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 562 毫秒
1.
2.
Alumina-supported Pd model catalysts were prepared by Pd evaporation onto a thin alumina film grown on a NiAl(110) substrate. Adsorption and co-adsorption of ethene, CO and hydrogen on Pd/Al2O3/NiAl(110) covered by carbon species, formed by ethene dehydrogenation at 550 K, was studied by temperature programmed desorption (TPD). TPD results show that carbon deposits do not prevent adsorption but inhibit dehydrogenation of di- bonded ethene. Carbon species suppress CO adsorption in the highly coordinated sites and also suppress the formation of hydrogen ad-atoms on the surface. The ethene hydrogenation reaction performed by co-adsorption of hydrogen and ethene is inhibited by the presence of carbon deposits. The inhibition is independent of particle size studied (1-3 nm). The effects are rationalized in terms of a site-blocking behavior of carbon species occupying highly coordinated sites on the Pd surface.  相似文献   

3.
The hydrogenation of ethene is an important reaction in heterogeneous catalysis and, despite its apparent simplicity, many aspects of the reaction mechanism remain unclear. By contrast, the mechanism using homogeneous catalysts such as Wilkinson's catalyst [(RhCl(PPh3)3] is thought to be well understood. To allow a comparison between the homogeneous and heterogeneous reactions we have studied ethene/hydrogen interactions on the (111) plane of rhodium in the temperature range 160–500 K. Under UHV conditions no catalytic reaction was detected. However, we have been able to observe stoichiometric hydrogenation and exchange in the chemisorbed layer. A mixed adlayer of either ethene/deuterium (or perdeuteroethene and hydrogen) was formed at ca. 160 K, and allowed to warm up. From previous spectroscopic studies, ethene is adsorbed at 165 K as partially rehybridised, bonded species with a C-C bond order of ca. 1.5, similar to ethene in Zeise's salt. At 190–210 K we observe coincident desorption of undeuterated ethene — the major species — together with much smaller quantities of deuterated ethane and partially deuterated ethenes. The influence of both hydrogen and ethene pre-coverage has been studied as has the relative extent of hydrogenation and exchange. The ethane formation results parallel those reported by other authors on Pd(110) and Pt(111) and Pt(110). We propose that on all three metals both hydrogenation and exchange follow the same pathway, with a common intermediate for exchange and hydrogenation. This isa weakly held, bonded species formed during the desorption process, which can be convertedreversibly into an adsorbed ethyl species. A detailed comparison indicates that the mechanism of heterogeneous hydrogenation closely parallels that in the homogeneous phase.  相似文献   

4.
The pyrolysis of benzonitrile under nitrogen at atmospheric pressure has been studied at temperatures of 823–873 K in a flow reactor. The results demonstrate that conversion to hydrogen cyanide occurs directly by a free radical mechanism. The dominant products detected are hydrogen cyanide, monocyanodiphenyl, benzene, dicyanodiphenyl and dicyanobenzene. Reaction orders and activation energies have been determined for product formation. A reaction scheme involving three competing chain reactions in the gas phase with chain carriers H., C6H and .C6H4CN is proposed to explain the observed kinetics. A mechanism is advanced for the formation of significant quantities of polymer, consistent with infra-red spectra and elemental analysis.  相似文献   

5.
The selective reduction of nitrogen dioxide and nitrogen monoxide by olefins (ethene, propene) has been studied over two different -aluminium oxides in the temperature range 473–873 K. Nitrogen dioxide was reduced more effectively than nitrogen monoxide with both, ethene and propene, as a reductant. At temperatures exceeding 700 K, ammonia was formed as a by-product over one type of alumina. Concentrations in the range 30–40 ppm were determined for propene in combination with both, NO and NO2, while no ammonia was produced with ethene as a reductant. In addition, significant formation of hydrogen cyanide up to 70 ppm was observed with propene over both aluminium oxides starting from either NO or NO2. In contrast, hydrogen cyanide formation remained below 10 ppm with ethene as a reductant. Nitrous oxide formation did not exceed 10 ppm for all investigations. The results show that for alumina catalysts ethene is a more suitable reductant than propene due to its lower tendency to form undesired by-products.  相似文献   

