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.
Supports of Mo and NiMo catalysts were prepared and tested in hydrotreatment model reactions. These supports, composed of zirconia and various amounts of yttria, were obtained by synthesis in molten salts.

It was found that the distribution of yttrium in zirconium oxide was homogeneous. Moreover, it was shown that the crystalline structure as well as the textural properties (especially the porosity) were stabilized.

These solids were then used as supports of Mo and NiMo sulphides and their activities were compared to those of a commercial NiMo alumina catalyst.

In biphenyl hydrogenation, with the same coverage of Mo (2.8 at/nm2), the activities per gram of both catalysts (supported on zirconia and alumina) were similar, while the activity per atom of Mo of the catalyst supported on ZrO2---Y2O3 was twice the activity of the catalyst supported on alumina.

For NiMo catalysts, a ratio R= Ni/(Ni+Mo)=0.4 with Mo = 2.8 at/nm2 and a co-impregnation of the Mo and Ni were required to have a good synergetic effect. The activity per atom of Mo in biphenyl hydrogenation was then enhanced more than twofold when compared to the NiMo/alumina catalyst. In an HDN model reaction (conversion of diethylanyline in the presence of quinoline) the results obtained with zirconia were much better than with alumina.  相似文献   


2.
The hydrodesulphurization (HDS) of thiophene at atmospheric pressure, hydrodenitrogenation (HDN) of 2,6 diethylaniline (DEA) and of 1,2,3,4 tetrahydroquinoline (THQ) at 70 atm in a dynamic flow microreactor - these molecules being present alone or mixed in the feed - and finally HDN of mixtures of quinoline (Q) at 70 atm or of phenanthridine (Ph) at 140 atm with DEA in a batch reactor were the test reactions used to compare the performances of (Ni)---Mo---S catalysts supported on MgAl2O4 and NiAl2O4 with a classical Ni---Mo---S/γ Al2O3 commercial catalyst. Some of the catalysts supported on NiAl2O4 appeared to be superior to the commercial one, whereas those supported on MgAl2O4 were found to be less active. HDS of thiophene, as well as the amount of carbon monoxide adsorbed at 0°C by sulphided catalysts permitted the prediction of the order of HDN activity of the catalysts for the conversion of some nitrogen-containing molecules. A strong inhibiting effect of molecules like Q, THQ or Ph (or some of their products formed during the process) on the HDN of DEA has been found and attributed to the competitive adsorption of these molecules on the supported active phase.  相似文献   

3.
A study on the catalytic properties of the transition metals (Ni,Co,Mo)-carbides, -nitrides for thiophene and dibenzothiophene hydrotreating was conducted. The (Ni,Co)-Mo carbides and the corresponding (Ni,Co)-Mo nitride phases showed a catalytic activity higher than conventional bimetallic (Ni,Co)-Mo sulfides. In addition, a study was done on the effect of the atomic ratios, i.e., 0.1 ≤ M+/(M+ + Mo) ≤ 0.9 where M+ stands for Ni or Co, and the concentration of promoters such as phosphorous, which was a structural stabilizing agent. The catalytic performance of the bimetallic NiMo and CoMo carbides and nitrides was studied using thiophene and dibenzothiophene hydrodesulfurization (HDS) as model reactions at 623 K and P = 1 atm. The catalytic activity of the dispersed carbide and nitride phases on the alumina carrier was more significant than that of the reference catalysts, alumina supported NiMo-S and CoMo-S. The metallic character of the NiMo and CoMo carbides was evidenced by their higher hydrogenation activity in thiophene HDS, while the nitrides favored both hydrogenation and hydrogenolysis type reactions.  相似文献   

4.
Hydrodenitrogenation of pyridine over alumina-supported iridium catalysts   总被引:1,自引:0,他引:1  
The catalytic properties of alumina-supported Ir catalysts (≈1 wt% Ir) were studied in the hydrodenitrogenation (HDN) of pyridine at 320°C and 20 bar of pressure in the absence as well as presence of parallel hydrodesulfurization (HDS) of thiophene. The effects of Ir precursor (Ir(AcAc)3, Ir4(CO)12, H2IrCl6, (NH4)2IrCl6), metal dispersion and sulfur addition were investigated. Ir4(CO)12 gave the most active catalyst which was ascribed to a lower amount of contaminants originated from the starting Ir compounds rather than to a better Ir dispersion. The decrease of Ir dispersion by sintering in air led to much higher decrease of the rate of C–N bond hydrogenolysis than that of pyridine hydrogenation. The Ir dispersion determined partly the HDN selectivity; a better dispersed Ir phase gave a lower amount of intermediate piperidine. Presulfidation of the reduced catalyst led to 20% decline of the rates of both consecutive HDN steps. An additional and much larger activity decline was caused by the simultaneous execution of HDS. The competitive adsorption of thiophene (or H2S) was selectively affecting C–N bond hydrogenolysis more than pyridine hydrogenation. The alumina-supported Ir catalysts possessed much higher HDN activity and HDN/HDS selectivity than a conventional NiMo system.  相似文献   

