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.
USY (ultrastabilized Y) and Ce/USY (5 wt.% supported) zeolite acidities were characterized by microcalorimetric and adsorption studies of pyridine using liquid phase (Cal-Ad), thermogravimetry, and infrared analysis. The average adsorption enthalpies determined by microcalorimetry were −125.0 kJ mol−1 for USY and −97.2 kJ mol−1 for Ce/USY. A heterogeneous distribution of acid sites with heats of adsorption ranging from −134.0 (maximum heat value for USY) to −73.5 (minimum heat value for Ce/USY) kJ mol−1 was found for both zeolites. A two-site model was best fitted by the Cal-Ad method for HUSY (n1 = 0.1385 mmol g−1 with ΔH1 = −134.0 kJ mol−1, and n2 = 0.7365 mmol g−1 with ΔH2 = −101.5 kJ mol−1) and Ce/HUSY (n1 = 0.0615 mmol g−1 with ΔH1 = −117.6 kJ mol−1, and n2 = 0.7908 mmol g−1 with ΔH2 = −83.6 kJ mol−1). DRIFTS measurements after pyridine adsorption showed that USY zeolite possesses only Brønsted acidity and that cerium impregnation leads to the appearance of Lewis sites. Based on these results, three families of acid strength were distinguished: (i) strong Brønsted sites (ΔH > −130 kJ mol−1); (ii) Brønsted sites with intermediate strength (−100 < ΔH < −130); and (iii) weak Brønsted and Lewis sites (ΔH < −100). Thermogravimetric analysis showed that the strongest sites were able to retain pyridine up to 800 °C and that cerium incorporation leads to a more stable zeolite. A loss of strength was observed after impregnation. The total number of sites desorbed after gas adsorption (0.88 and 0.95 mmol for HUSY and Ce/HUSY, respectively) supports the Cal-Ad results (0.88 and 1.19 mmol for HUSY and Ce/HUSY, respectively) and indicates that not all Al sites are available to pyridine. The methodology used in this work for solid acid characterization (Cal-Ad) proved to be efficient in the evaluation of acid strength, total number and distribution of acid sites. XRPD, ICP-AES, 27Al NMR, and FTIR were used for additional structural characterization.  相似文献   

2.
The gas phase catalytic hydrodechlorination (HDC) of mono- and di-chlorobenzenes (423 K ≤ T ≤ 593 K) over unsupported and silica supported Mo carbide (Mo2C) is presented as a viable means of detoxifying Cl-containing gas streams for the recovery/reuse of valuable chemical feedstock. The action of Mo2C/SiO2 is compared with MoO3/SiO2 and Ni/SiO2 (an established HDC catalyst). The pre- and post-HDC catalyst samples have been characterized in terms of BET area, TG-MS, TPR, TEM, SEM, H2 chemisorption/TPD and XRD analysis. Molybdenum carbide was prepared via a two step temperature programmed synthesis where MoO3 was first subjected to a nitridation in NH3 followed by carbidization in a CH4/H2 mixture to yield a face-centred cubic (-Mo2C) structure characterized by a platelet morphology. Pseudo-first order kinetic analysis was used to obtain chlorobenzene HDC rate constants and the associated temperature dependences yielded apparent activation energies that decreased in the order MoO3/SiO2 (80 ± 5 kJ mol−1) ≈ MoO3 (78 ± 8 kJ mol−1) > Ni/SiO2 (62 ± 3 kJ mol−1) ≈ -Mo2C (56 ± 6 kJ mol−1) ≈ -Mo2C/SiO2 (53 ± 3 kJ mol−1). HDC activity was lower for the dechlorination of the dichlorobenzene reactants where steric hindrance influenced chloro-isomer reactivity. Supporting -Mo2C on silica served to elevate HDC performance, but under identical reaction conditions, Ni/SiO2 consistently delivered a higher initial HDC activity. Nevertheless, the decline in HDC performance with time-on-stream for Ni/SiO2 was such that activity converged with that of -Mo2C/SiO2 after three reaction cycles. A temporal loss of HDC activity (less extreme for the carbides) was observed for each catalyst that was studied and is linked to a disruption to supply of surface active hydrogen as a result of prolonged Cl/catalyst interaction.  相似文献   