6.
Isolated atoms of carbon evaporated on to Pt(111) react with hydrogen atT170 K to form methine species, characterized with vibrational modesv(CH) at 2960 and (CH) at 800 cm–1. The high reactivity ofC ads is in line with their ability to take part as intermediates in the metanation reaction. CHads species are stable up toT 500 K; further heating leads to their dissociation accompanied by H2 desorption and formation of unreactive graphite-like islands.  相似文献   

7.
The bonding of the oxygen species held on a Ag/-Al2O3 catalyst has been studied by temperature programmed desorption and their reactivity in ethene epoxidation by temperature programmed reduction using ethene as the reductant. The Ag/-Al2O3 catalyst was produced by the thermal decomposition of a Ag oxalate/-Al2O3 precursor. Oxygen desorbs from this Ag/-Al2O3 catalyst in two states, one (peak maximum temperature 520 K) having a desorption activation energy of 140 kJ mol–1 – oxygen desorbing from Ag(111), and one (peak maximum temperature 573 K) having a desorption activation energy of 155 kJ mol–1 – oxygen desorbing from a highly stepped or defected Ag surface. Temperature programmed reduction of the two oxygen states existing on the surface of the Ag/-Al2O3 catalyst using ethene as the reductant produced two peaks at 373 and 473 K in which ethene epoxide and CO2 evolved coincidently. The peak at 373 K derives from the reduction of oxygen atoms adsorbed on Ag(111). The higher temperature peak (473 K) corresponds to the reduction of oxygen atoms adsorbed on highly stepped or defected Ag surface. The selectivity to ethene epoxide for the 373 K peak is ~ 57%, while that of the 473 K peak is 34%. The coincident evolution of ethene epoxide and CO2 shows that the selective and unselective reaction pathways have a common surface intermediate – probably an oxametallacycle. The higher selectivity of the oxametallacycle formed by the bonding of ethene to the weaker Ag-O bond is considered to result from its having a lower activation energy to cyclisation than that produced by ethene bonding to the higher Ag-O bond.  相似文献   

8.
Shale oil, obtained from an in-situ oil shale experiment, from the Green River formation in Southwestern Wyoming was thermally fractionated into naphtha, light distillate, heavy distillate and residue fractions. The naphtha and light distillate fractions were further separated into saturates, olefins and aromatic subfractions. 1H- and 13C-n.m.r. spectra and mass spectral data were obtained for the saturated hydrocarbons of the naphtha and light distillate fractions. Resonances in the n.m.r. spectra were assigned to normal alkanes and to methyl- and dimethyl-branched alkanes. The composition as determined by n.m.r. of the naphtha saturates was found to be ≈82% straight-chain alkanes with an average carbon-chain-length of ≈C11 and ≈18% branched/cyclo-alkanes. The composition of the light distillate saturates was found to be ≈ 69% straight-chain alkanes with an average carbon-chain-length of ≈C15and ≈31% branched/cyclo-alkanes. The dimethyl-branched alkanes in both saturate fractions were proposed to have the molecular substructure of saturated isoprenoids. In the naphtha saturates the isoprenoid substructure
is evident. However, in the light distillate saturates, both isoprenoid substructures,
and
, are evident. 13C spin-lattice re-laxation times also were determined for the dominant resonances observed in the spectra of the naphtha and light distillate saturates. Relaxation times for the molecular species in the naphtha saturate fraction were observed to be longer than those observed for the light distillate saturate fraction. It was found that the ratio of the overall average relaxation times for the naphtha and light distillate saturate fractions corresponded to the inverse ratio of the average molecular weights of each fraction as determined from average alkane carbon-chain-length determination. Intermolecular (segmental) motion of the carbon chain of the straight-chain alkanes in both saturate fractions is also evident from the relaxation time measurements.  相似文献   

