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 new approach to kinetic studies of fat hydrogenation is discussed. An experimental setup is described in detail. An example of reactor performance in hydrogenation of fatty acid methyl esters is given. aPresent address: AB Karlshamns Oljefabriker, S-292 00 Karlshamn, Sweden. Notation: c, outlet concentration, mol/m3 ; co, inlet concentration, mol/m3: cb, concentration in bulk fluid, mol/m3 ; cs concentration at catalyst surface, mol/m3 ; d,pore diameter, m; De, effective diffusivity, m2/s; E, activation energy of reaction, J/mol; h, heat transfer coefficient, J/m2 s K; ΔH, heat of reaction, J/mol; kc, mass transfer coefficient, m/s; p, partial pressure, Pa; po, saturated vapor pressure, Pa; qf, total flow to reactor, m3/s; qrec, recycle flow, m3/s; R, observed reaction rate per unit particle volume, mol/s m3 ; Rg, gas constant, J/mol K; r, observed reaction rate per unit mass of catalyst, mol/kg s; rp particle radius, m; Tb, temperature of gas in bulk flow, K; Ts, temperature of gas at catalyst surface, K;-v, molar volume, m3/mol; V, volume of catalyst bed, m3 ; W, catalyst mass, kg. Greek symbols: σ, surface tension, N/m; λe, effective thermal conductivity of catalyst particle, J/m s K.  相似文献   

2.
《Catalysis communications》2001,2(3-4):135-138
The hydrolysis of COS using γ-Al2O3 as catalyst at 30°C is described. The effect of additives (ca. 3% Na, Fe, Co, Ni, Cu, Zn) is described in detail. The additives reduce the surface area of the γ-Al2O3 markedly but have no effect on the powder X-ray diffraction pattern. All significantly enhance the initial intrinsic activity (mol COS hydrolysed/m2/h) when compared with unpromoted γ-Al2O3. However, this initial performance rapidly decays for the Na-, Fe-, Co- and Cu-modified catalysts. During this initial reaction period, of up to 5 h, the catalyst retains sulphur. In terms of specific activity (mol COS hydrolysed/g/h) only Ni- and Zn-promoted catalysts give a significant increase that is stable throughout the time-scale of these experiments.  相似文献   

3.
A 4-step kinetic model of CO2-assisted oxidative dehydrogenation (ODH) of propane to C2/C3 olefins over a novel MoOx/La2O3–γAl2O3 catalyst was developed. Kinetic experiments were conducted in a CREC Riser Simulator at various reaction temperatures (525–600 °C) and times (15–30 s). The catalyst was highly selective towards propylene at all combinations of the reaction conditions. Langmuir-Hinshelwood type kinetics were formulated considering propane ODH, uni- and bimolecular cracking of propane to produce a C1-C2 species. It was found that the one site type model adequately fitted the experimental data. The activation energy for the formation of propylene (67.8 kJ/mol) is much lower than that of bimolecular conversion of propane to ethane and ethylene (303 kJ/mol) as well as the direct cracking of propane to methane and ethylene (106.7 kJ/mol). The kinetic modeling revealed the positive effects of CO2 towards enhancing the propylene selectivity over the catalyst.  相似文献   

4.
Ni‐Al2O3 catalyst activity was tested for methane steam reforming using two different reaction systems: a catalyst particle bed (0.42–0.5 mm catalyst particles diluted in SiC) with a surface area‐to‐volume ratio SA/V of 910 m–1 and a porosity ? of 52 % and a catalyst‐coated metal monolith with an SA/V of 3300 m–1 and an ? of 86 %. Under a steam‐to‐carbon ratio of 2.5 and at a temperature of 700 °C, the highest specific reaction rates were found for the catalyst‐coated monolith. The high SA/V and ?, together with the high rate of heat transfer of the metal monolith were found to be responsible of this optimum behavior. However, in both systems, the Ni‐Al2O3 catalyst suffered a catalyst deactivation during operation.  相似文献   

