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 addition of Co to VPO formulations improves the yield of n-butane to maleic anhydride. In this work, different modes of impregnation and two different organic cobalt salts were used. The equilibrated catalysts were characterized using XRD, 31P SEM NMR, FT-IR and acetonitrile adsorption to evaluate Lewis acidity.

The best catalyst was obtained using Co acetyl acetonate for impregnation of the VOHPO4·0.5H2O precursor. This catalyst after equilibration had an optimum concentration of very strong Lewis acid sites, very low concentration of isolated V(V) centers, and no V(V) phases.  相似文献   


2.
Vanadium phosphate catalysts were synthesized via VOPO4·2H2O and were calcined in two different hydrocarbon reaction environments, i.e. n-butane/air and propane/air. Both catalysts are denoted VPDB and VPDP, respectively. Both catalysts exhibited a good crystalline with characteristic peaks of pyrophosphate phase. However, the peaks for VPDP are shown to be more prominent than those of VPDB. BET surface area showed that VPDB gave higher surface area (23 m2 g−1) compared to VPDP (18 m2 g−1). The average V valence state for VPDP is 4.08 and the higher V valence state for VPDB is 4.26 due to higher amount of VV for VPDB. Furthermore 14.2% of VIII was found for VPDP but none for VPDB. SEM micrographs clearly revealed that the morphologies of both catalysts composed of plate-like crystallite that was arranged into the characteristic of rosette cluster. However, the catalyst calcined in n-butane/air environment (VPDB) resulted in an increment of the amount of plate-like crystal formed in the rosette rosebud agglomerates. TPR in H2 profiles of both catalysts gave two reduction peaks corresponding to two kinetically different oxygen species which were associated with VV and VIV phases, respectively. VPDB removed larger amount of active oxygen species linked to VIV phase which eventually caused a higher conversion rate in the selective oxidation of n-butane and propane to maleic anhydride and acrylic acid, respectively.  相似文献   

3.
A new family of vanadium phosphorus oxides (VPO) catalysts has been identified. It consists in a mixture of VPO with niobium phosphate (NbPO). The amorphous NbPO material is introduced during the preparation of the VOHPO4·0.5H2O precursor. It is observed that the VPO–NbPO catalyst is more rapidly activated and gives better performances (n-butane conversion and maleic anhydride selectivity) for mild oxidation of n-butane to maleic anhydride. VPO phases and the NbPO material have been identified in the VPO–NbPO precursor and in the VPO–NbPO catalyst. Nb in VPO crystals and V in NbPO particles have been respectively observed by EDX-STEM. No other VPO crystals than the VOHPO4, 0.5H2O precursor, and (VO)2P2O7 for the catalyst, have been identified by XRD and 31P NMR. 31P NMR by spin echo mapping and 31P MAS NMR have confirmed an interaction of the VPO precursor with Nb and of the NbPO amorphous material with V, as evidenced by EDX-STEM. This should be the reason for the observed improvement of the catalytic results by a redox effect of niobium on the VPO catalyst modifying the V5+/V4+ balance in a favourable way.  相似文献   

4.
Different VOx/TiO2 catalyst have been catalytically tested and studied by in situ-spectroscopic methods (FT-IR, UV/vis, EPR) in the oxyhydrative scission (OHS) of 1-butene and n-butane to acetic acid (AcOH). While 1-butene OHS follows the sequence butene → butoxide → ketone → AcOH/acetate with a multitude of side products also formed, n-butane OHS leads to AcOH, COx and H2O only. Water vapour in the feed improves AcOH selectivity by blocking adsorption sites for acetate. The admixture of Sb2O3 was found to improve AcOH selectivity which is due to deeper V reduction under steady state conditions and lowering of surface acidity. VOx/TiO2 catalysts with sulfate-containing anatase are the most effective ones. Covalently bonded sulfate at the catalyst surface causes specific bonding of VOx, stabilizes active V species and ensures their high dispersity.  相似文献   