5.
In this paper, the effects of fluorine and phosphorus on the physical and chemical properties of Ni–Mo/Al2O3 catalysts and the hydrodenitrogenation (HDN) activity of quinoline were investigated. The acidity, pore structure, and dispersion of Mo of the catalysts were analyzed with TG-DTA, BET, and XRD techniques. The activities of hydrodenitrogenation and hydrogenation of the catalysts were investigated using hydrogenation of quinoline at high pressure in a micro-reactor. Experimental results verified that phosphorus can promote the formation of moderate and strong acidic sites, the dispersion of Mo, and the formation of the active phases; therefore, the hydrogenation activity of aromatic rings and the hydrogenolysis activity of C–N bonds increase. The hydrogenation and hydrogenolysis accelerate each other, which results in the increase of HDN activity. It is concluded that phosphorus is a promoter for HDN activity of the Ni–Mo/Al2O3 catalysts. Fluorine can promote the formation of weak and moderate acidic sites and the dispersion of Mo, but inhibit the formation of the active phases. Therefore, the hydrogenation activity of aromatic rings and the hydrogenolysis activity of C–N bonds decrease, which results in the decrease of HDN activity. It is concluded that fluorine is not a promoter for HDN activity of the Ni–Mo/Al2O3 catalysts. The possible promoting mechanism of fluorine and phosphorus for the Ni–Mo/Al2O3 catalyst is put forward and discussed.  相似文献   

6.
ZrO2–TiO2 mixed oxide (30–70 mol/mol) was prepared by low-temperature sol–gel followed by solvo-thermal treatment (1 day) at various temperatures (40, 80, 120, 160 and 200 °C). Selected samples of the corresponding single oxides were also prepared. Materials characterization was carried out by N2 physisorption, XRD, thermal analysis (TG-DTA) and UV–vis DRS, infra-red and Laser-Raman spectroscopies. Binary solids of enhanced pore volume and pore size diameter were obtained by increasing the post-treatment severity. Anatase TiO2 micro-segregation was evidenced by Raman spectroscopy for the mixed oxide solvo-treated at the highest temperature. This solid also showed the highest crystallization temperature to ZrTiO4 (702 °C). Mo impregnated (2.8 atom nm−2) on various mixed oxides was sulfided under H2S/H2 (400 °C, 1 h), the catalysts being tested in the dibenzothiophene hydrodesulfurization (HDS, T = 320 °C, P = 5.59 MPa). By increasing the severity of the solvo-treatment improved supports for MoS2 phase were obtained. The HDS activity of the catalyst with carrier post-treated at 200 °C was 40% higher (in per total mass basis) than that of sulfided Mo supported on the binary oxide solvo-treated at 80 °C. The ZrO2–TiO2-supported catalysts showed higher selectivity to products from the hydrogenation route than their counterparts supported on either single oxide.  相似文献   

7.
A series of Re-containing catalysts supported on activated carbon, with Re loading between 0.74 and 11.44 wt.% Re2O7, was prepared by wet impregnation and tested in the simultaneous hydrodesulphurisation (HDS) and hydrodenitrogenation (HDN) of a commercial gas oil. Textural analysis, XRD, X-ray photoelectron spectroscopy (XPS) and surface acidity techniques were used for physicochemical characterisation of the catalysts. Increase in the Re concentration resulted in a rise in the HDS and HDN activity due to the formation of a monolayer structure of Re and the higher surface acidity. At Re concentrations >2.47 wt.% Re2O7 (0.076 Re atoms nm−2) the reduction in the catalytic activity was related to the loss in specific surface area (BET) due to reduction in the microporosity of the carbon support. The magnitude of the catalytic effect was different for HDS and HDN, and depended strongly on the Re content and reaction temperature. The apparent activation energies were about 116–156 kJ mol−1 for HDS and 24–30 kJ mol−1 for HDN. This led to a marked increase in the HDN/HDS selectivity with decreasing temperature (values >3 at 325 °C), due to the large differences in the apparent activation energies of HDS and HDN found for all catalysts. A gradual increase in the HDN/HDS selectivity with increased Re loading was also found and related to the observed increase of catalyst acidity. The results are compared with those obtained for a series of Re/γ-Al2O3 catalysts.  相似文献   

