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 in-situ hydrodesulfurization (HDS) reaction of thiophene was performed by using hydrogen which was generated by a water gas shift reaction (WGSR) in a same catalyst bed. The catalyst used was commercial CoMo/γ-Al2O3 and it was used after presulfiding. The activity in the conversion of thiophene by using hydrogen generated in-situ from a WGSR was inferior to that by the pure hydrogen. The lower efficiency in the in-situ HDS with WGSR was attributed to water, carbon monoxide and carbon dioxide which were mixed after WGSR. The following rate equation, which was revised from that of Satterfield, was proposed for this in-situ HDS reaction of thiophene with WGSR to explain the observed phenomena.  相似文献   

2.
The conversion of used lubricating oil to transport fuel by direct cracking is a suitable way to dispose of waste oil. The aim of this research was to study the catalytic cracking of used lubricating oil, and thus change its classification from a waste produce to a liquid fuel as a new alternative for the replacement of petroleum fuels. The experiments were carried out in a 70-cm3 batch microreactor at a temperature of 648-698 K, initial hydrogen pressure of 1-4 MPa, and reaction time of 10-90 min over sulfated zirconia as a catalyst. The conditions that gave the highest conversion of naphtha (~20.60%) were a temperature of 698 K, a hydrogen pressure of 2 MPa, and a reaction time of 60 min, whereas, under the same conditions, kerosene, light gas oil, gas oil, long residues, hydrocarbon gases, and a small amounts of solids were present (~9.04%, 15.61%, 5.00%, 23.30%, 25.58%, and 0.87%, respectively). The liquid product consisted of C7-C15 of n-paraffins, C7-C9 of branched chain paraffin and aromatic compounds such as toluene, ethylbenzene and xylene. The kinetic study reveals that the catalytic cracking of waste lubricating oil follows a second order reaction, and the kinetic model is defined as k(s−1)=2.88×104exp[−(103.68 kJmol−1)/(RT)].  相似文献   

3.
The low temperature activity of Au/Fe2O3 catalysts towards CO oxidation was examined with respect to the temperature of pre-treatment and presence of water. The activity of all the prepared catalysts decreased as a result of a high temperature treatment (HTT) at 400 °C. The inclusion of water in the gas stream significantly enhanced the oxidation of CO at room temperature. When tested under water gas shift reaction (WGSR) conditions, significantly higher temperatures were required to convert CO to CO2, thereby excluding the possibility of the WGSR during CO oxidation in the presence of H2O at room temperature. The loss of activity for CO oxidation is attributed to the loss of hydroxyl groups and reduction of Au3+ to metallic gold during HTT. The observations are consistent with the model for hydroxyl promotion of the decomposition of a carbonate intermediate by transformation to less stable bicarbonate.  相似文献   

4.
In this paper, the liquid products from supercritical water gasification (SCWG) of glycerol were analyzed and some intermediates were identified. A simplified reaction pathway for gases production from SCWG of glycerol was proposed. The first quantitative kinetics model for describing the gaseous products (H2, CO, CH4 and CO2) of SCWG of glycerol was developed. The model comprises seven reactions to describe the typical reactions in SCWG, and the reaction rate constant of each reaction was obtained by using the nonlinear least-square fitting method. The reaction rate analysis showed that the main sources of hydrogen yield were glycerol pyrolysis and steam reforming of intermediates, while the hydrogen yield from water–gas shift reaction (WGSR) was very small. The temperature estimated by the kinetics model for completely SCWG of glycerol solution was given. In addition, the sensitivity analysis of rate constant of WGSR was done based on the model.  相似文献   

5.
This paper describes an investigation on CuO and CuO-ZnO catalysts supported on CeO2 and CeO2-La2O3 oxides, which were designed for the low temperature water-gas shift reaction (WGSR). Bulk catalysts were prepared by co-precipitation of metal nitrates and characterized by energy-dispersive spectroscopy (EDS), X-ray diffraction (XRD), surface area (by the BET method), X-ray photoelectron spectroscopy (XPS), and in situ X-ray absorption near edge structure (XANES). The catalysts' activities were tested in the forward WGSR, and the CuO/CeO2 catalyst presented the best catalytic performance. The reasons for this are twofold: (1) the presence of Zn inhibits the interaction between Cu and Ce ions, and (2) lanthanum oxide forms a solid solution with cerium oxide, which will cause a decrease in the surface area of the catalysts. Also the CuO/CeO2 catalyst presented the highest Cu content on the surface, which could influence its catalytic behavior. Additionally, the Cu0 and Cu1+ species could influence the catalytic activity via a reduction-oxidation mechanism, corroborating to the best catalytic performance of the Cu/Ce catalyst.  相似文献   