5.
A detailed study on the influence of the addition of molybdenum ions on the catalytic behaviour of a selective vanadium–magnesium mixed oxide catalyst in the oxidation of n-butane has been performed. The catalysts have been prepared by impregnation of a calcined V–Mg–O mixed oxides (23.8 wt% of V2O5) with an aqueous solution of ammonium heptamolybdate, and then calcined, and further characterised by several physico-chemical techniques, i.e. SBET, XRD, FTIR, FT-Raman, XPS, H2-TPR. MgMoO4, in addition to Mg3V2O8 and MgO, have been detected in all the Mo-doped samples. The incorporation of molybdenum modifies not only the number of V5+-species on the catalyst surface and the reducibility of selective sites but also the catalytic performance of V–Mg–O catalysts. The incorporation of MoO3 favours a selectivity and a yield to oxydehydrogenation products (especially butadiene) higher than undoped sample. In this way, the best catalyst was obtained with a Mo-loading of 17.3 wt% of MoO3 and a bulk Mo/V atomic ratio of 0.6. From the comparison between the catalytic properties and the catalyst characterisation of undoped and Mo-doped V–Mg–O catalysts, the nature of selective sites in the oxidative dehydrogenation of n-butane is also discussed.  相似文献   

6.
Molecular dynamics simulations were performed for ethane, propane, and n-butane in siliceous faujasite for different numbers of molecules per unit cell (loadings) at 300 K. Both the adsorbed molecules and the zeolite framework were modeled as flexible entities. A new semiempirical analytical potential function for the systems was constructed. From the mean-square displacement of the molecules, self-diffusion coefficients of 18.7 × 10−5, 13.3 × 10−5, and 4.3 × 10−5 cm2/s were calculated for ethane, propane, and n-butane, respectively at a loading of 8 molecules/unit cell. They compare well with experimental values from pulsed-field gradient NMR measurements (10 × 10−5, 9 × 10−5, and 6 × 10−5 cm2/s, respectively). Besides depending on the size of the hydrocarbon, the heats of adsorption and self-diffusion coefficients also strongly depend on the loading of adsorbate molecules. The results suggest that the new intermolecular force field can reasonably describe the adsorption and diffusion behavior of ethane, propane, and n-butane in faujasite zeolite.  相似文献   

7.
Commercial processes presently applied to manufacture iso-butane from n-butane involve continuous co-feeding of chlorine compounds and of hydrogen to enhance the activity and stability of the Pt/alumina catalysts. Aiming for alternative options, we investigated the isomerisation of n-butane at supercritical conditions, using commercially available sulphated zirconia extrudates as a catalyst. As compared to gas-phase conditions, the stability of the catalyst and the production capacities of iso-butane, amounting up to 3.0 kg iso-butane/kg catalyst and hour, could strongly be enhanced. Desulphurisation of the catalyst, causing release of hydrogen sulphide and concomitant catalyst deactivation, was observed at temperatures above 235 °C. Furthermore, the catalyst used in this study exhibited saturation behaviour. The results presented may point the way for the investigations and improvements necessary to make supercritical isomerisation an attractive processing alternative.  相似文献   

8.
Direct formation of isobutene from n-butane was investigated over zeolite TON and γ-Al2O3 supported platinum and platinum–copper catalysts. Addition of Cu decreases Pt dispersion, irrespective of preparation methods and nature of catalyst supports. The presence of potassium was found to reduce acidity and Pt dispersion. The experiments were performed in a fixed-bed microreactor system operating at 673 K and atmospheric pressure. Changing the support from γ-Al2O3 to TON, shows that n-butane conversion is almost independent of acidity. However, significant changes in products selectivities are observed. The selectivities to isobutene, C1–C3 fractions, and aromatics increases drastically from 3.5 to 32.6%, 20.3 to 27.2%, and 3.0 to 20.6%, respectively, for the TON-supported catalyst whereas dehydrogenation is largely predominant when γ-Al2O3 is used as support. Addition of Cu, as expected, has an adverse effect on n-butane conversion as less active sites are available due to the reduction in Pt dispersion. Though Cu addition has marginal effect on isobutene selectivity, it certainly suppresses hydrogenolysis which evidences a reduction in size of the Pt ensembles at the surface of the Pt particles.  相似文献   