8.
Hydrodenitrogenation of isoquinoline   总被引:1,自引:0,他引:1  
To determine the formation and reactivity of addition compounds produced during hydrodenitrogenation (HDN), we investigated the HDN of isoquinoline for a sulfided Ni–Mo/Al2O3 catalyst operated under a hydrogen pressure of 12 MPa (cold charge) in the temperature range 300–375°C. The reaction products were classified into five groups of compounds:

1. hydrogenated derivatives of isoquinoline (tetrahydroisoquinolines, decahydroisoquinolines, and their isomers);

2. nitrogen-containing ring-opened products (1-amino-2-(2-methylphenyl)ethane and 1-amino-1-(2-ethylphenyl)methane);

3. denitrogenated products (1-ethyl-2-methylbenzene, 1-ethyl-2-methylcyclohexane, and their isomers);

4. addition products (hydrocarbons with molecular weights of 238, 244, and 250 and nitrogen-containing compounds with molecular weights of 249, 251, and 257); and

5. cracked products (toluene, ethylbenzene, dimethylbenzenes, and their hydrogenated derivatives).

Most of the nitrogen-containing addition compounds appear to be substituted on the nitrogen atom. The HDN of isoquinoline was more than 10 times faster than the HDN of quinoline, whereas the hydrogenation of isoquinoline was difficult compared to the hydrogenation of quinoline. The reaction network for the HDN of isoquinoline is also presented.  相似文献   


9.
The effect of H2S on the rates of toluene hydrogenation of sulfided Mo, Ni, Ni---Mo and Ni---Mo---P/alumina catalysts has been determined at 6 MPa, 350°C over a large range of H2S partial pressure. In those conditions, Ni/alumina is nearly inactive. For the three other catalysts, similar trends are found with in particular no effect of H2S at high partial pressure. In the presence of NH3, the effect of H2S remains globally the same.  相似文献   

10.
Alumina supported RuMo sulphide catalysts having different weights of metal content and atomic composition ratios R=Ru/(Ru+Mo) were prepared by using ammonium heptamolybdate dissolved in water or a 10% (NH4)2S-H2O solution and RuCl3.3H2O as precursor compounds. Their activities were studied in the hydrodesulphurization (HDS) of thiophene and in the hydrogenation (HYD) of biphenyl, and optimized in terms of the preparation method and the sulphidation process. Some hydrodenitrogenation (HDN) tests were also performed on these catalysts. Electron microscopy and XPS measurements were performed, and the nature of the active phase was discussed.  相似文献   

11.
Selectivity in heterogeneous catalytic processes   总被引:2,自引:0,他引:2  
The selectivity of several catalytic systems was studied. Shape selectivity of Pt on carbon-fiber catalysts was demonstrated in the competitive hydrogenation of 1-hexene and cyclohexene and in the parallel dehydrogenation of cyclohexanol to cyclohexanone and phenol. Both reactions were carried out in a gas-phase fixed-bed reactor. Catalysts prepared on carbon fibers, containing pores with small constrictions (5 Å) yielded significantly higher rates of hydrogenation of 1-hexene compared to those of cyclohexene and selectively produced cyclohexanone from cyclohexanol. Other catalysts, supported on carbon fibers with large constrictions (7 Å) or activated carbon, displayed comparable rates of hydrogenation for both reactants and yielded cyclohexanone as well as phenol from cyclohexanol. Nitration of o-xylene with nitrogen dioxide was carried out in the gas phase over a series of solid acid catalysts packed in a fixed bed. Several zeolites, supported sulfuric acid, and sulfated zirconia were tested. Zeolite H-β was found to be the most active and selective catalyst for the production of 4-nitro-o-xylene. A preliminary kinetic model indicated that the selectivity to 4-nitro-o-xylene increased with decreasing concentration of nitrogen dioxide. Alkylation of phenol with methanol was performed on zeolites, supported sulfuric and phosphoric acids, and sulfated zirconia packed in a fixed-bed. The ratio of o- to c-alkylation, measured at 180°C and methanol to phenol feed molar ratio of unity, ranged from 4 with the supported acids to 2 with zeolite H-β. This ratio decreased with temperature. The ratio of o- to p-cresol changed from about 2 in zeolites in supported sulfuric acid and to 0.5 in phosphoric acid supported on carbon fibers.  相似文献   

12.
The effect of support material on the catalytic performance for methane combustion has been studied for bimetallic palladium–platinum catalysts and compared with a monometallic palladium catalyst on alumina. The catalytic activities of the various catalysts were measured in a tubular reactor, in which both the activity and stability of methane conversion were monitored. In addition, all catalysts were analysed by temperature-programmed oxidation and in situ XRD operating at high temperatures in order to study the oxidation/reduction properties.

