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 binary adsorption characteristics of methane and ethane on dry coal to 40 atm pressure have been calculated from pure-component isotherms. In some coal seams, pressures exceeding 40 atm have been recorded and the methane sampled from the virgin coal often shows a few percent of ethane. The binary adsorption characteristics were calculated by employing the ideal adsorbed solution theory of Myers and Prausnitz, and experimentally-determined (Type I) pure gas isotherms at 0, 30 and 50 °C. The coal used in this investigation was high-volatile ‘A’ bituminous (hvab) from the Pennsylvania Pittsburgh seam. Gas nonideality was accounted for by replacing pressure with fugacity. Adsorption of methane on dry coal is purely physical; the isosteric heat of adsorption does not exceed 2.4 kcal/mol* at 30 °C on the above coal. Isobars on the resulting binary equilibrium diagram exhibited an unexpected phenomenon of intersecting each other which might be attributable to the above nonideality considerations. The region of a few percent of ethane, which is of practical importance from the viewpoint of coal seams, was expanded and reduced to an equation: V(CH4) = −21.52 + 7.18(VF) + 16.88(VF)2 −0.395(P) − 0.00661(P)2 + 0.824(T) − 0.00030(T)2 + 0.928(VF)(P) − 0.858(VF)(T). V(Total) = 25.9 − 23.6(VF) + 0.655(P) − 0.00875(P)2 − 0.795(T) + 0.743(VF)(T) where V(CH4) and V(Total) = cm3(STP)CH4 and total gas respectively adsorbed per g dry coal; VF = vol. fraction of methane as analysed at 1 atm (0.94 VF 1.0); P = seam pressure, atm (0 P 40); T=seam temperature, °C(−10 T 50).  相似文献   

2.
Mesoporous γ-Ga2O3 was prepared by calcination (at 773 K) of a gallia gel obtained by adding ammonia to an ethanolic solution of gallium nitrate. The corresponding powder X-ray diffraction pattern was found to be similar to that of γ-alumina; all diffraction lines could be indexed by assuming a cubic spinel-type structure having a lattice parameter a0=0.830 nm. Nitrogen adsorption–desorption at 77 K showed the material to have a BET surface area of 120 m2 g−1 and a most frequent pore radius of 2.1 nm. The surface chemistry was studied by FTIR spectroscopy of adsorbed carbon monoxide at liquid nitrogen temperature. Partially hydroxylated γ-Ga2O3 gave main O–H stretching bands at 3692 and 3637 cm−1. These hydroxyl bands were significantly perturbed by adsorbed CO, thus showing a Brønsted acid character. Lewis acidity was monitored by analyzing the C–O stretching mode of carbon monoxide adsorbed on coordinatively unsaturated Ga3+ ions; main IR absorption bands of Ga3+CO adducts were found at 2220 and 2193 cm−1.  相似文献   

3.
The dehydration and rehydration processes of mesolite belonging to NAT group of zeolites were investigated using in situ Fourier Transform Infrared spectroscopy (FTIR). The thermal induced variations of the water molecule bending (ν2), the stretching (ν3 and ν1) modes and the corresponding second order modes in the wavenumber region 4000–8000 cm−1 were followed as indicative of the dehydration process. Observed spectral variations were well correlated with thermogravimetric studies and indicate that the mesolite dehydrates in three stages (470; 510 and 650 K). Anomalous spectral variations of 4600 cm−1 at first stage indicate that the dehydration is triggered by the expulsion of water coordinating AlO4 tetrahedron. Partial rehydration (up to 85%) at second stage indicates disordering of natrolite and scolecite layers. During the third stage mesolite completely dehydrates, causing the destruction of framework.  相似文献   