5.
The reaction-extraction-regeneration system for the liquid-phase oxidation of benzene to phenol in the benzene-water-oxygen system was investigated. Phenol was extracted in the extractor to reduce the concentration of phenol in the benzene phase. As vanadium catalyst was oxidized to inactive species after the oxidation reaction, the regenerator was installed in the system to reduce the oxidation state of vanadium catalyst from V4+ or VO2+ to the active V3+ under H2 flow. The effects of various operating parameters including concentration of VCl3 catalyst, O2 and H2 flow rates, benzene bubble size, pH, surface area of Pt regeneration catalyst, the metal species, and amount of ascorbic acid were investigated. Ascorbic acid was employed as a reducing agent for helping reduce the V4+ form to the active form and therefore improving the activity of vanadium catalyst. VCl3 catalyst concentration of 10 mol/m3 with pH of 3–4 in the reactor and Pt surface area of 0.05 m2 in the regenerator showed optimal conditions for the system.  相似文献   

6.
New catalysts for olefin metathesis are obtained upon interaction between rhenium trisacetylacetonate and -Al2O3 surface. After high temperature treatment in flows of O2 and N2 and addition of organometallic compound as a cocatalyst the activity of the resulting catalyst exceeds that of known Re/Al2O3 catalysts prepared by impregnation. The catalysts exhibit maximal activity at a cocatalyst surface concentration of 3 × 10–7 mol/m2. Further increase of the cocatalyst concentration leads to deactivation.  相似文献   

7.
Nanocrystalline LaPxOy with various starting P to La ratios from 0.5 to 2.0 catalysts were prepared by a sol?Cgel method using cetyl trimethylammonium bromide (CTAB) as template. The catalysts were thoroughly characterized by N2 physisorption, powder X-ray diffraction (XRD), temperature programmed desorption (TPD) of NH3, solid state 31P and 1H nuclear magnetic resonance (NMR), and transmission electron microscopy (TEM) techniques. XRD results indicate the presence of predominantly monazite LaPO4 with minor amounts of (??3.0 wt%) rhabdophane LaPO4 phase in the samples with starting P/La ratios of 1.0 and 1.5. NH3-TPD results show an increasing trend in the total acidity with increase in P/La ratio. These catalysts were tested in the selective ethanol dehydration in the temperature range between 250 and 400?°C. The catalyst activity (??mol/h/m2) is increased with P/La ratio and the catalyst with highest P/La ratio of 2.0 exhibiting the highest ethanol dehydration activity. The ethanol conversion increased with reaction temperature, reaching 100% at 350?°C and remains unchanged at higher temperatures. On the other hand, the ethylene selectivity is also increased up to 350?°C and then decreased with further increase of reaction temperature. At a P/La ratio of 2, the CTAB templated LaPxOy catalyst showed higher catalytic activities compared to the LaPxOy by hydrothermal method without any template.  相似文献   

8.
The effect of catalyst pore size has been studied for the hydroliquefaction of a West Virginia coal in the presence of Co/Mo/Al2O3 catalyst. The alumina supports used for catalyst preparation had relatively sharp, unimodal pore size distribution with average pore diameters in the range of 100 Å to almost 1000 Å. Loading of MoO3 and CoO on the Al2O3 supports was in the constant weight ratio of 5:1, but the absolute loading was in direct proportion to the surface area of the support. Two series of catalyst were studied: “High loading”, with 9.7 × 10?4 g MoO3/m2 Al2O3, and “low loading”, with 4.5 × 10?4 g MoO3/m2 Al2O3; both loadings were less than the amount necessary for monolayer distribution of MoO3 on Al2O3. The weight of catalyst charged in each autoclave run was varied so that the same weight of MoO3 and CoO was present for each experiment.The principal results were: (1) Al2O3 alone is not catalytic, even in large amount; (2) conversion of coal increases as catalyst pore diameter increases; from 100 Å to at least 500 Å; (3) the increased conversion with increasing pore size is manifested mainly as increased yield of asphaltenes at 400°C, so the ratio of oil to oil-plus-asphaltenes decreases as pore diameter increases; and (4) catalysts with “low loading” of MoO3 and CoO on the Al2O3 surface give higher liquefactions than their counterparts with “high loading”. Most of the results are consistent with an expected low diffusion rate of large, coal-derived molecules through the catalyst pore system. The higher liquefaction with “low loading” of the Al2O3 surface might result from slow desorption of large product molecules (asphaltenes) exhibiting multiple-site adsorption to Mo neighbors on the surface.  相似文献   