6.
《Applied Catalysis A: General》2001,205(1-2):101-108
Hydrodesulfurization (HDS) and hydrodenitrogenation (HDN) of a commercial medium cycle oil (MCO) were performed over a commercial NiMoS/Al2O3 catalyst through both single- and two-stage hydrotreatments at 340°C. The reaction atmosphere was replaced with fresh hydrogen, with or without additional dose of catalyst, for the second-stage treatment to determine the mechanism of reduced activity. Sulfur and nitrogen molecular species in MCO were identified by gas chromatography with an atomic emission detector (GC-AED) to quantify their respective reactivities and susceptibilities to inhibition. Under single-stage (30 min) conditions, the reactivity orders in HDS and HDN were BT>DBT>4-MDBT>4,6-DMDBT and In>alkylIn>Cz>1-Cz>1,8-Cz, respectively. Additional reaction time beyond the initial 30 min, without atmosphere or catalyst replacement, gave little additional conversion. Replacement of the first-stage gas with fresh hydrogen strongly improved second-stage conversions, particularly those of the more refractory species. An additional dose of catalyst for the second stage with hydrogen renewal facilitated additional HDS of dibenzothiophene (DBT), 4-monomethylated DBT (4-MDBT), and 4,6-dimethylated DBT (4,6-DMDBT) which was independent of their initial reactivity, while HDN of carbazole (Cz), 1-Cz, and 1,8-Cz was improved, the least reactive species being most denitrogenated. Such results suggest the strong inhibition of the gaseous products H2S and NH3. The catalyst deactivation was most marked with HDN of 1,8-Cz, suggesting that acidity is essential to the reaction. H2S is suspected to inhibit both S elimination and hydrogenation of S and N species at the level of concentration obtained during desulfurization. The inhibition by remaining substrates may still influence the HDS and HDN of refractory species in the second stage, even if their contents were reduced by the first stage. It appears very important to clarify the inhibition factor of all species on the refractory sulfur species, and to determine the inhibition susceptibility of these species at their lowered concentration to enable the effective achievement of 50 ppm sulfur level in distillate products. The conversions of inhibitors must be accounted for during reactions. Catalyst and reaction configuration to reduce the inhibition by the gaseous products are the keys for deep refining.  相似文献   

7.
We demonstrate ways to prepare active and stable Pt–CeO x catalysts for the water-gas shift reaction (WGSR). Various synthesis protocols are shown to work; the best being the coprecipitation/gelation method, which suppresses the crystal growth of ceria during calcination by the incorporation of some platinum in the bulk oxide. Metallic platinum nanoparticles are not necessary for an active WGSR catalyst. Reaction light-off occurs at ∼120 °C, where Pt2+ species bound to ceria are still present. No activation period, and no hysteresis phenomena were found. During reaction at reducing conditions, some Pt is reduced, but it is reoxidizable. The stability of these low-content (<2 at.%) Pt/CeO x catalysts is high even in realistic reformate gas streams. To avoid cerium(III) hydroxycarbonate formation at room temperature-shutdown with water condensation, a small amount (∼1%) of gaseous oxygen is added to the reaction gas mixture. Cyclic shutdown/startup operation is thus possible without catalyst degradation.  相似文献   

8.
Heat of reaction and kinetic parameters were determined by differential scanning calorimetry for decomposition of hydrogen peroxide, reaction of hydrogen peroxide with lignocellulosic materials, glucose and pinitol, and for the reaction of the same materials with produced or introduced oxygen. The heat of decomposition of hydrogen peroxide obtained in N2 (720 cal/g H2O2) was in fair agreement with literature data, considering the different temperature and pressure conditions. The heats of reaction of hydrogen peroxide and lignocelluloses were higher when determined in N2 (1670–2500 cal/g H2O2) than in O2 (1450–2020 cal/g H2O2) atmosphere. The activation energy for decomposition of hydrogen peroxide amounted to 20.3 kcal/mol in N2 and 15.9 kcal/mol in O2 with frequency factors of 5.7 × 109 and 3.7 × 107 min?1, respectively. The activation energies for the reaction of hydrogen peroxide and lignocellulosic materials tested were similar and not influenced by the atmospheric composition, ranging overall between 19.7 and 22.4 kcal/mol. The corresponding frequency factors ranged between 2.77 × 109 and 2.23 × 1011.  相似文献   