4.
The reduction of nitric oxide by propene in the presence of oxygen over platinum-group metals supported on TiO2, ZnO, ZrO2, and Al2O3 has been investigated by combined diffuse reflectance FT-IR spectroscopy and catalytic activity studies under flow reaction conditions at 523–673 K and atmospheric pressure. The catalytic activity for the selective reduction of nitric oxide and the intensity of the IR bands due to reaction species depended strongly on the nature of the support, type of supported metal, reaction time and temperature. The main surface species detectable by IR were adsorbed hydrocarbons (2900–3080 cm−1), isocyanate (2180, and 2232–2254 cm−1), cyanide (2125 cm−1), nitrosonium (1901 cm−1), CO2 (2343–2357 cm−1), CO (2058 cm−1) and carbonate (1300–1650 cm−1) species. In the case of rhodium containing catalysts, when supported on Al2O3, they exhibited both the highest concentration of surface species and the highest activity for nitric oxide reduction and selectivity to nitrogen. The catalytic activity and the IR intensities of the nitrosonium and isocyanate bands increased with reaction temperature, reached their maximum between 570 and 620 K, and then decreased at higher temperatures. The IR band intensities due to nitrogen containing surface species were found to be strongly correlated to the activity for nitric oxide conversion and only slightly related to the selectivity to dinitrogen.  相似文献   

5.
Pure Fe and Fe/Cr alloys (3% and 12% Cr) were oxidized at 250–260°C and subsequently electrochemically reduced in borate buffer, pH 8.4. The reduction was followed by in situ Raman spectroscopy as well as chronopotentiometric and chronoamperometric measurements. At several time intervals ex situ Raman control investigations and ESCA analyses were undertaken. Although at −0.7 V vs sce partial reduction of the oxide film was observed, -Fe2O3 remained unchanged and reduced at −0.9 V. A mixture of spinel structured oxide, possibly Fe3O4, and a ferrous deposit were observed when the oxide film was galvanostatically reduced at 100 μA cm−2; after prolonged reduction the bare metal resulted. The thermally created oxide of the Cr alloy possessed an inner Cr rich layer which during the galvanostatic reduction of the oxide film at 50 μA cm−2 is most likely responsible for an increase in the overpotential of more than 80 mV.  相似文献   

6.
Nanometer perovskite-type oxides La1−xSrxMO3−δ (M = Co, Mn; x = 0, 0.4) have been prepared using the citric acid complexing-hydrothermal-coupled method and characterized by means of techniques, such as X-ray diffraction (XRD), BET, high-resolution scanning electron microscopy (HRSEM), X-ray photoelectron spectroscopy (XPS), temperature-programmed desorption (TPD), and temperature-programmed reduction (TPR). The catalytic performance of these nanoperovskites in the combustion of ethylacetate (EA) has also been evaluated. The XRD results indicate that all the samples possessed single-phase rhombohedral crystal structures. The surface areas of these nanomaterials ranged from 20 to 33 m2 g−1, the achievement of such high surface areas are due to the uniform morphology with the typical particle size of 40–80 nm (as can be clearly seen in their HRSEM images) that were derived with the citric acid complexing-hydrothermally coupled strategy. The XPS results demonstrate the presence of Mn4+ and Mn3+ in La1−xSrxMnO3−δ and Co3+ and Co2+ in La1−xSrxCoO3−δ, Sr substitution induced the rises in Mn4+ and Co3+ concentrations; adsorbed oxygen species (O, O2, or O22−) were detected on the catalyst surfaces. The O2-TPD profiles indicate that Sr doping increased desorption of the adsorbed oxygen and lattice oxygen species at low temperatures. The H2-TPR results reveal that the nanoperovskite catalysts could be reduced at much lower temperatures (<240 °C) after Sr doping. It is observed that under the conditions of EA concentration = 1000 ppm, EA/oxygen molar ratio = 1/400, and space velocity = 20,000 h−1, the catalytic activity (as reflected by the temperature (T100%) for EA complete conversion) increased in the order of LaCoO2.91 (T100% = 230 °C) ≈ LaMnO3.12 (T100% = 235 °C) < La0.6Sr0.4MnO3.02 (T100% = 190 °C) < La0.6Sr0.4CoO2.78 (T100% = 175 °C); furthermore, there were no formation of partially oxidized by-products over these catalysts. Based on the above results, we conclude that the excellent catalytic performance is associated with the high surface areas, good redox properties (derived from higher Mn4+/Mn3+ and Co3+/Co2+ ratios), and rich lattice defects of the nanostructured La1−xSrxMO3−δ materials.  相似文献   