3.
Relative acid strength and acid amount of solid acids (alumina, silica-alumina, sulfated zirconia, mordenite, ZSM-5, beta, Y, and reduced MoO3) are determined by argon adsorption technique. To obtain the heat of Ar adsorption and saturated adsorption amount, the adsorption isotherm is analyzed using the theory reported by Cremer and Flügge. The obtained heats of Ar adsorption and saturated adsorption amounts of sulfated zirconia catalysts and proton-type zeolites correspond well with the activities of acid-catalyzed pentane isomerization of these catalysts. The heats of adsorption were −22 kJ mol−1 for sulfated zirconia, and ca. −19 kJ mol−1 for mordenite, ZSM-5, and beta. Molybdenum oxides reduced at 623 and 773 K show large heat of adsorption (−19.3 and −19.7 kJ mol−1, respectively), and these are classified into the superacid.  相似文献   

4.
The electro-oxidation of dimethyl ether (DME) on PtMe/Cs (Me = Ru, Sn, Mo, Cr, Ni, Co, and W) and Pt/C electro-catalysts were investigated in an aqueous half-cell, and compared to the methanol oxidation. The addition of a second metal enhanced the tolerance of Pt to the poisonous species during the DME oxidation reaction (DOR). The PtRu/C electro-catalyst showed the best electro-catalytic activity and the highest tolerance to the poisonous species in the low over-potential range (<0.55 V, 50 °C) among the binary electro-catalysts and the Pt/C, but at the higher potential (>ca. 0.55 V, 50 °C), the Pt/C behaved better than PtRu/C. The apparent activation energy for the DOR decreased in the order: PtRu/C (57 kJ mol−1) > Pt3Sn/C (48 kJ mol−1) ≈ Pt/C (46 kJ mol−1). On the other hand, the activation energy for the MOR showed a different turn, decreased in the following order: Pt/C (43 kJ mol−1) > Pt3Sn/C (35 kJ mol−1) ≈ PtRu/C (34 kJ mol−1). The temperature dependence of the DOR was greater than that of the oxidation of methanol (MOR) on the PtRu/C.  相似文献   

5.
Well-known, yet undefined, changes in the conditions and activity of palladized zerovalent iron (Fe/Pd) over an extended period of time hindered a careful study of dechlorination kinetics in long-term experiments. A short-term experimental method was, therefore, developed to study the effects of temperature and solvent on the dechlorination of monochlorobiphenyls (MCBs), 2-chlorobiphenyl (2-ClBP), in particular by Fe/Pd. The experiments started with specified initial conditions and lasted only for 10 min. The average value (k) of the first-order rate constant for the dechlorination of 2-ClBP was 0.13 ± 0.03 L m−2 h−1, not significantly different from the average values for 3-chlorobiphenyl and 4-chlorobiphenyl. The apparent activation energy was 20 ± 4 kJ mol−1 and 17 ± 7 kJ mol−1, in a temperature range between 4 °C and 60 °C, for the dechlorination of 2-ClBP using two batches of Fe/Pd catalyst. The k values decreased significantly in mixtures with a methanol concentration higher than 10%. The values of the rate constant were slightly influenced by the initial concentrations in the experiments at a low temperature and in a solution with a high methanol concentration. The concentration dependence was described with a Langmuir equation, based on the Langmuir–Hinshelwood mechanism that includes an adsorption step of a single species preceding a rate-determining catalytic reaction.  相似文献   

6.
This paper investigates the reduction of ionic concentration and carbon oxygen demand (COD) in dairy process waters modelled by one volume of skim milk diluted with two volumes of water using shear-enhanced reverse osmosis. Initial COD and conductivity were, respectively, 36,000 mg O2 L−1 and 2000 μS cm−1. We have compared the performances of a VSEP vibratory pilot and of a single rotating disk-stationary membrane module equipped with the same Desal AG membrane (Osmonics). Membrane shear rates were varied by changing the vibration frequency in the VSEP and the disk rotation speed or adding radial vanes in the other module. In all tests the permeate COD was reduced below 15 mg O2 L−1. Permeate fluxes reached a maximum of 180 L h−1 m−2 at a transmembrane pressure (TMP) of 4 MPa at initial concentration with the VSEP at its resonant frequency and with the disk equipped with 6 mm high vanes rotating at 2000 rpm. Permeate conductivity fell from 60 μS cm−1 at 1 MPa to about 18 μS cm−1 at 4 MPa. In concentration tests, corresponding permeate fluxes at the maximum volume reduction ratio reached (VRR = 8), were 55 L h−1 m−2 for the VSEP and 60 L h−1 m−2 for the rotating disk at a TMP of 4 MPa. Permeate conductivities increased exponentially with VRR from 18 to 210 μS cm−1 for the rotating disk and to 250 μS cm−1 for the VSEP. However the mean conductivity of collected permeate varied from 38 μS cm−1at highest shear rate to 60 at lower shear rates. This study shows that these filtration systems permit to obtain reusable water from this high initial COD model effluent with one single reverse osmosis step.  相似文献   