9.
Yuying Shu 《Carbon》2005,43(7):1517-1532
A series of nickel, molybdenum, and tungsten metal phosphides deposited on a carbon black support (Ni2P/C, MoP/C, and WP/C) were synthesized by means of temperature-programmed reduction. The samples were characterized by BET surface area, CO uptake, X-ray diffraction (XRD), elemental analysis, and extended X-ray absorption fine structure (EXAFS) measurements. The activity of these catalysts was measured at 613 K and 3.1 MPa in a three-phase, packed-bed reactor for hydrodesulfurization (HDS) and hydrodenitrogenation (HDN) with a model liquid feed containing 500 ppm sulfur as 4,6-dimethyldibenzothiophene (4,6-DMDBT), 3000 ppm sulfur as dimethyl disulfide, and 200 ppm nitrogen as quinoline. The Ni2P/C catalyst was found to exhibit the best hydroprocessing performance based on equal CO chemisorption sites (70 μmol) loaded in the reactor. An optimum Ni loading for HDS and HDN activity was found as 1.656 mmol g−1 (11.0 wt.% Ni2P) which gave an HDS conversion of 99% and an HDN conversion of 100% at a molar space velocity of 0.88 h−1. These were much higher than those of a commercial Ni-Mo-S/γ-Al2O3 catalyst which gave an HDS conversion of 68% and an HDN conversion of 94%, and a previously reported best Ni2P/SiO2 catalyst which gave an HDS conversion of 76% and an HDN conversion of 92%. The use of carbon instead of silica as a support gave rise to other differences, which included smaller particle size, higher CO uptake, lessened retention of P on the support, and reduced sulfur deposition. The stability of the 11.0 wt.% Ni2P/C catalyst was also excellent with no deactivation observed over 110 h of time on stream. The activity and stability of the Ni2P/C catalyst were affected by the phosphorous content, both reaching a maximum with an initial Ni/P ratio of 1/2. EXAFS and elemental analysis of the spent samples indicated the formation of a surface phosphosulfide phase on the Ni2P, which was beneficial for hydrotreating activity, while the bulk structure of the phosphides was maintained during the course of reaction as revealed from the XRD patterns.  相似文献   

10.
The reaction route of the hydrogen electrode reaction on Ni, Rh, Au, and Pt at the reversible potential was investigated through the exchange reaction between deuterium and light water. Various reaction routes involving hydrogen adatom H(a) and adsorbed hydrogen molecule-ion H+2(a) as the reaction intermediates were considered. In order to single out the correct reaction route, these routes were examined, without assuming a priori the existence of a single rate-determining step, for satisfactory interpretation of the isotopic composition of the hydrogen gas during the exchange reaction. It is concluded that the reaction route consists of two consecutive steps,
and
, where B represents H2O or OH?.  相似文献   

11.
Film-type electrodes of hydrogen absorbing intermetallic compound and alloys, LaNi5, Ni0.11 Ti0.89, Ni0.50 Ti0.50 and Ni0.76 Ti0.24 were prepared by a flash evaporation method. The hydrogen electrode reaction characteristics of the LaNi5 and NiTi alloy films in 1 M NaOH are very similar to each other. The reaction proceeds via the Volmer-Tafel reaction route with mixed rate-determining characteristics. The exchange current densities of the constituent steps, as well as the overall reaction, are in the range of 10?6 A cm?2 (true). Surface analysis by an XPS technique has shown that La or Ti on the electrode surface exists as an electrocatalytically innert oxide of La2O3 or TiO2. Close similarities of these electrodes with pure Ni electrodes indicate that Ni is responsible for the electrocatalytic activity. No synergistic effect is thus noticeable.  相似文献   