7.
Pdn+/Cen+/Na+/γ-Al2O3-type materials used as FCC additives for CO/NOx control were characterized by extended X-ray absorption fine structure (EXAFS) spectroscopy and in situ FTIR. The EXAFS data indicate that in freshly prepared samples palladium is present in the form of highly dispersed PdO species. Reduction with H2 at 500 °C leads to the formation of small Pd clusters incorporating on average approximately six to eight metal atoms at a Pd−Pd bond distance of 2.76 Å. All components of these materials can interact with NO and promote the formation of nitrate/nitrite species, essentially “trapping” NOx species on the catalyst surface. However, the Na+ species dominate the surface chemistry and readily form sodium nitrates with a characteristic IR band at 1370–1385 cm−1. Finally, hydroxyls from the support are also actively participating in the formation of HNOx type compounds with characteristic stretching vibrations in the 3500–3572 cm−1 region.  相似文献   

8.
Medium energy (5–25 keV) 13C+ ion implantation into diamond (100) to a fluence ranging from 1016 cm−2 to 1018 cm−2 was performed for the study of diamond growth via the approach of ion beam implantation. The samples were characterized with Rutherford backscattering/channelling spectroscopy, Raman spectroscopy, X-ray photoemission spectroscopy and Auger electron spectroscopy. Extended defects are formed in the cascade collision volume during bombardment at high temperatures. Carbon incorporation indeed induces a volume growth but the diamond (100) samples receiving a fluence of 4 × 1017 to 2 × 1018 at. cm−2 (with a dose rate of 5 × 1015 at. cm−2 s−1 at 5 to 25 keV and 800 °C) showed no He-ion channelling. Common to these samples is that the top surface layer of a few nanometers has a substantial amount of graphite which can be removed by chemical etching. The rest of the grown layer is polycrystalline diamond with a very high density of extended defects.  相似文献   

9.
It was found that TeCl4 in methylene chloride (containing tetrabutylammonium perchlorate as a supporting electrolyte) is reduced to TeCl2, Te0 and Te−2 in potential applied. Two reduction waves are observed during the reduction of TeCl4 at a platinum rde (Ep1 = 0.08 ± 0.02 V, Ep2 = −1.10 ± 0.02 V) due to the reduction of TeCl4 to Te and Te−2 respectively. A cathodic deposition of tellurium from TeCl4 is followed by anodic stripping wave (Ep1 = 0.42 ± 0.02 V), corresponding to the oxidation of Te to TeCl2. It has been shown that reduction of TeCl4, or TeCl−26 to Te causes coating of the electrode with metallic tellurium, on whose surface chloride ions are strongly bonded. It was found that when the electrolysis solution contains excess of chloride ion with respect to TeIV the peak potentials of the cathodic waves shift to more cathodic values (Ep1 = 0.30 ± 0.02 V, Ep2 = −1.25 ± 0.02 V). After cathodic deposition of at least a monolayer of tellurium from such a solution two anodic waves appear (Ep1 = 0.12 ± 0.03 V, Ep2 = 0.24 ± 0.03 V) which are both due to the oxidation of Te to TeII.  相似文献   