9.
C2H6 reactions with O2 only form CO2 and H2O on dispersed Pt clusters at 0.2–28 O2/C2H6 reactant ratios and 723–913 K without detectable formation of partial oxidation products. Kinetic and isotopic data, measured under conditions of strict kinetic control, show that CH4 and C2H6 reactions involve similar elementary steps and kinetic regimes. These kinetic regimes exhibit different rate equations, kinetic isotope effects and structure sensitivity, and transitions among regimes are dictated by the prevalent coverages of chemisorbed oxygen (O*). At O2/C2H6 ratios that lead to O*-saturated surfaces, kinetically-relevant CH bond activation steps involve O*O* pairs and transition states with radical-like alkyls. As oxygen vacancies (1) emerge with decreasing O2/alkane ratios, alkyl groups at transition states are effectively stabilized by vacancy sites and CH bond activation occurs preferentially at O** site pairs. Measured kinetic isotope effects and the catalytic consequences of Pt cluster size are consistent with a monotonic transition in the kinetically-relevant step from CH bond activation on O*O* site pairs, to CH bond activation on O** site pairs, to O2 dissociation on ** site pairs as O* coverage decrease for both C2H6 and CH4 reactants. When CH bond activation limits rates, turnover rates increase with increasing Pt cluster size for both alkanes because coordinatively unsaturated corner and edge atoms prevalent in small clusters lead to more strongly-bound and less-reactive O* species and lower densities of vacancy sites at nearly saturated cluster surfaces. In contrast, the highly exothermic and barrierless nature of O2 activation steps on uncovered clusters leads to similar turnover rates on Pt clusters with 1.8–8.5 nm diameter when this step becomes kinetically-relevant at low O2/alkane ratios. Turnover rates and the O2/alkane ratios required for transitions among kinetic regimes differ significantly between CH4 and C2H6 reactants, because of the different CH bond energies, strength of alkylO* interactions, and O2 consumption stoichiometries for these two molecules. Vacancies emerge at higher O2/alkane ratios for C2H6 than for CH4 reactants, because their weaker CH bonds lead to faster scavenging of O* and to lower O* coverages, which are set by the kinetic coupling between CH and OO activation steps. The elementary steps, kinetic regimes, and mechanistic analogies reported here for C2H6 and CH4 reactions with O2 are consistent with all rate and isotopic data, with their differences in CH bond energies and in alkyl binding, and with the catalytic consequences of surface coordination and cluster size. The rigorous mechanistic interpretation of these seemingly complex kinetic data and cluster size effects provides useful kinetic guidance for larger alkanes and other catalytic surfaces based on the thermodynamic properties of these molecules and on the effects of metal identity and surface coordination on oxygen binding and reactivity.  相似文献   

10.
The products of pyrolysis at 525 and 840 °C of two asphaltites from South-Eastern Turkey have been analysed and compared with the bitumen obtained by solvent extraction. The yield of oil product is reasonably similar for all three treatments, with gas (hydrogen, ethene, C1C4 alkanes and hydrogen sulphide) being liberated during pyrolysis. Greater percentages of alkanes with shorter chain lengths (along with some alkenes), and of pentane-soluble aromatic oils with reduced molecular masses, are generated during pyrolysis, at the expense of asphaltenes. The extra alkanes are generated partly by the cracking of aromatic side-chains and also from kerogen. Pyrolysis reduces the number of sulphur linkages in the oil, but nitrogen- and oxygen-containing structures are liberated from kerogen during heating.  相似文献   