7.
The oxidation of CH4 over Pt–NiO/δ-Al2O3 has been studied in a fluidised bed reactor as part of a major project on an autothermal (combined oxidation–steam reforming) system for CH4 conversion. The kinetic data were collected between 773 and 893 K and 101 kPa total pressure using CH4 and O2 compositions of 10–35% and 8–30%, respectively. Rate–temperature data were also obtained over alumina-supported monometallic catalysts, Pt and NiO. The bimetallic Pt–NiO system has a lower activation energy (80.8 kJ mol−1) than either Pt (86.45 kJ mol−1) and NiO (103.73 kJ mol−1). The superior performance of the bimetallic catalyst was attributed to chemical synergy. The reaction rate over the Pt–NiO catalyst increased monotonically with CH4 partial pressure but was inhibited by O2. At low partial pressures (<30 kPa), H2O has a detrimental effect on CH4 conversion, whilst above 30 kPa, the rate increased dramatically with water content.  相似文献   

8.
For the production of siloxane fluids, the viability of using a multi-channel monolith as a catalyst support system in a three-phase reactor has been studied. The catalyst was tripotassium phosphate (K3PO4). Experiments were performed in a single-channel flow reactor (15 mm i.d. and 500 mm catalyst coated length). The rate of reaction was followed by monitoring the disappearance of the hydroxyl group (–OH). Reaction experiments were performed at a hydroxyl group concentration range from 150 to 170 mol m−3, T=373–413 K and P=7.9 kPa with a nitrogen purge. The maximum temperature of operation was restricted to 413 K to avoid the formation of undesirable by-products. In the regime controlled by chemical kinetics, reaction was of an apparent first order with respect to –OH concentration, and in the apparent rate constant, the pre-exponential factor was 4.19×10−4 ms−1, and the apparent activation energy was 16.1 kJ mol−1. These are only valid for the operating pressure and purge gas flowrate used, as both of these are shown to affect water removal from the liquid phase and, hence, reaction rates. Mass transfer coefficients from the liquid to the catalyst surface were estimated and these increased rapidly with flowrate and were higher than expected for a falling liquid film.  相似文献   

9.
The single gas H2 and N2 permeability of a 4 μm thick dense fcc-Pd66Cu34 layer has been studied between room temperature and 510 °C and at pressure differences up to 400 kPa. Above 50 °C the H2 flux exhibits an Arrhenius-type temperature dependence with JH2=(5.2±0.3) mol m−2 s−1 exp[(−21.3 ± 0.2) kJ mol−1/(R·T)]. The hydrogen transport rate is controlled by the bulk diffusion although the pressure dependence of the H2 flux deviates slightly from Sieverts’ law. A sudden increase of the H2 flux below 50 °C is attributed to embrittlement.  相似文献   

10.
以碳化硼(B4C)、二硼化钛(TiB2)、碳化钛(TiC)为原料,采用无压烧结法在2 130℃制备了含20%(质量分数)和30%TiB2的B4C基复相陶瓷,分析所制样品的密度、硬度、弯曲强度和断裂韧性。结果表明:在2 130℃,直接加入30%的TiB2亚微米颗粒,复相陶瓷的抗弯强度和断裂韧性分别达到277.6 MPa和5.38 MPa·m1/2。陶瓷中颗粒拔出和裂纹微桥接对复相陶瓷增韧作用显著。B4C–TiB2复相陶瓷的增韧机理主要是由于TiB2与B4C热膨胀系数不匹配产生的残余应力导致的微裂纹增韧和裂纹偏转增韧。  相似文献   