10.
The three-dimensional orientation change during thermally induced structural change was monitored in a ‘real-time’ mode with polarized FTIR–ATR spectroscopy using a double-edged parallelogram crystal in a temperature-controlled ATR set-up. Upon cold-crystallization of stretched PTT sample, the growth of crystalline phase along the stretching direction was much faster than those along the perpendicular directions. The cold-crystallization of melt-quenched amorphous PTT was found to produce crystallites of no orientation, as expected. The formation of crystalline phase along three orthogonal directions was followed by changes of three attenuation indices of 1358 cm−1 band which is associated with the wagging vibration motion of CH2 groups in crystalline phase. The orientation of CH2 groups in amorphous phase estimated from three attenuation indices of 1385 cm−1 band along three directions was very small even for the cold-crystallized, anisotropic PTT sample. This work appears to be the first successful observation in a ‘real-time’ mode on the dynamic change of three-dimensional orientation of polymeric materials using temperature-controlled polarized FTIR–ATR spectroscopy.  相似文献   

11.
Three procedures were employed for the preparation of Fe-zeolites with ZSM-5 (MFI), ferrierite (FER) and beta (BEA) structures: ion exchange from FeCl3 solution in acetyl acetonate and solid-state ion exchange from FeCl2 using an oxygen or nitrogen stream. A combination of UV–VIS–NIR spectra, IR spectra of skeletal vibrations and of adsorbed NO, as well as voltammetry provided information on the type of Fe species introduced. Single Fe(III) ion complexes (Fe(H2O)6−xOHx) in hydrated zeolites were reflected in the charge-transfer bands at 33 100, 37 300 and 45 600 cm−1. The single Fe(II) ions at cationic sites in evacuated zeolites yielded (through perturbation of framework T–O bonds) characteristic bands (910–950 cm−1) in the region of the skeletal window. These Fe(II) ions with adsorbed NO were also reflected in vibrations at 1880 cm−1. Dinuclear Fe–oxo complexes yielded the Vis band at 28 200 cm−1. Voltammetry indicated the presence of Fe oxides (hematite) through the reduction peak at −0.7 V. Such oxide-like species were also reflected in the absorption edge at 19 800 cm−1, and a doublet at 11 000 and 11 800 cm−1 in the Vis spectra. Fe(II)–NO vibrations at 1840, 1810 and 1760 cm−1 belonged to the undefined exposed Fe cations, probably originating from supported oxides. Using an ion exchange procedure, employing FeCl3 in acetyl acetonate, exclusively Fe ions at cationic sites could be introduced at low concentrations (Fe/Al < 0.1). At higher Fe loadings, dinuclear Fe–oxo complexes were formed preferably in Fe-ZSM-5, but were absent in Fe-beta. Exclusively single Fe species could not be prepared at Fe concentrations above Fe/Al > 0.2; all three types of Fe species, single Fe ions, dinuclear Fe–oxo complexes and Fe oxides were formed.  相似文献   

12.
In this paper, single-crystalline SnO2 nanowires have been successfully prepared by a carbothermal reduction process employing SnO as the starting material and CuO as the catalyst. Their morphologies, purity and sizes of the products were characterized by transmission electron microscopy (TEM), selected area electron diffraction, X-ray diffraction, field emission scanning electron microscopy (FESEM) and Raman spectroscopy, respectively. The FESEM images reveal wire-like and rod-shaped nanowires of about 100–800 μm in length and 30–200 nm in the transverse dimensions. The three observed Raman peaks at 474, 634 and 774 cm−1 indicate the typical rutile phase of the SnO2 which is in agreement with the X-ray diffraction results. The influence of some reaction parameters, including the temperature and the reaction duration, on the forming, morphology and particle size of SnO2 crystallize is discussed.  相似文献   

13.
Small-angle neutron scattering experiments in the range of q2 from 0.01 to 25 nm−2 have been carried out on branched epoxide resins based on bisphenol-A at the Institute Laue—Langevin (I.L.L) in Grenoble (q=(4π/λ) sin(θ/2)). Measurements were made with six samples in the range of MW from 1500 to 19 000 and four concentrations between 1.3 and 10% (w/w) in deuterated diglyme. The results are as follows: (i) The mean square radius of gyration follows a relationship S2z=4.69×10−4M1.20W (nm2). (ii) In all cases fairly large second virial coefficients A2 are obtained which, however, decrease strongly with molecular weight. Above MW=2500, the virial coefficient follows the relationship A2=1.6M−0.85W (mol cm3g−2). (ii) The reciprocal particle scattering factor as a function of q2 exhibits only a slight upturn and otherwise shows the behaviour of a randomly branched polycondensate. The slight upturn is discussed as being caused by the finite volume of the monomeric unit. Possible reasons for the high exponent in the S2z versus MW dependence are briefly discussed.  相似文献   