9.
Selectivities for skeletal isomerizations of n-butane and n-pentane catalyzed by typical solid acids such as Cs2.5H0.5PW12O40 (Cs2.5), SO42−/ZrO2, WO3/ZrO2, and H-ZSM-5 and their Pt-promoted catalysts were compared. High selectivities for n-butane and low selectivity for n-pentane were observed over Cs2.5 and SO42−/ZrO2, while H-ZSM-5 was much less selective, and WO3/ZrO2 was highly selective for both reactions. The Pt-promoted solid acids were usually selective for these reactions in the presence of H2 except for Pt-H-ZSM-5 for n-butane isomerization. Both the acid strength and pore structure would be factors influencing the selectivity. Mechanism of skeletal isomerization of n-butane was investigated by using 1,4-13C2-n-butane over Cs2.5 and Pt–Cs2.5. It was concluded that n-butane isomerization proceeded mainly via monomolecular pathway with intramolecular rearrangement on Pt–Cs2.5, while it occurred through bimolecular pathway with intermolecular rearrangement on Cs2.5. The higher selectivity on Pt–Cs2.5 would be brought about by the monomolecular mechanism. In the skeletal isomerization of cyclohexane, Pt–Cs2.5/SiO2 was highly active and selective, while Pt–Cs2.5 was less selective. Control in the acid strength of Cs2.5 by the supporting would be responsible for the high selectivity.  相似文献   

10.
High surface area (>300 m2 g−1) nano-structured TiO2 oxides (ns-T) were used as CoMo hydrodesulfurization catalyst support. Cylindrical extrudates were impregnated by incipient wetness with Mo (2.8 Mo at. nm−2) and Co (atomic ratio Co/(Co + Mo) = 0.3). Characterization of impregnated precursors was carried out by N2 physisorption, XRD and atomic absorption and laser-Raman spectroscopies. Sulfided catalysts (400 °C, H2S/H2) were studied by X-ray photoelectronic spectroscopy. As indicated by XRD and after various preparation steps (extrusion, Mo and Co impregnation and sulfiding) the nano-structured material was well preserved. XPS analyses showed that Co and Mo dispersion over the ns-T support was much higher than that on alumina. Very high surface S concentration suggested that even ns-T was partially sulfided during catalyst activation. Dibenzothiophene hydrodesulfurization activity (5.73 MPa, 320 °C, n-hexadecane as solvent) of CoMo/ns-T was two-fold to that of an alumina-supported commercial CoMo catalyst. The improvement was even more remarkable in intrinsic pseudo kinetic constant basis. No important differences in selectivity over the catalysts supported on either Al2O3 or ns-T were observed, where direct desulfurization to biphenyl was favored. Both Mo dispersion and sulfidability were enhanced on the ns-T support where Mo4+ fraction was notably increased (100%) as to that found on CoMo/Al2O3.  相似文献   

11.
The effects of reaction conditions on the synthesis of isobutanol and methanol from syngas over Zr---K and Zr---Mn---K catalysts have been investigated, with the reactions carried out at pressures between 8 and 16 MPa, temperatures between 380 and 440 °C, and gas hourly space velocity (GHSV) between 3 000 and 8 000 h−1. It was found that the Zr---K catalyst exhibited the highest activity and selectivity at 420 °C, 10.0 MPa, and 5 000 h−1, and that the Mn-promoted catalyst increased the yield of isobutanol with the corresponding optimum temperature being 400 °C. Moreover, the higher pressure and GHSV favored isobutanol synthesis. At 16. 0 MPa, 400 °C, 5 000 h−1, the space time yield of isobutanol was about 21.5 ml per ml cat. h−1, and selectivity reached 98% over the Zr---Mn---K catalyst.  相似文献   

