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 15 毫秒
1.
In this study, the decomposition of methanol into the CO and H species on the Pd/tungsten carbide (WC)(0001) surface is systematically investigated using periodic density functional theory (DFT) calculations. The possible reaction pathways and intermediates are determined. The results reveal that saturated molecules, i.e., methanol and formaldehyde, adsorb weakly on the Pd/ WC(0001) surface. Both CO and H prefer three-fold sites, with adsorption energies of −1.51 and −2.67 eV, respectively. On the other hand, CH3O stably binds at three-fold and bridge sites, with an adsorption energy of −2.58 eV. However, most of the other intermediates tend to adsorb to the surface with the carbon and oxygen atoms in their sp3 and hydroxyl-like configurations, respectively. Hence, the C atom of CH2OH preferentially attaches to the top sites, CHOH and CH2O adsorb at the bridge sites, while COH and CHO occupy the three-fold sites. The DFT calculations indicate that the rupture of the initial C–H bond promotes the decomposition of CH3OH and CH2OH, whereas in the case of CHOH, O–H bond scission is favored over the C–H bond rupture. Thus, the most probable methanol decomposition pathway on the Pd/WC(0001) surface is CH3OH → CH2OH → trans-CHOH → CHO → CO. The present study demonstrates that the synergistic effect of WC (as carrier) and Pd (as catalyst) alters the CH3OH decomposition pathway and reduces the noble metal utilization.  相似文献   

2.
In this work, mechanistic aspects of the partial oxidation of methanol (POM) to hydrogen and carbon dioxide over Cu/ZnO catalysts have been investigated. The data obtained with different catalyst compositions and different Cuo metal surface areas showed that the reaction depends on the presence of both the phases ZnO and Cuo. On the other hand, for catalysts with Cu concentrations in the range 40-60 wt%, the copper metal surface area seems to be the main factor determining the reaction rate. Kinetic isotope effects using CH3OH and CH3OD showed that both C–H and O–H bonds are at least partially involved in the rate-limiting step. TPD experiments with pure Cuo, pure ZnO and the catalyst Cu/ZnO showed that methanol can be activated by both ZnO and copper. On the ZnO surface methanol can form intermediates which in the presence of copper might react and desorb more easily probably via a reverse spillover process. The isotopic product distribution of H2, HD, D2, H2O, HDO and D2O in the temperature-programmed reaction of CH3OD revealed a slight enrichment of the products with H, suggesting that during methanol activation on the ZnO some of the D atoms might be retained by the support. The effect of oxygen partial pressure suggests that oxygen atoms on the copper surface strongly promote methanol activation and H2 and CO2 formation. It is proposed that oxygen atoms participate in methanol activation by the abstraction of the hydroxyl H atom to form methoxide and OHsurf. This OHsurf species rapidly loses H to the surface regenerating the Osurf.  相似文献   

3.
Based on the information from GC-MS on-line measurement and thermodynamic analysis, the reaction network of gas-phase hydroxylation of benzene with nitrous oxide over Fe-ZSM-5 zeolite was systematically investigated. The main reactions and side reactions were identified, and a kinetic reaction network was proposed as follows: benzene+N2O→phenol→CO/CO2. According to the mechanism, the experimental results were interpreted reasonably. The hydroxylation kinetic experiments were carried out in an isothermal integral microreactor under the conditions of n(benzene)/n(N2O)=8–12, T=663–763 K and atmospheric pressure. Based on the reaction network proposed, the parameters in the rate model of power-law were estimated by means of Gauss-Newton optimal method with the Levenberg-Marquardt modifications, and the results were in good agreement with the experimental data.  相似文献   