14.
The reaction mechanism of the reduction of NO by propene over Pd-based catalysts was studied by FTIR spectroscopy. It was observed that the reaction between NO and propene most probably goes via isocyanate (2256–2230 cm−1), nitrate (1310–1250 cm−1) and acetate (1560 and 1460 cm−1) intermediates formation. Other possible intermediates such as partially oxidized hydrocarbons, NO2, and formates were also detected. The reaction between nitrates and acetates or carbonates reduced nitrates to N2 and oxidized carbon compounds to CO2. In situ DRIFT provides quick and rather easily elucidated data from adsorbed compounds and reaction intermediates on the catalyst surface. The activity experiments were carried out to find out the possible reaction mechanism and furthermore the kinetic equation for NO reduction by propene.  相似文献   

15.
Rate data have been obtained for CO hydrogenation on a well-characterized 11.7% Co/TiO2 catalyst in a differential fixed bed reactor at 20 atm, 180–240°C, and 5% conversion over a range of reactant partial pressures. The resulting kinetic parameters can be used to model precisely and accurately the kinetics of this reaction within this range of conditions. Turnover frequencies and rate constants determined from this study are in very good to excellent agreement with those obtained in previous studies of other cobalt catalysts, when the data are normalized to the same conditions of temperature and partial pressures of the reactants. Based on this comparison CO conversion and the partial pressure of product water apparently have little effect on specific rate per catalytic site. The data of this study are fitted fairly well by a simple power law expression of the form −rCO=kPH20.74PCO−0.24, where k=5.1×10−3 s−1 at 200°C, P=10 atm, and H2/CO=2/1; however, they are best fitted by a simple Langmuir–Hinshelwood (LH) rate form −rCO=aPH20.74PCO/(1+bPCO)2 similar to that proposed by Yates and Satterfield.  相似文献   

16.
Secondary electron emission from boron doped diamond polycrystalline membranes (hole concentration 5×1018 cm−3), prepared by microwave plasma assisted CVD, was investigated in both the reflection and transmission configurations. The model of secondary electrons behavior taking into account the distribution and diffusion mechanism of secondary electrons is proposed to explain the yield dependencies on primary electron energy in both configurations. The model predicts the SEE yield K=19 at the primary electron energy E0 close to 1 keV for reflection configuration and K=3–7 at E0=15–30 keV for transmission configuration for polycrystalline films used in the study. Experimental measurements of the SEE yield vs. primary electron energy (18 at E0=950 eV for the reflection scheme and 3.5–4 at E0=25 keV for the transmission one) are found to accord well with the theoretical results. Estimations, which were made using the model, show that SEE yield in transmission configuration can be increased up to 60 for the primary electron energy of about 10 keV. Since such high yields in transmission scheme may be obtained in monocrystalline membrane, another approach using porous polycrystalline diamond membranes is considered. Porous diamond membranes having SEE yield in transmission scheme of more than 10 at the primary electron energy E0=1 keV were fabricated.  相似文献   