12.
A Co-N on alumina catalyst yielded high performance in the oxidative cracking of n-butane to ethylene and propylene. A total of 47.7 wt.% yield of olefins including 31% of ethylene and 13% of propylene were obtained at 82% of n-butane conversion at 600 °C. Catalyst characterization by SEM, X-ray photoelectron spectroscopy (XPS), XRD and TPR studies suggested that a cobalt oxynitride phase was formed. This resulted in lowering the oxygen binding energy leading to enrichment in mobile, low energy, oxygen species that significantly accelerates the formation of lower olefins.  相似文献   

13.
An evaluative investigation of the Fischer–Tropsch performance of two catalysts (20%Co/Al2O3 and 10%Co:10%Mo/Al2O3) has been carried out in a slurry reactor at 2 MPa and 220–260 °C. The addition of Mo to the Co-catalyst significantly increased the acid-site strength suggesting strong electron withdrawing character in the Co-Mo catalyst. Analysis of steady-state rate data however, indicates that the FT reaction proceeds via a similar mechanism on both catalysts (carbide mechanism with hydrogenation of surface precursors as the rate-determining step). Although chain growth, , on both catalysts were comparable (  0.6), stronger CH2 adsorption on the Co-Mo catalyst and lower surface concentration of hydrogen adatoms as a result of increased acid-site strength was responsible for the lower individual hydrocarbons production rate compared to the Co catalyst. The activation energy, E, for Co (96.6 kJ mol−1), is also smaller than the estimate for the Co-Mo catalyst (112 kJ mol−1). Transient hydrocarbon rate profiles on each catalyst are indicative of first-order processes, however the associated surface time constants are higher for alkanes than alkenes on individual catalysts. Even so, for each homologous class, surface time constants for paraffins are greater for Co-Mo than Co, indicative that the adsorption of CH2 species on the Co-Mo surface is stronger than on the monometallic Co catalyst.  相似文献   

14.
Addition of molybdenum to nickel catalysts has a favourable effect on their properties in the steam reforming of hydrocarbons. Some attempts were made to explain the mechanism of promoter action by determining the properties of such catalysts in hydrogenolysis and cracking reactions. With small amounts of promoter (≤0.5 wt.%) advantageneous changes in selectivity of steam reforming and disadvantageneous changes in n-butane hydrogenolysis were observed. The promoter does not affect practically the catalyst properties in n-butane cracking. The effect of molybdenum was compared with that of potassium promoter applied in the industry.  相似文献   

15.
Mesostructured lamellar, hexagonal and cubic vanadium–phosphorus-oxide (VPO) phases were prepared employing cationic, anionic and alkylamine surfactants under mild conditions and low pH. The obtained mesophases displayed desirable vanadium oxidation states (+3.8 to +4.3) and P/V molar ratios 1.0 for the partial oxidation of n-butane to maleic anhydride. As-synthesized mesostructured VPO underwent phase transformations to various mesostructured and dense VPO phases depending on the post-synthesis treatment. The phase transformations of mesostructured VPO during Soxhlet extraction and thermal treatment in N2 have been observed for the first time. These transformations were explained by the changes in the surfactant packing parameter, g. Calcination in air produced more disordered mesostructures and dense VPO phases such as γ-VOPO4 and (VO)2P2O7.  相似文献   

16.
This study focuses on the direct sulfur recovery process (DSRP), in which SO2 can be directly converted into elemental sulfur using a variety of reducing agents over Ce1−xZrxO2 catalysts. Ce1−xZrxO2 catalysts (where x = 0.2, 0.5, and 0.8) were prepared by a citric complexation method. The experimental conditions used for SO2 reduction were as follow: the space velocity (GHSV) was 30,000 ml/g-cat h and the ratio of [CO (or H2, H2 + CO)]/[SO2] was 2.0. It was found that the catalyst and reducing agent providing the best performance were the Ce0.5Zr0.5O2 catalyst and CO, respectively. In this case, the SO2 conversion was about 92% and the sulfur yield was about 90% at 550 °C. Also, a higher efficiency of SO2 removal and elemental sulfur recovery was achieved in the reduction of SO2 with CO as a reducing agent than that with H2. In the reduction of SO2 by H2 over the Ce0.5Zr0.5O2 catalyst, SO2 conversion and sulfur yield were about 92.7% and 73%, respectively, at 800 °C. Also, the reduction of SO2 using synthetic gas with various [CO]/[H2] molar ratios over the Ce0.5Zr0.5O2 catalyst was performed, in order to investigate the possibility of using coal-derived gas as a reducing agent in the DSRP. It was found that the reactivity of the SO2 reduction using the synthetic gas with various [CO]/[H2] molar ratios was increased with increasing CO content of the synthetic gas. Therefore, it was found that the Ce1−xZrxO2 catalysts are applicable to the DSRP using coal-derived gas, which contains a larger percentage of CO than H2.  相似文献   