The activity of the monometallic palladium catalyst decreases under steady-state conditions, even at a temperature as low as 470 °C. In situ XRD results showed that no decomposition of bulk PdO into metallic palladium occurred at temperatures below 800 °C. Hence, the reason for the drop in activity is probably not connected to the bulk PdO decomposition.

All Pd–Pt catalysts, independently of the support, have considerably more stable methane conversion than the monometallic palladium catalyst. However, dissimilarities in activity and ability to reoxidise PdO were observed for the various support materials. Pd–Pt supported on Al2O3 was the most active catalyst in the low-temperature region, Pd–Pt supported on ceria-stabilised ZrO2 was the most active between 620 and 800 °C, whereas Pd–Pt supported on LaMnAl11O19 was superior for temperatures above 800 °C. The ability to reoxidise metallic Pd into PdO was observed to vary between the supports. The alumina sample showed a very slow reoxidation, whereas ceria-stabilised ZrO2 was clearly faster.  相似文献   


13.
Mixed oxides of alumina and zirconia having a relative composition of 50, 80 and 100% Zr2O were synthesized by means of sol–gel methods. The catalysts were sulfated with H2SO4 1N, and were loaded with 0.3% Pt metal using the incipient wetness technique. The characterization of the physicochemical properties was carried out using XRD, N2-adsorption at 78 K, and SEM. The catalytic properties of the Al2O3–ZrO2 series were studied by means of dehydration of 2-propanol at 180°C and isomerization of n-hexane at 250°C, 1 atm. The sulfated solids presented a high surface acidity and a limited crystallinity, together with high activity for alcohol dehydration (i.e. 2-propanol). On the other hand, the Al2O3–ZrO2 solid solutions (i.e. those having a 20–80% composition) turned out to be the most active ones for the isomerization of n-hexane.  相似文献   

14.
Zirconia catalysts in Meerwein-Ponndorf-Verley reduction of citral   总被引:3,自引:0,他引:3  
Zirconium-containing catalysts were found to be active in the Meerwein-Ponndorf-Verley (MPV) reduction of citral. Good activity and selectivity to the reduced alcohol, geraniol and nerol, were observed over hydrous zirconia and zirconium 1-propoxide supported on silica. In particular, hydrous zirconia calcined at temperatures below 300 °C was highly active. Hydrous zirconia catalysts modified by NaOH, NH4F, PO43− and SO42− had lower activity. Surface hydroxyl groups are postulated to be involved in ligand exchange with the reductant, 2-propanol. Zr-zeolite beta showed high activity, but poorer selectivity than the other two samples, due to subsequent dehydration of the product formed. For all catalysts, the trans-isomer of citral was preferentially reduced over the cis-isomer.  相似文献   

15.
The siliceous and the metal substituted (B or Al)-SBA-15 molecular sieves were used as a support for NiMo hydrotreating catalysts (12 wt.% Mo and 2.4 wt.% Ni). The supports were characterized by X-ray diffraction (XRD), scanning electron microscopy and N2 adsorption–desorption isotherms. The SBA-15 supported NiMo catalysts in oxide state were characterized by BET surface area analysis and XRD. The sulfided NiMo/SBA-15 catalysts were examined by DRIFT of CO adsorption and TPD of NH3. The HDN and HDS activities with bitumen derived light gas oil at industrial conditions showed that Al substituted SBA-15 (Al-SBA-15) is the best among the supports studied for NiMo catalyst. A series of NiMo catalysts containing 7–22 wt.% Mo with Ni/Mo weight ratio of 0.2 was prepared using Al-SBA-15 support and characterized by BET surface area analysis, XRD and temperature programmed reduction and DRIFT spectroscopy of adsorbed CO. The DRIFT spectra of adsorbed CO showed the presence of both unpromoted and Ni promoted MoS2 sites in all the catalysts, and maximum “NiMoS” sites concentration with 17 wt.% of Mo loading. The HDN and HDS activities of NiMo/Al-SBA-15 catalysts were studied using light gas oil at temperature, pressure and WHSV of 370 °C, 1300 psig and 4.5 h−1, respectively. The NiMo/Al-SBA-15 catalyst with 17 wt.% Mo and 3.4 wt.% of Ni is found to be the best catalyst. The HDN and HDS activities of this catalyst are comparable with the conventional Al2O3 supported NiMo catalyst in real feed at industrial conditions.  相似文献   