11.
Separation properties of a mordenite membrane for water–methanol–hydrogen mixtures were studied in the temperature range from 423 to 523 K under pressurized conditions. The mordenite membrane was prepared on the outer surface of a porous alumina tubular support by a secondary-growth method. It was found that water was selectively permeated through the membrane. The separation factor of water/hydrogen and water/methanol were 49–156 and 73–101, respectively. Even when only hydrogen was fed at 0.5 MPa, its permeance was as low as 10−9 mol m−2 s−1 Pa−1 up to 493 K, possibly suggesting that water pre-adsorbed in the micropores of mordenite hindered the permeation of hydrogen. The hydrogen permeance dramatically increased to 6.5 × 10−7 mol m−2 s−1 Pa−1 at 503 K and reached to 1.4 × 10−6 mol m−2 s−1 Pa−1 at 523 K because of the formation of cracks in the membrane. However, the membrane was thermally stabilized in the presence of steam and/or methanol.  相似文献   

12.
We have fabricated an array of cone-shaped diamond tips for use as a field electron emitter by air-plasma etching a polycrystalline diamond film with a silicon oxide mask. The difference in etching speed between the mask and the diamond resulted in the formation of cone-shaped diamond tips. Post-treatment with hydrogen plasma was effective in cleaning the diamond tips and increasing the surface conductivity. The emission from the diamond tips was measured with a diode configuration. The threshold field was 3 V μm−1, and the emission current was 0.8 nA tip−1 when the field was raised to 10 V μm−1.  相似文献   

13.
Adsorption of dihydrogen onto the zeolites Na-ZSM-5 and K-ZSM-5 renders the fundamental H–H stretching mode infrared active. The corresponding infrared absorption bands were found at 4101 and 4112 cm−1 for H2/Na-ZSM-5 and H2/K-ZSM-5, respectively. Thermodynamic characterization of the adsorbed state was carried out by means of variable-temperature infrared spectroscopy; simultaneously measuring integrated band intensity, temperature and equilibrium pressure of the gas phase. For the H2/Na-ZSM-5 system, the standard adsorption enthalpy and entropy resulted to be Δ = −10.3 (±0.5) kJ mol−1 and Δ = −121 (±10) J mol−1 K−1. For H2/K-ZSM-5 corresponding values were −9.1 (±0.5) kJ mol−1 and −124 (±10) J mol−1 K−1, respectively.  相似文献   

14.
H2/D2 exchange (473–583 K), 1,3-butadiene hydrogenation (418–513 K) and tetrahydrothiophen hydrodesulphurisation (428–557 K) have been studied over powdered Co9S8 (surface area, 7 m2 g−1) using D2 as an isotopic tracer. Hydrogen exchange proceeded as a first order process at a modest rate (k540 = 1.0 h−1 m−2) with an apparent activation energy of 67 kJ mol−1. Butadiene hydrogenation was diagnostic as to the surface state of Co9S8; samples showed either predominant 1:2-addition or 1:4-addition of hydrogen, interpreted as indicating the presence in the surface of single sites or pair/ensemble sites, respectively. Reactions at 473 K in the presence of D2 gave butenes containing 0–6 D-atoms: exchange patterns obtained from these D-distributions showed that a proportion of butadiene molecules underwent extensive dehydrogenation during the normal progress of hydrogenation. At 633 K this dehydrogenation activity was evident as self-hydrogenation which occurred in the absence of D2. Tetrahydrothiophen was desulphurised in the presence of D2 to thiophen (void of D), butadiene (containing 0–5 D-atoms) and 1-butene (containing mostly 0 and 4 D-atoms). Increase in temperature or in deuterium pressure favoured butene formation so that it became the dominant product (88%). Tetrahydrothiophen also underwent self-hydrodesulphurisation in the absence of D2. A mechanism is proposed, consistent with this D-tracer information, that accommodates dehydrogenation, desulphurisation and hydrogenation steps in the overall process. The activity of powdered Co9S8 exceeded that of powdered MoS2.  相似文献   

15.
An evaluative investigation of the Fischer–Tropsch performance of two catalysts (20%Co/Al2O3 and 10%Co:10%Mo/Al2O3) has been carried out in a slurry reactor at 2 MPa and 220–260 °C. The addition of Mo to the Co-catalyst significantly increased the acid-site strength suggesting strong electron withdrawing character in the Co-Mo catalyst. Analysis of steady-state rate data however, indicates that the FT reaction proceeds via a similar mechanism on both catalysts (carbide mechanism with hydrogenation of surface precursors as the rate-determining step). Although chain growth, , on both catalysts were comparable (  0.6), stronger CH2 adsorption on the Co-Mo catalyst and lower surface concentration of hydrogen adatoms as a result of increased acid-site strength was responsible for the lower individual hydrocarbons production rate compared to the Co catalyst. The activation energy, E, for Co (96.6 kJ mol−1), is also smaller than the estimate for the Co-Mo catalyst (112 kJ mol−1). Transient hydrocarbon rate profiles on each catalyst are indicative of first-order processes, however the associated surface time constants are higher for alkanes than alkenes on individual catalysts. Even so, for each homologous class, surface time constants for paraffins are greater for Co-Mo than Co, indicative that the adsorption of CH2 species on the Co-Mo surface is stronger than on the monometallic Co catalyst.  相似文献   