11.
The activation of O2 over SmOF was studied by in situ laser Raman spectrometry and temperature programmed desorption (TPD). When the hydrogen- and helium-treated (1 h for each gas at 973 K) SmOF sample was cooled to 303 K in oxygen, Raman bands which correspond to the existence of O 2 2– , O 2 n– (2 >n > 1), O 2 and O 2 - (1 > > 0) species were observed. From 303 to 973 K, there was no O2 desorption but the Raman bands observed at 303 K reduced in intensity and vanished completely at 973 K, even though the sample was under an atmosphere of oxygen. We suggest that as the sample temperature increased, dioxygen species were converted to mono-oxygen species such as O which were undetectable by Raman spectrometry. O2 desorption occurred above 973 K, giving a TPD-peak at 1095 K. When C2 H6 was pulsed over the sample pretreated with oxygen and helium at 973 K, C2H4 selectivity was 91.8%. We conclude that the mono-oxygen species is responsible for the oxidative dehydrogenation of ethane to ethene.  相似文献   

12.
The kinetic of ethene polymerization catalyzed by (nBuCp)2ZrCl2/AliBu3/[Me2PhNH]+[B(C6F5)4]? has been investigated at 100 and 140 °C at pressures from 2 to 7 MPa. The initial polymerization rate Rp0 increases linearly with increasing catalyst concentration, whereas a second‐order dependence of Rp0 on the ethene concentration is found the number‐average molecular weight M n increases with increasing ethene pressure (ethene concentration). The relation between M n and ethene concentration can be explained by an equation based on a kinetic model involving a single center, two‐state catalyst system.

Influence of the Zr concentration on the initial polymerization rate Rp0. pEthene = 7 MPa, solvent (toluene/ethene) = 275 mL.  相似文献   


13.
Summary Gas chromatography has been employed to determine partial molar free energies of sorption, C1 s, as well as partial molar free energies of mixing, C1 , of atactic poly(styrene) with linear and branched alkanes (C6-C9), alkenes (C8), cyclohexane and alkylbenzenes (Ph-H to PhC6H13) within the temperature range from 403 to 463 K. The influence of nature and constitution of the solute molecule on sorption and mixing properties in poly(styrene) are discussed in terms of the competing group contributions of the components. Knowledge of this influence may be transduced to understand polymer-polymer compatibility. The cohesive energy density concept has been used to calculate the infinite dilution solubility parameter for the polymer, with the aid of G 1 . From the high temperature range the HILDEBRAND parameter 2 was extrapolated to 298 K. The value obtained, 9.14, indicates that gas chromatography is an promising alternative to the conventional methods for determination of the solubility parameter for polymers.Herrn Prof. Dr. R. C. Schulz zu seinem 65. Geburtstag herzlichst gewidmet  相似文献   

14.
With a small amount of H2 (3 6%) addition into methane feed, coke formation on 6 wt% Mo catalyst during the methane dehydroaromatization reaction was effectively suppressed and the catalyst stability was increased evidently under the reaction conditions of 1023K, 0.3MPa and 2520 mL g-MFI-1 h-1 of methane space velocity.  相似文献   

15.
The polyurethanes are synthesized from the biphenyl-4,4′-diol (mesogenic biphenol) and 1,3-Bis(isocyanatomethyl) cyclohexane, using (CH2) of 2, 6 and 11 units as flexible alkylene spacer, respectively. FTIR detects the hydrogen bond in the thermotropic liquid crystalline polyurethane. FTIR spectra show a new CO absorption with lower wavenumber at around 1658 cm?1 is assigned to “bifurcated” hydrogen bonded CO group—a CO with higher strength hydrogen bonds. The distributions of “bifurcated” hydrogen bonded CO are increased substantially along with increasing the flexible spacer length in polymer backbone. The “bifurcated” hydrogen bond existed not only at the temperature below Tg, but also existed at the temperature far higher than Tm and Ti. It almost is independent of temperature and exhibits a stable interaction (or strucuture) throughout a wide temperature range, differences from the normal liquid crystalline polyurethanes. It is worthy of predicting the thermotropic liquid crystalline polyurethane with “bifurcated” hydrogen bond would enhance its performances.  相似文献   