4.
Summary The relative efficiencies of organic, polar solvents and of solvent-water pairs for use in the extraction of gossypol and related compounds from cottonseed flakes were determined in a specially devised glass laboratory extractor. Of the solvents tested a butanone-water pair containing 10% of water by volume was the most effective, and chlorine-substituted hydrocarbons were the least effective. Under equilibrium conditions maximum extraction of gossypol was obtained with a butanone solvent containing 2.5% of water by weight. The rate of extraction of gossypol from cottonseed meal with butanone-water pairs increased with increase in the amount of water in the system and with increase in temperature of the extraction system. The greater amounts of water in the extraction system resulted in swelling and packing of the flakes and in a decrease in extraction efficiency. Flakes extracted at 26°C. contained 0.08% free gossypol and those extracted at 71°, 0.054%. This decrease may be due, in part, to the reaction of gossypol with the protein to form bound gossypol. Report of a study carried on under the Research and Marketing Act of 1946. This paper is No. 9 in the series on “Processing of Cottonseed” from the Southern Regional Research Laboratory. References to other papers in this series are: J. Am. Oil Chem. Soc.24, 97–108 (1947); J. Am. Oil Chem. Soc.24, 276–283 (1947); J. Am. Oil Chem. Soc.24. 362–369 (1947); J. Am. Oil Chem. Soc.26, 28–34 (1949); Oil Mill Gaz.54 (2), 12–15 (1949); Cotton Gin and Oil Mill Press51 (9), 18–20 (1950); Official Proc. Natl. Cottonseed Products Asso,55, 32–34, 36 (1951); and J. Am. Oil Chem. Soc. (in press), “The Effect of Screw Press and Hydraulic Press Processing Conditions on Pigment Glands of Cottonseed”, by D. M. Batson. F. H. Thurber, and A. M. Altschul. One of the laboratories of the Bureau of Agricultural and Industrial Chemistry, Agricultural Research Administration, U. S. Department of Agriculture.  相似文献   

5.
Steam reforming of ethanol, C2H5OH+H2O→2CO+4H2, was carried out over Co/Al2O3, Co/SiO2, Co/MgO, Co/ZrO2 and Co/C. The properties of the Co catalysts were greatly affected by the supports. Co/Al2O3 exhibited the highest selectivity for steam reforming of ethanol by suppression of methanation of CO and decomposition of ethanol. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

6.
Structural change of Cu/ZnO by reduction of ZnO in Cu/ZnO with methanol   总被引:1,自引:0,他引:1  
The reducibility of ZnO was investigated in the temperature range of 523–623 K in a stream of a reducing agent such as H2, CO, and methanol. ZnO was reduced only in the presence of copper in the vicinity of ZnO with CO and methanol, but it was not reduced with H2. Methanol was a stronger reducing agent in the reduction of ZnO than CO, while CO was stronger in the reduction of CuO than methanol. Two types of brass were observed resulting from the reduction of ZnO in the Cu/ZnO sample by XRD. Zanghengite brass started to be formed at 573 K in addition to α-brass which was observed at the temperature above 523 K in the temperature range of 523–623 K during the ZnO reduction with methanol. The carbon monoxide chemisorption showed that the copper surface areas decreased during the reduction of ZnO with methanol. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

7.
Surface science studies of heterogeneous catalysis use model systems ranging from single crystals to monodispersed nanoparticles in the 1–10 nm range. Molecular studies reveal that bond activation (C–H, H–H, C–C, C≡O) occurs at 300 K or below as the active metal sites simultaneously restructure. The strongly adsorbed molecules must be mobile to free up these sites for continued turnover of reaction. Oxide–metal interfaces are also active for catalytic turnover. Examples using C–H and C = O activation are described to demonstrate these properties. Polymerization catalysis demonstrates a strong dependence upon catalyst surface structure, which allows for the selectivity to be tuned by the choice of Ziegler-Natta surface preparation. Novel preparation methods of model catalyst arrays in two and three dimensions are opening the door to a complete understanding of catalytic reaction selectivity.  相似文献   

8.
The effect of carbon monoxide (CO) co-adsorption on the dissociation of water on the Ni(111) surface has been studied using density functional theory. The structures of the adsorbed water molecule and of the transition state are changed by the presence of the CO molecule. The water O–H bond that is closest to the CO is lengthened compared to the structure in the absence of the CO, and the breaking O–H bond in the transition state structure has a larger imaginary frequency in the presence of CO. In addition, the distances between the Ni surface and H2O reactant and OH and H products decrease in the presence of the CO. The changes in structures and vibrational frequencies lead to a reaction energy that is 0.17 eV less exothermic in the presence of the CO, and an activation barrier that is 0.12 eV larger in the presence of the CO. At 463 K the water dissociation rate constant is an order of magnitude smaller in the presence of the CO. This reveals that far fewer water molecules will dissociate in the presence of CO under reaction conditions that are typical for the water-gas-shift reaction.  相似文献   