12.
《Journal of Catalysis》2006,237(2):213-219
Tubular catalytic membranes (TMCs) active in the direct synthesis of hydrogen peroxide were prepared, characterized, and tested using methanol as the reaction medium. Low hydrogen peroxide selectivity was found when only palladium was used as a catalyst, whereas palladium/platinum bimetallic samples gave higher productivity and selectivity, with an optimum molar ratio of 18. The H2O2 decomposition rate is influenced by the feed gases. O2 improves H2O2 stability, whereas H2 causes hydrogen peroxide to decompose at a higher rate. The most likely decomposition pathway should be the reduction of H2O2 to water by H2. Bromide ion was used as a promoter and when used in excess (60 ppm) causes a decrease in overall catalytic activity.  相似文献   

13.
The exchange reaction between deuterium gas and light water was conducted with Ni, Rh, and Pt catalyst. Through analysis of isotopic composition of the gaseous hydrogen during the exchange reaction, the exchange rates of the individual steps, H2 ? 2H(a) and H(a) + B ? H+B + e, of the hydrogen electrode reaction were determined at various hydrogen pressures and solution pH's, where H(a) is a hydrogen adatom and B = H2O or OH?. The rates of the steps were widely different among the metals studied but, throughout these metals, the hydrogen pressure dependence of the rate of the first step was with the power of 1·0–0·9, and that of the second step, 0·2–0·5. Both the rates were independent of solution pH. Generally, the former step is virtually rate-determining under low hydrogen pressures, whereas the latter takes over the role with increasing hydrogen pressure. Under ordinary pressures, the first step is virtually rate-determining on Pt but neither step is singly rate-determining on Rh and Ni.  相似文献   

14.
To obtain an idea of the magnitudes of the ozone loss rates rO3 in practical applications of ozone, an overall determination of the ozone decay profiles and rate constants was carried out in four different systems. These systems resemble different conditions for industrial application of ozone and the peroxone process, such as in the field of micro electronics, drinking water purification, disinfection, etc. Therefore, the behavior of ozone was monitored in the pH range from 4.5 to 9.0, in pure water and phosphate buffered systems in absence and presence of small amounts of hydrogen peroxide (10?7 M to 10?5 M H2O2). First the reproducibility of the ozone decay profiles was checked and from the various kinetic formalism tests, the reaction order 1.5 for the ozone decay rate has been selected. As expected, hydrogen peroxide increases the decay rates. In pure systems, added concentrations of 10?7M H2O2 already cause a remarkable acceleration of the ozone decay in the acidic and neutral pH range compared to the pure systems. However for alkaline pH conditions almost no effect of the low hydrogen peroxide concentrations was noticed. Contradictory to literature data, in the absence of hydrogen peroxide, ozone displays faster decays in the buffered systems of low ionic strength of 0.02 compared to pure water. This acceleration is more pronounced for acidic pH conditions. Low concentrations of phosphate may indeed accelerate the ozone decay in the presence of organic matter. Adding H2O2 concentrations below 10?5M to phosphate buffered solutions has a negligible effect on the ozone decay rate compared with pure water systems, except for pH 7. It appears that phosphate masks the effect of hydrogen peroxide below 10?5 M as tested here. Thus the application of AOP's by adding low concentrations of hydrogen peroxide is not well feasible in the presence of phosphate buffers in pure water systems.  相似文献   

15.
Zinc oxide catalyses both methanol synthesis and the forward and ‘everse water-gas shift reaction (f- and r- WGSR). Copper also catalyses both reactions, but at lower temperatures than ZnO. Presently the combination of Cu and ZnO stabilized by Al2O3 is the preferred catalyst for methanol synthesis and for the f- and r- WGSR. On Cu, the mechanism of methanol synthesis is by hydrogenation of an adsorbed bidentate formate [1] (the most stable adsorbed species in methanol synthesis), while the f- and r- WGSR proceeds by a redox mechanism. The f-WGSR proceeds by H2O oxidizing the Cu and CO, reducing the adsorbed oxide and the r-WGSR proceeds by CO2 oxidising the Cu and H2, reducing it [2–5]. Here we show that the mechanisms of both reactions are subtly different on ZnO. While methanol is shown to be formed on ZnO through a formate intermediate, it is a monodentate formate species which is the intermediate; the f- and r-WGS reactions also proceed through a formate – a bidentate formate - in sharp contrast to the mechanism on Cu.  相似文献   