9.
A mechanism is developed for the initiated nonbranched-chain formation of ethylene glycol in methanol-formaldehyde solutions at formaldehyde concentrations of 0.1–3.1 mol dm?3 and temperatures of 373–473 K. At a formaldehyde concentration of 1.4 mol dm?3 and T = 473 K, the radiation-chemical yield of ethylene glycol is 139 molecules per 100 eV. The effective activation energy of ethylene glycol formation is 25 ± 3 kJ mol?1. The quasi-steady-state treatment of the reaction network suggested here led to a rate equation accounting for the nonmonotonic dependence of the ethylene glycol formation rate on the concentration of the free (unsolvated) form of dissolved formaldehyde. It is demonstrated that the peak in this dependence is due to the competition between methanol and CH2=O for reacting with the adduct radical HOCH2CH2O?.  相似文献   

10.
Density functional calculations at the B3LYP/6-311G**//B3LYP/6-311G level of theory were used to study the initial process of ethylene degradation on the TiO2 photocatalyst by adopting the dimeric titanium structure Ti2O6H4 as a model of the catalyst surface. Adsorption energies of water and ethylene were calculated to be 31.9 and 20.4 kcal mol–1. The photogenerated OH radical does not desorb from the catalyst surface and the further reaction with ethylene proceeds since the adsorption energy was estimated to be 39.3 kcal mol–1. Our calculation also indicated that under steady illumination, ethylene directly attacks the OH radical bound to the TiO2 surface even though the surface has vacant sites available for ethylene adsorption.  相似文献   

11.
《Ceramics International》2021,47(19):27201-27209
The Cobalt ferrite (CoFe2O4) and cobalt ferrite incorporated with Cu (Cu–CoFe2O4) and Mg (Mg–CoFe2O4) have been synthesized by wet chemical route. In fixed-bed reactor, the CoFe2O4, Cu–CoFe2O4 and Mg–CoFe2O4 were used as catalysts for direct cracking of methane at temperature of 800 °C and 20 mL/min of feed gas flow rate for hydrogen and nano-carbon production. Different characterizations techniques namely XRD, FESEM, XPS, Raman spectroscopy, TGA, BET for fresh and spent catalysts have been executed. The X-ray powder diffraction and Raman spectra confirmed that the fresh catalysts possess a cubic spinel crystal structure. The FESEM images for spent catalysts displayed that filamentous carbons were not formed over catalysts surfaces except a few amount was observed over the spent CoFe2O4 catalyst. XPS results confirmed the purity of the synthesized catalysts and qualitatively evaluate the cation distribution between the tetrahedral and octahedral sites from Co 2p3/2 and Fe 2p3/2 spectra. BET surface area revealed no significant effects in surface area and pore size of CoFe2O4 catalyst by incorporation of Mg and Cu metals. Activity studies showed that incorporation Mg metal on CoFe2O4 improved methane conversion up to 40.03% and hydrogen formation rate of 79.90 mol H2 g−1 min−1 as compared to 31.61% and 66 mol H2 g−1 min−1 for CoFe2O4 catalyst. Whilst, Cu metal incorporation in CoFe2O4 catalyst led to decline the catalyst performance. Structure activity relationship was described in details.  相似文献   

12.
Prussian Blue-modified graphite electrodes (G/PB) with electrocatalytic activity toward H2O2 reduction were obtained by PB potentiostatic electrodeposition from a mixture containing 2.5 mm FeCl3 + 2.5 mm K3[Fe(CN)6] + 0.1 m KCl + 0.1 m HCl. From cyclic voltammetric measurements, performed in KCl aqueous solutions of different concentrations (5 × 10−2–1 m), the rate constant for the heterogeneous electron transfer (k s) was estimated by using the Laviron treatment. The highest ks value (10.7 s−1) was found for 1 m KCl solution. The differences between the electrochemical parameters of the voltammetric response, as well as the shift of the formal potential, observed in the presence of Cl and NO3 compared to those observed in the presence of SO42− ions, points to the involvement of anions in the redox reactions of PB. The G/PB electrodes showed a good electrochemical stability proved by a low deactivation rate constant (0.8 × 10−12 mol cm2 s−1). The electrocatalytic efficiency, estimated as the ratio , was found to be 3.6 (at an applied potential of 0 mV vs. SCE; Γ = 5 × 10−8 mol cm−2) for a H2O2 concentration of 5 mm, thus indicating G/PB electrodes as possible H2O2 sensors.  相似文献   