9.
Methanol steam reforming (MSR) and preferential CO oxidation (PROX) were studied with the view of improving the generation of H2-rich gases. In MSR, conventional catalysts of methanol synthesis were tested, various Cu-based catalysts were prepared and studied. A theoretic kinetic model (based on the reaction mechanism established using independent methods [1]) is developed and checked out. PROX was studied over various Ru/Al2O3 catalysts using a flow “quasi-adiabatic” reactor. On-line recording of gas temperature in the catalyst bed and CO residual concentration at varied reaction conditions allowed to observe ignition and extinction of the catalyst surface and the transition states of the process. It is shown that in the ignition mode a sharp decrease in CO residual concentration can be achieved. The combination of proposed catalyst and the control of the macrokinetic regime of PROX allows high degree of CO removal from gaseous mixtures produced by MSR. Residual CO content in a H2-rich gaseous mixture can be lowered to < 15 ppm at GHSV∼100 m3/(kg cat)/h and O2/CO ratio of 1. Obtained data show the possibility of designing a high-throughput set-up for generation of H2-rich gases from methanol with one-step cleaning from the CO impurity.  相似文献   

10.
The intrinsic kinetics of the three-phase dimethyl ether (DME) synthesis from syngas over a bi-functional catalyst has been investigated in a agitated slurry reactor at 20–50 bar, 200–240 °C and H2/CO feed ratio from 1 to 2. The bi-functional catalyst was prepared by physical mixing of CuO/ZnO/Al2O3 as methanol synthesis catalyst and H-ZSM-5 as methanol dehydration catalyst. The three reactions including methanol synthesis from CO and H2, methanol dehydration and water gas shift reaction were chosen as the independent reactions. A kinetic model for the combined methanol and DME synthesis based on a methanol synthesis model proposed by Graaf et al. [G.H. Graaf, E.J. Stamhuis, A.A.C.M. Beenackers, Kinetics of low pressure methanol synthesis, Chem. Eng. Sci. 43 (12) (1988) 3185; G.H. Graaf, E.J. Stamhuis, A.A.C.M. Beenackers, Kinetics of the three-phase methanol synthesis, Chem. Eng. Sci. 43 (8) (1988) 2161] and a methanol dehydration model by Bercic and Levec [G. Bercic, J. Levec, Intrinsic and global reaction rate of methanol dehydration over γ-Al2O3 pellets, Ind. Eng. Chem. Res. 31 (1992) 399–434] has been fitted our experimental data. The obtained coefficients in equations follow the Arrhenius and the Van’t Hoff relations. The calculated apparent activation energy of methanol synthesis reaction and methanol dehydration reaction are 115 kJ/mol and 82 kJ/mol, respectively. Also, the effects of different parameters on the reactor performance have been investigated based on the presented kinetic model.  相似文献   

11.
The tracer method was used to numerically study the effect of nitric oxides (NO and NO2) on the oxidation of rich hydrogen-air mixtures during adiabatic self-ignition at low and high initial temperatures and a pressure of 0.1 MPa. At low temperatures, the added NO interacts with HO2 to form NO2, and NO2 then interacts with H to form NO. When NO2 is added at the same temperatures, a two-stage mechanism takes place: NO formed by the reaction NO2 + H is not involved in the reaction until NO2 is almost completely consumed. In the temperature range 900–1200 K, NO2 inhibits self-ignition through participation in the reaction with H, leading to the replacement of part of the completely branched chain H → (O, OH) → 3H by the unbranched chain H → OH → H. At low initial temperatures, NO effectively promotes hydrogen oxidation due to replacement of the unbranched chain H → HO2 → H2O2 → OH → H by a chain with branching.  相似文献   

12.
The ring-opening mechanism influencing effect of a major product in the cyclohexene oxide–D2 system was investigated over a Cu/SiO2 catalyst. This product is cyclohexene, thus, the hydrogenative ring opening of a 1:1 cyclohexene oxide–cyclohexene mixture was studied in the presence of D2 at 403 K in a closed circulation reactor. It was found that the mechanism of single C–O scission was not affected, but that of the double C–O scission was changed. Simultaneous bond cleavage was the major route of ring opening in the additive-free system and it became consecutive on cyclohexene addition. Added cyclohexene was hydrogenated with a very low rate, but it transformed the surface of the catalyst and, thus, facilitated the change in the mechanism. An explanation concerning the seemingly anomalous lack of deuterium in a product (cyclohexane) not seen in the additive-free system is also suggested. This revised version was published online in November 2006 with corrections to the Cover Date.  相似文献   

