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.
The effects of temperature and Fe loading over multi wall carbon nano tube catalysts in thermal catalytic cracking of naphtha to produce light olefins have been studied in this paper. The CCD method was utilized and a set of experiments were designed and carried out. The temperature and loading varied from 572 to 628 °C and 0.34 to 11.66 wt.% Fe, respectively. In order to determine the effects of the variation of the operating conditions on the yield distributions, a set of statistical models were utilized and the maximum point of the yield of each product was determined. The maximum yield of ethylene (18.84 wt.% of product) and propylene (12.85 wt.% of product) was obtained at 628 °C and 10.6 wt.% loading of Fe over CNTs. Finally, thermal cracking of naphtha was carried out and was compared with thermal catalytic cracking of naphtha. As a result, at 620 °C, the yield of ethylene and propylene in thermal-catalytic cracking was 6.3% and 4.7%, respectively, more than those in thermal cracking of naphtha.  相似文献   

2.
《Comptes Rendus Chimie》2015,18(11):1205-1210
Nickel–aluminium and magnesium–aluminium hydrotalcites were prepared by co-precipitation and subsequently submitted to calcination. The mixed oxides obtained from the thermal decomposition of the synthesized materials were characterized by XRD, H2-TPR, N2 sorption and elemental analysis and subsequently tested in the reaction of methane dry reforming (DRM) in the presence of excess of methane (CH4/CO2/Ar = 2/1/7). DMR in the presence of the nickel-containing hydrotalcite-derived material showed CH4 and CO2 conversions of ca. 50% at 550 °C. The high values of the H2/CO molar ratio indicate that at 550 °C methane decomposition was strongly influencing the DRM process. The sample reduced at 900 °C showed better catalytic performance than the sample activated at 550 °C. The catalytic performance in isothermal conditions from 550 °C to 750 °C was also determined.  相似文献   

3.
It was studied that the influences of the aging, temperature, addition of the polymer and cosolvent on the stability of the vesicles spontaneously formed in the mixtures of zwitterionic surfactant (dodecyl carboxyl betaine, C12BE) and double-tailed anionic surfactant (sodium bis(2-ethylhexyl) sulfosuccinate, AOT) under the inducement of salt by means of freeze-fracture and negative-staining transmission electron microscopy (TEM), dynamic light scattering (DLS) and turbidity measurements. It is found that the vesicles can exist over a long period of aging (about 300 days) at room temperature, show good stability after a heating–cooling cycle of 90–25 °C and a freeze–thaw cycle of −10 to 25 °C, respectively, and may be transformed from spherical vesicles to tubelike structures induced by high temperature 90 °C. Under the effect of (PEO)13(PPO)30(PEO)13 (L64), the transition from unilameller vesicles to large multivesicular vesicles. The presence of ethanol may decrease the stability of vesicles, resulting in the fusion among vesicles to form large vesicles. The excessive amount of ethanol may destroy the vesicles, and the order of ability of destroying vesicles was obtained to be C5H11OH > C4H9OH > C3H7OH > C2H5OH > CH3OH.  相似文献   

4.
Activated carbon prepared from silk fibroin, which is free of metal elements, showed a high catalytic activity for the oxygen-reduction reaction (ORR). The activated carbon had a very high onset potential of Eonset = 0.83 V (vs. RHE) in oxygen-saturated 0.5 M H2SO4 at 60 °C. The ORR on the activated carbon proceeded by a four-electron process in the high-electrode-potential region; this gradually decreased to a 3.5-electron reaction below about 0.6 V (vs. RHE). Only about 1% of nitrogen atoms (mostly quaternary) remained in the activated carbon by heat-treatment at up to 1200 °C are responsible for the high catalytic activity. The open circuit voltage of a polymer electrolyte fuel cell using the activated carbon as the cathode and a platinum/carbon black anode under pure oxygen and hydrogen gases, respectively, both at one atmosphere, was 0.96 V at 27 °C.  相似文献   

5.
A continuous flow reactor was operated at 420 °C and feed rate of 0–1.5 kg h−1 for catalytic degradation of polyethylene (PE) over SA-1 silica–alumina in order to investigate the effect of catalyst on the reaction rate and the quantity and quality of degradation products. SA-1 was either mixed with the PE inside reactor or placed in a catalyst cage, the efficiency being slightly higher in the first case. The catalyst did not have a significant effect on the reaction rates but the volatile products clearly had lower molecular weights. More gases were produced on SA-1 compared to thermal degradation, containing higher amounts of C4 and less amounts of C2 compounds.  相似文献   