13.
The carbonylation of methanol with CO using CH3O exchange resin as a heterogeneous catalyst at temperatures near 350 K is examined systematically in an attempt to derive kinetic rate expressions for the reaction. The activation energies for the carbonylation and decarbonylation reactions are found to be 68 kJ/mol and 105 kJ/mol, respectively. The CH3O exchange resin is also shown to suffer no degradation of catalytic activity upon repeated separation and re-use at 353 K.  相似文献   

14.
The results from investigating the influence of temperature, concentration, and flow rate on the catalytic oxidation of vapors of volatile organic compounds (VOCs) in the presence of Pd/γ-Al2O3 catalyst on cellular supports are presented. The activity of Pd/γ-Al2O3 catalysts on ceramic and metal monolith supports with a cellular structure during the catalytic neutralization of VOC (ethanol, ethyl acetate) vapors under laboratory conditions was determined, and the most stable catalyst for the preliminary study of a large batch was chosen. A pilot unit was created to test a large batch of cellular monolith catalyst in neutralizing VOC vapors under conditions of flexographic production. It was established that a high rate of conversion (> 99 %) was achieved for VOC concentrations of 0.5 g/m3 at space velocities of up to ∼104 h−1, and for VOC concentrations of 5.0 g/m3 at space velocities of up to ∼5 × 105 h−1. The change in the activity of the catalysts on metal (nickel alloyed by aluminum) and ceramic cellular supports in service was investigated. After 300–500 min of operation, virtually complete deactivation of catalyst on a metal support was observed, accompanied by the formation of nickel oxide and acetate. Pilot unit tests with catalyst on cellular supports having a volume of 14.5 l in neutralizing the ventilation exhausts of flexographic production confirmed the possibility of more than 90% conversion at VOC concentrations of ∼0.1 g/m3 and more than 97% at VOC concentrations of over 1 g/m3. A consistently high conversion of VOC was observed during a 100 h test of the pilot unit. A system for recovering the heat released during VOC oxidation lowers the operating costs of the pilot unit.  相似文献   

15.
Biodiesel has been identified as a suitable resource that can be produced from biomass such as Styrax confusus Hemsl. In the current study, biodiesel was synthesized from Styrax confusus Hemsl oil catalyzed by a magnetic solid acid heterogeneous catalyst S2O82?/ZrO2‐TiO2‐Fe3O4, which had a high recovery rate and reusability. The catalyst was prepared by co‐precipitation and characterized by Fourier transform infrared spectroscopy, X‐ray diffraction and Brunauer, Emmett and Teller (BET) adsorption. The properties of the catalyst, including the recovery rate, usage count, magnetic susceptibility and catalytic efficiency, were studied. The results showed that the catalyst has a BET pore diameter of 1.74 nm, BET area of 7.3 m2/g, molar magnetic susceptibility of 1.83 × 10?5 m3/kg and tetragonal structure. In addition, the influences of reaction conditions on yields of biodiesel were also discussed. A fatty acid methyl ester (FAME) yield of 90.02 % was obtained under the conditions of reaction time 1.5 h, reaction temperature 373 K, catalyst amount 5 %, and methanol‐to‐oil molar ratio 8:1. A FAME yield of 65.5 % was obtained when the catalyst was used for the fourth time.  相似文献   

16.
Adsorption equilibrium constants for methyl oleate and methyl linoleate in vapor phase on supported copper and nickel catalysts have been determined using the technique of pulse gas chromatography. The results are discussed in relation to selectivity in fat hydrogenation. Notation: A, column cross-section, m2 ; an,bn, nth Fourier coefficients; c, concentration of adsorbate in bulk flow, mol/m3 ; c* = c/ ∫ 0 cdt, normalized concentration of adsorbate in bulk flow; Ci, concentration of adsorbate in catalyst pores, mol/m3 ; ca, concentration of adsorbate on catalyst surface, mol/kg; cTOT, active area of catalyst as measured by hydrogen adsorption, mol/kg; De, effective diffusion coefficient of adsorbate in catalyst, m2/s; Dea, axial dispersion coefficient based on void cross-section, m2/s; hn, nth coefficient in Hermite polynomial expansion; Hn nth Hermite polynomial; ΔHA, adsorption enthalpy, kJ/mol; ΔHvap , heat of vaporization, kJ/moll; ka, adsorption rate constant, m3/kgs;KA, adsorption equilibrium constant, m3/kg; K0 , preexponential factor defined in Eqn. 8, m3/kg; kf, mass transfer coefficient, m/s; L, bed length, m; q, flow rate, m3/s; R, particle radius, m; Rg, gas constant; t, time, s; T, temperature, K; TF, period of Fourier expansion, s; u = q/A, linear velocity, m/s; z, length coordinate in packed column, m. Greek symbols: δ(t), Dirac delta function; ∈B, void fraction of bed; ∈-p, particle void fraction, ρrp, particle density, kg/m3 ; ξ, radial coordinate in particle, m; μ1, first absolute moment, μ2, second central moment.  相似文献   

