possibly due to good dispersion of catalysts as a result of catalyst size reduction during ultrasound irradiation. Moreover, between the two supported copper catalysts, the Cu/Al provided promising catalytic performance by giving higher 4-CP and TOC removal accompanied with efficient H2O2 consumption. Experiments with a homogeneous copper catalyst revealed that use of ultrasound in a homogeneous system shows an adverse effect on decomposition of 4-CP.  相似文献   

18.
The fracture toughness, Gc, of the interface between a nitrogen plasma-treated poly(ethylene terephthalate) (PET) film and a poly(styrene-co-maleic anhydride) (PSMA) substrate was measured by using asymmetric double cantilever beam method. The effects of plasma treatment condition on PET films and post-plasma bonding treatment of the bi-material on the adhesion and the failure mechanism were investigated. For a given plasma pressure and energy, the amount of incorporated nitrogen on the PET surface as determined from X-ray photoelectron spectrometry (XPS) increased with increasing plasma treatment time and reached a plateau value of 7.7 at.%. XPS measurement showed that the incorporated nitrogen was primarily in the form of amine and amide. For bonding temperatures between 130 °C and 160 °C, the fracture toughness increased with increasing nitrogen incorporation on PET surface and reached a saturation Gc which significantly depended on the bonding temperature. The saturation Gc increased from 10 J/m2 at 130 °C to 40 J/m2 at 140 °C, reached a maximum of 120 J/m2 at 150 °C, and then decreased to 60 J/m2 at 160 °C. The location of failure also changed drastically with the bonding temperature. SEM and XPS measurements showed that for bonding temperature < 140 °C, failure occurred at the PET/PSMA interface. For bonding temperature = 150 °C, the interfacial adhesion exceeded that of the cohesive strength of PET film and failure occurred within the PET film. At the bonding temperature of 160 °C, failure occurred within PSMA bulk material. XPS measurement was used to measure the areal joint density, Σcross of PSMA chains pinned on the functionalized PET film surface. A transition in areal joint density below which Gc scales linear with Σcross and above which Gc scales with was found. The transition was identified as the transition from the pure chain scission of in situ formed copolymers to plastic deformation of the interface.  相似文献   

19.
Three zirconia-supported platinum group metal (Pt, Ru and Pt–Ru) catalysts were prepared by impregnation. The activity of these catalysts toward the oxidative steam reforming of ethanol (OSRE) was examined in a fixed-bed reactor in the temperature range of 260–380 °C. The catalysts were characterized by X-ray diffraction (XRD), temperature-programmed reduction (TPR), transmission electron microscopy (TEM) and nitrogen adsorption at −196 °C. Activity results indicated that the optimized experimental conditions involved a reforming temperature of close to 300 °C and the molar ratios of O2/EtOH and H2O/EtOH of 0.44 and 4.9, respectively. An ethanol conversion (CEtOH) approaching 100% and a hydrogen yield (YH2) exceeding 3.0 mole/mole ethanol were noticed at 280 °C over all the catalysts. Among these catalysts, the Pt–Ru/ZrO2 catalyst was an excellent OSRE catalyst at low temperature. The maximum YH2 was 4.4 and the CO distribution was 3.3 mol% at 340 °C.  相似文献   

20.
Ti3SiC2 bulk materials were synthesized from the starting powders of 1Ti/1Si/2TiC–xAl and 3Ti/1SiC/1C–xAl (molar ratios, x ranges from 0.05 to 0.15) at temperatures between 1100 and 1400 °C for 15 min by pulse discharge sintering technique. X-ray diffraction and scanning electron microscopy were used to characterize the synthesized materials. It was found that the addition of Al decreases the content of TiC in the sintered samples and expands the optimal temperature range for the synthesis of Ti3SiC2 bulk materials. By addition of Al, Ti3SiC2 bulk materials of high phase-purity have been synthesized at 1100 and 1200 °C from 1Ti/1Si/2TiC and 3Ti/1SiC/1C starting powders, respectively.  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