13.
Using TiO2 as carrier, CuO/TiO2 catalysts with different CuO loading were prepared by the impregnation method. The catalytic activities in NO+CO reaction were examined with a micro-reactor gas chromatography reaction system and the methods of TPR, XPS and NO-TPD. It was found that the catalytic activities were affected by pretreatment atmosphere, i.e. H2 atmosphere > reduction–reoxidation > 10%CO/He > reaction gas (fresh sample). NO decomposition was better by low-valence Cu species than by high-valence Cu species, i.e. Cu0>Cu+>Cu2+. The XPS results indicated that Cu species on CuO/TiO2 were Cu0, Cu+, normal Cu2+(Cu2+(I)) and chain-structured Cu2+(Cu2+(II)) as –Cu–O–Ti–O–. The activities of Cu2+(II) were much higher than that of Cu2+(I), but both species were very unstable in the reaction atmosphere and easily reduced by CO, which accounted for the variable activities of fresh catalysts with increasing reaction temperature. In NO+CO reaction, the redox process was a cycle of Cu+–Cu2+(I) at low reaction temperature but was a cycle of Cu0–Cu+ at high reaction temperature. As shown by NO-TPD, high catalytic activities could be attributed to the following factors, e.g. oxygen caves on the catalyst’s surface after pretreatment with H2 and reduction–reoxidation, formation of Cu0 after pretreatment with H2, and increment of Cu species dispersion and formation of Cu2+(II) after pretreatment with reduction–reoxidation.  相似文献   

14.
Estimate of methane production from rumen fermentation   总被引:1,自引:0,他引:1  
A method for the assessment of CH4 emission from dairy cows, based on in vitro volatile fatty acids (VFA) production, is described. 15 energy rich feedstuffs, 12 protein feedstuffs and 15 forages were in vitro fermented using rumen fluid as inoculum. Methane production for each feedstuff was calculated from net concentration of volatile fatty acids after 24 hours of fermentation according to the following stoechiometry: Glucose + 2 H2O → 2 acetate + 2 CO2 + 4 H2; Glucose + 2 H2 → 2 propionate + 2 H2O ; Glucose → 1 butyrate + 2 CO2 + 2 H2; CO2 + 4 H2 → CH4 + 2 H2O and taking in account the `dilution rate' of feedstuffs, estimated according to the Cornell University model. These data were used to estimate the CH4 production in cows (live weight 650 kg, annual milk production:7550 kg, fat content 3.8%). Calculated annual methane emissions based on in vitro trials were: 182.74 kg CH4 /cow or 137.05 kg C–CH4 /cow (diet with corn silage); 182.56 kg CH4 /cow or 136.92 kg C–CH4/cow (diet without corn silage). If compared with estimastes obtained from IPPC (1996) detailed methodology, the above estimates is 35% higher. A critical evaluation of reliability of in vitro data is given. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

15.
The reduction of nitric oxide by reaction with non-hydrocarbon fuels under reducing conditions at comparatively higher temperature has been studied with a detailed chemical kinetic model. The reaction mechanism consists of 337 elementary reactions between 65 chemical species based on the newest rate coefficients. The experimental data were adopted from previous work. Analyses by comparing existing experimental data with the modeling predictions of this kinetic mechanism indicate that, at comparatively high temperature, apart from the reaction path NO→HNO→NH→N2, NO+N→N2 is also prominent. In the presence of CO, NO is partly converted to N by reaction with CO. Based on present model, the reduction of NO at high temperature, which was usually underestimated by previous work, can be improved to some extent. This work was presented at the 7 th China-Korea Workshop on Clean Energy Technology held at Taiyuan, Shanxi, China, June 26–28, 2008.  相似文献   

16.
The reaction kinetics of the dimethyl ether synthesis from hydrogen-rich synthesis gas over bi-functional catalyst was investigated using an isothermal integral reactor at 220–260°C temperature, 3–7 MPa pressure, and 1,000–2,500 mL/g·h space velocity. The H2/CO ratio of the synthetic gas was chosen between 3 : 1 and 6 : 1. The bi-functional catalyst was prepared by physically mixing commercial CuO/ZnO/Al2O3 and γ-alumina, which act as methanol synthesis catalyst and dehydration catalyst, respectively. The three reactions, including methanol synthesis from CO and CO2 as well as methanol dehydration, were chosen as independent reactions. The Langmuir-Hinshelwood kinetic models for dimethyl ether synthesis were adopted. Kinetics parameters were obtained using the Levenberg-Marquardt mathematical method. The model was reliable according to statistical and residual error analyses. The effects of different process conditions on the reactor performance were also investigated.  相似文献   