17.
《Catalysis communications》2007,8(9):1305-1309
Selective hydrogenation of m-dinitrobenzene to m-nitroaniline (m-NA) catalyzed by polylvinylpyrrolidone stabilized Ru/Al2O3 (PVP-Ru/Al2O3) was studied experimentally. The effects of solvents, metal cation additives and reaction conditions were examined. The highest total yield of m-NA was obtained with 97.9% selectivity at 100% conversion when Sn4+ used as modifier (the molar ratio of m-DNB to catalyst was 477:1, the molar ratio of Sn4+ to ruthenium was 1:4) under suitable conditions.  相似文献   

18.
During the operation of continuous catalyst regeneration reformers, the problem of optimizing the catalyst circulation ratio in the reactor-regenerator system arises. This problem is solved by a combination of real and computational experiments to investigate the regularities of coking on a catalyst’s surface. Based on TGA results for industrial Pt-Sn/γ-Al2O3 catalyst, it is concluded that amorphous coke is formed on the catalyst’s surface during reforming, its quantity at the reactor block outlet being 4–6%, depending on the feed composition and technological parameters of the process. The specific surface of samples is 152 m2/g for the fresh catalyst, 140 m2/g after regeneration, and 118 m2/g at the reactor outlet, which correlates with the quantity of coke on the surface of samples. Mathematical analysis of the coking processes in a reformer with a moving bed show that the catalyst circulation ratio must be maintained in the range of 0.008–0.010 m3/m3 to increase the operating efficiency of an industrial unit. Maintaining optimal conditions enables us to control the coking process, keeping coke concentration as low as possible and the catalyst specific surface as high as possible.  相似文献   

19.
The absorption rates of NO into aqueous solutions of NaClO2 blended with Na2CO3 were investigated in a stirred reactor. The experimental results showed that the reaction rate could be expressed as . An equation for a second‐order reaction rate constant between NO and NaClO2, kmn = 5.79 · 1015 exp(–5557.26/T), was obtained. The addition of Na2CO3 into the solution of NaClO2 decreased the reaction rate constant. The optimal absorption conditions involved a NO concentration of 540 ppm, NaClO2 concentration of 50 mol m–3, Na2CO3 concentration of 10 mol m–3, temperature of 333 K and O2 concentration of 4 %, which were determined by an orthogonal experiment. Under these optimal conditions, it was possible for the absorption rate to reach up to 1.9271 · 10–5 mol/(s m2).  相似文献   

20.
The effects of catalyst support pore size and reaction conditions (T=40-100 °C; ethylene pressure=1.4 MPa; 1-hexene concentration=0-47 mol/m3) on gas-phase polymerization rates and product properties were studied. Catalysts were prepared by impregnation of mesoporous molecular sieves (pore sizes of 2.5-20 nm) with methylaluminoxane and (n-BuCp)2ZrCl2. Temperature rising elution fractionation, differential scanning calorimetry and size exclusion chromatography were used to characterize the products. The results showed that these catalysts contained multiple types of catalytic sites and that the types of sites were a strong function of the support pore size. Ethylene polymerization and 1-hexene incorporation rates were strong functions of support pore size, 1-hexene concentration, and temperature. Large-pored catalysts had higher 1-hexene incorporation rates and the rate of 1-hexene incorporation was a function of polymerization time. Highest polymerization rates were obtained at 80 °C and 1-hexene concentration of 4-12 mol/m3; high 1-hexene concentrations resulted in large decreases in polymerization rates.  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