16.
The hydrogenation of CO over mixed oxides (RhVO4, Rh2MnO4) supported on SiO2 has been studied after H2 reduction at 300°C and at 500°C, and the results compared with those of unpromoted Rh/SiO2 catalysts. Rh was more highly dispersed (40 Å) after the decomposition of RhVO4 by the H2 reduction than those of Rh2MnO4/SiO2 and unpromoted Rh/SiO2 catalysts. The activity and the selectivity to C2 oxygenates of the mixed-oxide catalysts after the H2 reduction were higher than those of the unpromoted Rh/SiO2 catalysts, but the activity of the RhVO4/SiO2 catalyst increased more dramatically after the decomposition by the H2 reduction at 300°C, and hence the yield of C2 oxygenates increased. These results suggest that a strong metal–oxide interaction (SMOI) was induced by the decomposition of the mixed oxides after the H2 reduction. The catalytic activity and selectivity were reproduced repeatedly by the calcination and reduction treatments of the spent (used) catalyst because of the regeneration of RhVO4 and redispersion of Rh metal.  相似文献   

17.
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   

18.
The complete catalytic oxidation of 1,2-dichloroethane (DCE) and trichloroethylene (TCE) over alumina supported noble metal catalysts (Pt and Pd) was evaluated. Experiments were performed at conditions of lean hydrocarbon concentration (around 1000 ppm) in air, between 250°C and 550°C in a conventional fixed bed reactor. The catalysts were prepared in a range of metal contents from 0.1 to 1 wt%. Palladium catalysts resulted to be more active than platinum catalysts in the oxidation of both chlorinated volatile organic compounds. DCE was completely destructed at 375°C, whereas TCE required 550°C. HCl was the only chlorine-containing product in the oxidation of DCE in the range of 250–400°C. Tetrachloroethylene was observed as an intermediate in the oxidation of TCE, being formed to a significant extent between 400°C and 525°C. CO was also detected in the oxidation of both DCE and TCE over Pd catalysts, though at temperatures of complete destruction, CO2 was the only carbon-containing product. The Pt catalysts were selective to CO2 at the studied conditions.  相似文献   

19.
Nano-scale, binary, 4.5 wt.% Fe–0.5 wt.% M (M = Pd, Mo or Ni) catalysts supported on alumina have been shown to be very effective for the decomposition of lower alkanes to produce hydrogen and carbon nanofibers or nanotubes. After pre-reduction at 700 °C, all three binary catalysts exhibited significantly lower propane decomposition temperatures and longer time-on-stream performances than either the non-metallic alumina support or 5 wt.% Fe/Al2O3. Catalytic decomposition of propane using all three catalysts yielded only hydrogen, methane, unreacted propane, and carbon nanotubes. Above 475 °C, hydrogen and methane were the only gaseous products. Catalytic decomposition of cyclohexane using the (4.5 wt.% Fe–0.5 wt.% Pd)/Al2O3 catalyst produced primarily hydrogen, benzene, and unreacted cyclohexane below 450 °C, but only hydrogen, methane, and carbon nanotubes above 500 °C. The carbon nanotubes exhibited two distinct forms depending on the reaction temperature. Above 600 °C, they were predominantly in form of multi-walled nanotubes with parallel walls in the form of concentric graphene sheets. At or below 500 °C, carbon nanofibers with capped and truncated stacked-cone structure were produced. At 625 °C, decomposition of cyclohexane produced a mixture of the two types of carbon nanostructures.  相似文献   

20.
Sulfidation of trimetallic CoNiMo/Al2O3 catalysts was studied by thermogravimetry at 400 °C under flow and pressure conditions. Results were compared with those obtained on prepared and industrial CoMo/Al2O3 and NiMo/Al2O3 catalysts. The amount of sorbed H2S on the sulfided solids was measured at 300 °C in the H2S pressure range 0–3.5 MPa at constant H2 pressure (3.8 MPa). The adsorption isotherms were simulated using a model featuring dissociated adsorption of H2S on supported metal sulfides and bare alumina. The amount of sulfur-vacancy sites could thus be determined under conditions close to industrial practice. A relationship with activity results for thiophene HDS and benzene hydrogenation was sought for.  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