17.
We report that ultrastable faujasite-based ruthenium zeolites are highly active catalysts for N2O decomposition at low temperature (120–200°C). The faujasite-based ruthenium catalysts showed activity for the decomposition of N2O per Ru3+ cation equivalent to the ZSM-5 based ruthenium catalysts at much lower temperatures (TOF at 0.05 vol.-% N2O: 5.132 × 10−4 s−1 Ru−1 of Ru-HNaUSY at 200°C versus 5.609 × 10−4 s−1 Ru−1 of Ru-NaZSM-5 at 300°C). The kinetics of decomposition of N2O over a Ru-NaZSM-5 (Ru: 0.99 wt.-%), a Ru-HNaUSY (Ru: 1.45 wt.-%) and a Ru-free, Na-ZSM-5 catalyst were studied over the temperature range from 40 to 700°C using a temperature-programmed micro-reactor system. With partial pressures of N2O and O2 up to 0.5 vol.-% and 5 vol.-%, respectively, the decomposition rate data are represented by: −dN2O/dt=itk(PN2O) (PO2)−0.5 for Ru-HNaUSY, −dN2O/dt=k(PN2O) (PO2)−0.1 for Ru-NaZSM-5, and −dN2O/dt=k(PN2O)−0.2 (PO2)−0.1 for Na-ZSM-5. Oxygen had a stronger inhibition effect on the Ru-HNaUSY catalyst than on Ru-NaZSM-5. The oxygen inhibition effect was more pronounced at low temperature than at high temperature. We propose that the negative effect of oxygen on the rate of N2O decomposition over Ru-HNaUSY is stronger than Ru-NaZSM-5 because at the lower temperatures (<200°C) the desorption of oxygen is a rate-limiting step over the faujasite-based catalyst. The apparent activation energy for N2O decomposition in the absence of oxygen is much lower on Ru-HNaUSY (Ea: 46 kJ mol−1) than on Ru-NaZSM-5 (Ea: 220 kJ mol−1).  相似文献   

18.
The quadratic electrooptic coefficients g1111 and g2211 have been measured interferometrically in ADP at room temperature. In addition, ‖g1111-g2211‖ was determined by a dynamic polarimetric technique at the same temperature. Both experiments yield values for g1111-g2211 at 21 C and λ = 0.633 μm with an average of (-5.9±0.5)×10-20 m2/V2. An estimate is made, using this value, of the spontaneous antipolarization appearing in the antiferroelectric phase of ADP. This antipolarization is found to be comparable in magnitude with the spontaneous polarizations observed in ferroelectrics belonging to the KDP family of crystals.  相似文献   

19.
In situ investigations of the interaction of methanol on zeolites of the faujasite, pentasil and mordenite types were performed. We suppose that CH3)- or (CH3)2O… species exist on the surface of the H-ZSM-5 zeolite in the temperature range between 380 and 470 K which desorb above 470 K and form dimethyl ether (DME). The band at 1502 cm−1 represents aromatics being intermediates in the aromatization of methanol. The band at 1595 cm−1 is due to polymeric “coke” which cannot be removed from the zeolite up to 670 K. In the case of the H-Y zeolite no band at 1500 cm−1 assigned to aromatics could be observed.  相似文献   

20.
Polycrystalline (PbS)1.14(TaS2)2, a misfit layer sulfide, was used as cathodic material for lithium secondary battery. One molar LiClO4 in propylene carbonate (PC) was used as electrolyte. The cell could be galvanostatic discharged down to x = 4.6 [Lix(PbS)1.14(TaS2)2] when the current density was 65 μA cm−2 and the cell was cycled more than 100 times between 3.5 and 1.5 V at a current density of 260 μA cm−2. Lattice expansion increased linearly with lithium content and was less than that reported for the Li/TaS2 system. Chemical diffusion coefficients were determined by a modified version of the galvanostatic intermittent titration technique and they were fairly constant in the composition range 0.2 < x < 1, and an average value of 8.1 × 10−11 cm2 s−1 was calculated. Sodium intercalation was also accomplished, but the uptake of this ion resulting in a significant lattice expansion compared with that observed for lithium ions. Moreover, a similar dependence of the sodium chemical diffusion coefficient on the composition was observed with an average value of 1.4 × 10−10 cm2 s−1, somewhat higher than that of lithium ion. We believe that differences in lattice expansion may be responsible for the differences found in the chemical diffusivity values.  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