16.
The oxidative dehydrogenation and cracking of ethane and propane over LiDyMg mixed oxides is reported. High yields of olefins and only moderate formation of carbon oxides was observed. Both are primary products that hardly interconvert under the reaction conditions used. Addition of chloride increases the rate of reaction, while slightly decreasing the selectivity to olefins. The addition of carbon dioxide strongly decreases the rate of reaction, the negative order of 0.5 indicating that two active Li+sites are blocked by the adsorption of one CO2molecule. The reaction proceeds at low oxygen pressure primarily via elimination of dihydrogen, while at higher oxygen partial pressure the hydrogen elimination occurs via water formation. It is speculated that dehydrogenation and cracking involve Li+and a rather nucleophilic oxygen site.  相似文献   

17.
The influence of crystal phase of zirconia on the performance of Au/ZrO2 catalysts for low temperature water gas shift reaction was investigated. Au/ZrO2 catalysts with pure tetragonal and monoclinic phases of ZrO2 were prepared by the deposition-precipitation method with similar gold loading and dispersion. It was found that the Au/m-ZrO2 catalyst showed much higher activity than that of the Au/t-ZrO2 catalyst, which could be attributed to the higher CO adsorption capacity of the Au/m-ZrO2 catalyst. The chemical state of gold that was strongly related to the pretreatment atmosphere also played an essential role in determining the catalytic activity for water gas shift reaction. Hydrogen and/or helium pretreated samples only contained Au0 species and exhibited higher activity than that of the sample pretreated with oxygen-containing atmosphere which resulted in the co-existence of Au0 and Au+ species. FTIR study further revealed that the formate species formed by the reaction of the adsorbed CO on gold nanoparticle with the hydroxyl groups on the surface of m-ZrO2 acted as the most important intermediates for the water gas shift reaction.  相似文献   

18.
The hydrodenitrogenation (HDN) of piperidine, decahydroquinoline (DHQ) and orthopropylaniline (OPA) has been studied over NiMo(P)/Al2O3 catalysts at 593 K and 3.0 MPa in order to understand the effect of phosphorus on the elementary HDN reaction steps. Phosphorus exhibited a negative effect on the HDN of piperidine and DHQ, both on the C-N bond cleavage reaction and on the subsequent hydrogenation reaction of alkene to alkane. A P/Al2O3 catalyst showed no HDN activity at all, neither with piperidine, nor with DHQ. A positive effect of phosphorus was observed in the HDN of OPA, where hydrogenation of the aromatic ring is needed and is rate limiting. It is suggested that introduction of phosphorus to NiMo/Al2O3 catalysts on the one hand decreases the available support surface area, and as a consequence the dispersion of the Ni-Mo-S phase and thus the capacity for C-N bond breaking and olefin hydrogenation. On the other hand, phosphorus induces either new or more active sites for the hydrogenation of aromatics.  相似文献   

19.
In this study, ammonia is presented as a feasible fuel for solid oxide fuel cells (SOFCs). Ammonia has several interesting features as fuel due to low‐production cost, high‐energy density and, focusing on fuel cells and hydrogen application, ammonia is an excellent H2 carrier thanks to high value of volumetric and gravimetric densities. The paper reports experimental test performed to evaluate the feasibility of NH3 directly fed to a 50 cm2 single cell SOFC. A test plan was developed to compare pure ammonia with an equivalent mix of ammonia, nitrogen, and hydrogen and the study of temperature and voltage values strongly indicates that a two stage oxidation of ammonia can be predicted and a previous cracking reaction occurs in the cell due to the nickel catalytic contribution. The study of temperatures and of heat flows show how the cell is cooled down to lower temperature because of heat adsorbed by the reaction and by flow mix entering the anode. The study shows also how for operative temperatures below 800 °C the cracking reaction takes place in the cell active area. Efficiency test demonstrates that the cell can operate at 300 mW cm–2 and 30% efficiency based on ammonia LHV.  相似文献   

20.
Hydrodenitrogenation (HDN) of quinoline using a Mo-based dispersed catalyst was studied in a batch reactor using H2 generated in situ via the water gas shift reaction. In situ H2 was found to be more active for HDN than externally supplied H2. Both water and H2S have an effect on HDN.  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