16.
A generally applicable stoichiometric and kinetic model was developed for the polymerization of dimer fatty acids with ethylenediamine in the presence of phosphoric acid as catalyst. Rate equation used is based on second order and the temperatures are between 405 and 475 K with mixing rate of 75 rpm. The parameters of the rate equations are determined using nonlinear regression analysis. Comparison of the model predictions with the experimental data show that the approach is useful in predicting the polymerization kinetic. Equilibrium constant changes from 2.432 to 17.765. Frequency factor and activation energy for forward rate constant are 362306102.681 kg mol−1 min−1 and 83.24 kJ mol−1, respectively. The equilibrium constant is independent of temperature with frequency factor and activation energy values of 17317.97 and 28.65 kJ mol−1, respectively.  相似文献   

17.
The degradation of high-density polyethylene (HDPE) was studied alone and in presence of silicoaluminophosphate type silicoaluminophosphate (SAPO-37) as catalyst. This material was synthesized by the hydrothermal method using tetrapropylammonium hydroxide and tetramethylammonium chloride as organic templates. The characterization by X-ray diffraction, infrared spectroscopy, thermogravimetry and scanning electron microscopy showed that typical faujasite structure for the SAPO-37 was obtained. The total acidity, determined by n-butylamine adsorption, it was equivalent to 0.558 mmol g−1, corresponding to moderate acid strength. For catalytic reaction, a physical mixture of 25%SAPO-37/HDPE was decomposed in a thermobalance at heating rates of 5, 10 and 20 °C min−1, from 380 to 520 °C. At the maximum degradation rate, the products were collected in a cold trap and analyzed by a coupled gas chromatograph/mass spectrometer. The degradation of HDPE without catalyst was carried out at the same conditions for comparison with the obtained data with SAPO-37. The HDPE alone suffers decomposition to a wide range of hydrocarbons (C5–C25) while in the presence of catalyst, light hydrocarbons (C2–C12) were obtained. By the application of the Vyazovkin model-free kinetic method, it was observed that the activation energy decreased from 290 kJ mol−1 for HDPE alone, to 220 kJ mol−1 for 25%SAPO-37/HDPE, evidencing that SAPO-37 is an effective catalyst for polyethylene degradation.  相似文献   

18.
-Alumina-supported MFI zeolite membranes were modified by on-stream catalytic thermal cracking of methyldiethoxysilane (MDES) molecules inside the zeolitic channels during the separation of H2/CO2 gas mixture at 450 °C and atmospheric pressure. The MDES vapor was carried by the H2/CO2 feed gas and the effect of modification was monitored continuously through online analysis of the permeate stream. The modified membrane exhibited a significant increase in H2 selectivity over CO2 with a moderate decrease in H2 permeance. At 450 °C, the modified MFI membrane obtained a H2/CO2 permselectivity of 17.5 with H2 single gas permeance of 1.86 × 10−7 mol m−2 s−1 Pa−1 as compared to a permselectivity of 2.78 and permeance of 2.75 × 10−7 mol m−2 s−1 Pa−1 for the membrane before modification. The modified membrane also showed good performance and stability in separation of H2/CO2 gas mixture containing up to 28.4% water vapor at 450 °C and atmospheric pressure.  相似文献   

19.
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   

20.
J.M.G Cowie  G.H Spence 《Polymer》1998,39(26):7139-7141
Gels of crosslinked β-cyclodextrin have been prepared using dimethylacetamide containing lithium, sodium and potassium triflate salts.

Compositions were adjusted to produce materials with dry surfaces that showed no evidence of solvent leakage. Alternating current conductivity (σ) measurements of ion transport in these systems were made over the temperature range 290–360 K. Systems containing KCF3SO3 exhibited the best range of conductivity values from σ=10−4 S cm−1 (293 K) to σ=1.8×10−3 S cm−1 (360 K). These systems also show a linear dependence of log conductivity on 1/temperature, with activation energies for ion transport in the range 32–48 kJ mol−1.  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