17.
Clavibacter sp. ALA2 converts linoleic acid into many novel oxygenated products including hydroxy FA and tetrahydrofuranyl unsaturated FA (THFA). One of them was tentatively identified by GC-MS as 12,13,16-trihydroxy-9(Z)-octadecenoic acid (12,13,16-THOA) (Hou, C.T., H.W. Gardner, and W. Brown, J Am. Oil Chem. Soc. 78∶1167–1169, 2001). We have separated and purified 12,13,16-THOA from its isomer, 12,13,17-THOA, by silica gel column chromatography and by preparative TLC. Its structure was then confirmed by proton and 13C NMR analyses. Purified 12,13,16-THOA was used as a substrate to study the biosynthesis of THFA. Within 24 h of incubation, cells of strain ALA2 converted 12,13,16-THOA to both 12-hydroxy-13,16-epoxy-9(Z)-octadecenoic acid (12-hydroxy-THFA) and 7,12-dihydroxy-13,16-epoxy-9(Z)-octadecenoic acid (7,12-dihydroxy-THFA). The relative abundance of 7,12-dihydroxy-THFA increased with incubation time, whereas that of 12,13,16-THOA and of 12-hydroxy-THFA decreased. Therefore, the biosynthetic pathway of THFA from linoleic acid by strain ALA2 is as follows: linoleic acid→12,13-dihydroxy-9(Z)-octadecenoic acid→12,13,16-THOA→12-hydroxy-THEA→7,12-dihydroxy-THFA.  相似文献   

18.
A numerical modeling study of the propagation of a laminar flat homogeneous gas flame has shown that in a hydrogen-air flame, a rapid increase in the concentration of OH radicals begins in the range of low temperatures and the concentration profile has two maxima. The first maximum in the low-temperature region of the front is related to the diffusion of H atoms, formation of HO2 radicals, and the quadratic branching reaction H + HO2 → OH + OH. The second maximum in the OH concentration profile is due to the classical high-temperature branching reactions H + O2 → OH + O and O + H2 → OH + H. __________ Translated from Fizika Goreniya i Vzryva, Vol. 43, No. 2, pp. 3–9, March–April, 2007.  相似文献   

19.
Hollow silica nano spheres containing Rh, Ir or Ru metal particles were synthesized by Rh(NH3)6Cl3 aq, Ir(NH3)3Cl3 aq or Ru(NH3)6Cl3 aq/NP-6/cyclohexane reversed micelle system. Hydrolysis of TEOS surrounding metal ammine complex crystals inside the micelle caused the formation of the hollow, which contained small metal particles inside and tiny metal clusters in the silica network. The amounts of H2 adsorption over Rh and Ir nanocomposites were two to three times more in the cases of hollow-SiO2 catalysts compared with those of non-hollow ones, suggesting the occlusion of hydrogen inside the hollows of Rh–SiO2 or Ir–SiO2. CO molecules could also permeate into the silica wall and be adsorbed on the metal clusters in the silica wall after 573 K pretreatment. Especially in the case of Ru nanocomposite the amount of adsorbed CO was much more than that of H2, suggesting some unique character of Ru metal nanoparticles. After 773 K pretreatment, however, the amount of CO(a) decreased drastically to less than 1/10 of H(a), indicating the densification of Si–O–Si bonds and the formation of ultra-micropores in the silica wall where only H2 can selectively permeate. Selective formation of methane was observed in the CO–H2 reaction over these nanocomposite catalysts, provably because of the higher concentration of hydrogen inside the hollow and silica network.  相似文献   

20.
Substituent effects have been used to probe the nature of the transition state to catalytic carbon–halogen bond breaking. Kinetics measurements have determined the activation energies (E act to C–Cl bond breaking on the Pd(111) surface and C–I bond breaking on the Pd(111) and Ag(111) surfaces. These barriers have been measured using alkyl halides with varying degrees of fluorine substitution. The activation energies have been correlated with the inductive or field substituent constants (σF) of the fluorinated alkyl groups in order to determine reaction constants (E act=E0+ρσF) for the dehalogenation reactions. In all three cases it has been found that the barriers are insensitive to inductive substituent effects and the reaction constants are all relatively small: ρ= −0.5± 1.0 kcal/mol for C–Cl cleavage on Pd(111), ρ= −0.3±0.8 kcal/mol for C–I cleavage on Pd(111), and ρ= −2.9±0.4 kcal/mol for C–I cleavage on Ag(111). This implies that the transition state for dehalogenation is homolytic and occurs early in the reaction coordinate. The implications of this result are discussed for catalytic dehalogenation processes such as hydrodechlorination. This revised version was published online in July 2006 with corrections to the Cover Date.  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