17.
An assessment of the influence of the crystal structure, surface hydroxylation state and previous oxidation/reduction pretreatments on the activity of sulfate-zirconia catalysts for isomerization of n-butane was performed using crystalline and amorphous zirconia supports. Different sulfation methods were used for the preparation of bulk and supported SO42−-ZrO2 with monoclinic, tetragonal and tetragonal+monoclinic structures. Activity was important only for the samples that contained tetragonal crystals. The catalysts prepared from pure monoclinic zirconia showed negligible activity. SO42−-ZrO2 catalysts prepared by sulfation of crystalline zirconia displayed sites with lower acidity and cracking activity than those sulfated in the amorphous state. Prereduction of the zirconia samples with H2 was found to greatly increase the catalytic activity, and a maximum rate was found at a reduction temperature of 550–600 °C, coinciding with a TPR peak supposedly associated with the removal of lattice oxygen and the creation of lattice defects. A weaker dependence of catalytic activity on the density or type of surface OH groups on zirconia (before sulfation) was found in this work.

A model of active site generation was constructed in order to stress the dependence on the crystal structure and crystal defects. Current and previous results suggest that tetragonal structure in active SO42−-ZrO2 is a consequence of the stabilization of anionic vacancies in zirconia. Anionic vacancies are in turn supposed to be related to the catalytic activity for n-butane isomerization through the stabilization of electrons from ionized intermediates.  相似文献   


18.
In this study, the PEG-derived VPO precursors were subject to water refluxing (90 °C, 8 h) and in situ activation in a steam-containing environment. For comparison, the VPO precursors without water refluxing were also activated in a similar manner. The IR measurement indicated that the majority of PEG in the precursor has been removed during water refluxing, and the VPO is essentially unenwrapped by PEG. The consequence is an increase of particle size and crystallinity of VPO as well as decreases in surface acidity and site density. The activation of catalysts in a stream-containing environment has an influence on the content of V5+ species and on the reduction behavior of the VPO catalysts. The VPO catalyst activated with 20% water vapor (by volume) in the feed shows the highest crystallinity. Compared with the non-PEG-derived VPO precursor, the PEG-derived one undergoes structure changes of higher severity during the water/steam treatments. The VPO catalyst generated from the PEG-derived precursor with water refluxing and activated with steam (20%) exhibited a propane conversion of 25% and a (AA + HAc) selectivity of 70%. The superior catalyst behavior can be interpreted in terms of the higher crystallinity of the (VO)2P2O7 phase, the lower content of V5+ species, and the milder surface acidity as caused by the water/steam treatments.  相似文献   

19.
Dehydrogenation-cracking double-stage (tandem) reaction of n-butane was studied using a Pt-Sn type dehydrogenation catalyst and a cracking catalyst (rare earth-loaded HZSM-5). n-Butane was firstly dehydrogenated to n-butene (1- and 2-butene) over the Pt-Sn catalyst loaded at the upper part of the reactor. Then n-butene was successively converted to ethylene and propylene over the cracking catalyst loaded at the lower part of the reactor. The yield of light olefins (ethylene+propylene) was 58% at 650 °C. The key to obtaining ethylene and propylene in high yield was to determine how the bimolecular reactions of olefins to aromatic and heavier products can be inhibited. It was proved that the loaded rare earths played an important role in inhibiting the bimolecular reactions.  相似文献   

20.
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   

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