16.
Investigation on the Kinetics of Diffusion of Ethene on 4A Zeolites The kinetics of diffusion of ethene on 4A zeolites was examined. Conclusions on the transport mechanism were obtained by the loading dependence on the diffusion coefficients. Contrary to propene and the n-butene isomers, ethene can cross the windows to the cavities blocked up by Na+ ions in the NaA and MeNa0,8A zeolites (Me = Mg, Ca, Zn, Ba) by a translatory jump motion. In NaKA zeolites – by reason of additional blocking up of windows with K+ ions – the ethene molecules overcome the barrier only through the ethene/Na+(1) addition complexes.  相似文献   

17.
A detailed procedure for the quantitative analysis of aromatic and aliphatic hydrogen based on infrared spectroscopy was set up and implemented on some carbon-based materials produced from organic precursors (naphthalene pitch) and/or relevant in combustion field (asphaltenes, carbon particulate matter, carbon black), spanning in the H/C atomic ratio range from 0.1 to 1. The quantitative FT-IR analysis involved the spectral deconvolution in the CH vibrations regions and the calibration factors of diverse standard species having spectral characteristics suitable for the detailed peak-to-peak analysis of the CH stretching (3100–2800 cm−1) and aromatic CH bending (900–700 cm−1) regions. The good agreement between the H/C atomic ratio obtained by quantitative FT-IR analysis and elemental analysis showed a reasonable reliability of the procedure. The major value of the developed FT-IR quantitative technique relies also on the capacity of discriminating between the different kinds of aliphatic and aromatic hydrogen. The quantitative and detailed analysis of hydrogen in form of CH3, CH2 and CH groups and in form of solo, duo and trio/quatro aromatic hydrogens showed to be useful also for inferring the structure of the aromatic moieties constituting the CC backbone of carbon materials.  相似文献   

18.
Compensation between adsorption entropies and enthalpies results in less than a two-fold variation in adsorption equilibrium constants for C3–C6 alkanes at temperatures relevant for monomolecular cracking; the size-independent activation energy for CC bond activation in C3–C6 alkanes indicates that the marked increase in monomolecular cracking turnover rates observed with alkane chain size reflects a concurrent increase in activation entropies. Thermodynamic treatments for non-ideal systems rigorously describe confinement effects within zeolite channels and show that pre-exponential factors depend on solvation effects of the zeolite-host environment through variations in the thermodynamic activity of the zeolitic proton. Observed differences in rates and selectivities of monomolecular alkane activation with zeolite structure, after normalization to intrazeolitic concentrations, reflect differences in intrinsic rate constants.  相似文献   

19.
A comparison of slurry versus fixed-bed reactor design principles for methanol and Fischer-Tropsch distillate production.Notation a gas-liquid interfacial area, m–1 - CCat catalyst concentration, kg mole/m3 - C HG hydrogen concentration in gas phase, kg mole/m3 - C * HL hydrogen concentration, liquid, in equilibrium with gas, kg mole/m3 - C HL hydrogen concentration in the liquid phase, kg mole/m3 - D I.D. of reactor, m - GHSV Gas hourly space velocity, Nm3 (H2 + CO)/[h · m3 reactor volume], (reactor volume is expanded slurry height times cross section area) - H solubility coefficient of hydrogen =C HG/C * C HL - l Inlet ratio of CO/H2 - k L liquid side mass transfer coefficient, m/s - k H effective reaction rate constant for hydrogen consumption, s–1 (note that to agree with space velocity in Nm3/[s · kgCat],k H =k H ·CCat wherek H is in m3/[kg · s] - L Length of expanded slurry bed, m - P pressure, kPa - r rate of hydrogen consumption,r =k H ·C HL, kg moles/[m3 · s] - SV Space velocity in actual m3 inlet gas/[s · m3] - T temperature, K - U Usage ratio of CO/H2 - X H hydrogen fractional conversion per pass (IfU=l,X H =X CO) - contraction factor, = [m3/s(X H 2+CO = 1)-m3/s(inlet)]/[m3/s(inlet)] - * contraction factor modified for H2 conversion, * = · (1 +U)/(1 +l) - L fractional liquid hold-up  相似文献   

20.
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