6.
《Comptes Rendus Chimie》2014,17(7-8):672-680
Experimental studies on diesel soot oxidation under a wide range of conditions relevant for modern diesel engine exhaust and continuously regenerating particle trap were performed. Hence, reactivity tests were carried out in a fixed bed reactor for various temperatures and different concentrations of oxygen, NO2 and water (300–600 °C, 0–10% O2, 0–600 ppm NO2, 0–10% H2O). The soot oxidation rate was determined by measuring the concentration of CO and CO2 product gases. The parametric study shows that the overall oxidation process can be described by three parallel reactions: a direct C–NO2 reaction, a direct C–O2 reaction and a cooperative C–NO2–O2 reaction. C–NO2 and C–NO2–O2 are the main reactions for soot oxidation between 300 and 450 °C. Water vapour acts as a catalyst on the direct C–NO2 reaction. This catalytic effect decreases with the increase of temperature until 450 °C. Above 450 °C, the direct C–O2 reaction contributes to the global soot oxidation rate. Water vapour has also a catalytic effect on the direct C–O2 reaction between 450 °C and 600 °C. Above 600 °C, the direct C–O2 reaction is the only main reaction for soot oxidation. Taking into account the established reaction mechanism, a one-dimensional model of soot oxidation was proposed. The roles of NO2, O2 and H2O were considered and the kinetic constants were obtained. The suggested kinetic model may be useful for simulating the behaviour of a diesel particulate filter system during the regeneration process.  相似文献   

7.
Zijuan tea theabrownins (ZTTBs) was extracted from a type of fermented Zijuan tea and separated into fractions according to molecular weight. The extract was found to contain predominantly two fractions: <3.5 kDa and >100 kDa. These two fractions were analyzed for chemical composition, structural characteristics by Curie-point pyrolysis–gas chromatography–mass spectroscopy (CP-Py–GC/MS). The affects of pyrolysis temperature on pyrolytic products were also investigated. The fraction >100 kDa produced 50 GC/MS peaks during pyrolysis at 280 °C, 70 peaks at 386 °C, and 134 peaks at 485 °C. Fourteen of the products formed at 280 °C, 12 of those formed at 386 °C, and 21 of those formed at 485 °C were identified with match qualities of greater than 80%. The fraction <3.5 kDa gave 51 peaks during pyrolysis at 280 °C, 99 peaks at 386 °C, and 257 peaks at 485 °C. Six products formed at 280 °C, four products formed at 386 °C, and 61 products formed at 485 °C were identified with match qualities of greater than 80%. Pyrolysis temperatures of 485 °C and 386 °C were found suitable for the two fractions respectively. CP-Py–GC/MS revealed that, the fraction >100 kDa mainly consisted of phenolic pigments, esters, proteins, and polysaccharides, while the fraction <3.5 kDa contained no polysaccharide. CP-Py–GC/MS is an effective tool for the composition difference and structural characteristics of ZTTBs as well as other complex macromolecular plant pigments.  相似文献   

8.
Group 4 [η13-tert-butyl(dimethylfluorenylsilyl)amido]dimethyl complexes [t-BuNSiMe2Flu]MMe2 (M = Ti, 1; Zr, 2; Hf, 3) were synthesized in a one-pot synthesis starting from the ligand, MeLi and MCl4 (M = Ti, Zr, Hf), respectively. The structures of these complexes were determined by X-ray crystallography and the results obtained revealed that the fluorenyl ligand coordinates to center metal in a η3-manner irrespective of center metal employed. Propylene polymerization was conducted at 0 or 20 °C in toluene by 13 combined with dried methylaluminoxane (MAO), which was prepared from the toluene solutions of MAO by removing free trialkylaluminiums, and HNMe2PhB(C6F5)4 in the presence of triisobutylaluminium. The 1–dried MAO system gave the polymer with syndiotactic triad (rr) of 63% at 0 °C, whereas 2 and 3 did not give any polymer in the same conditions. The 2–dried MAO system gave the polymer with the highest syndiotacticity (rr = 97%) at 20 °C, although the activity was low. The 3–dried MAO system did not give any polymer even at 20 °C. When HNMe2PhB(C6F5)4 was used in place of dried MAO at 20 °C, 1 gave almost atactic polymer, while 2 and 3 gave highly syndiotactic one (rr > 90%). These results indicate that the catalytic performance strongly depended on the center metal of the ansa-fluorenylamidodimethyl complexes as well as cocatalysts employed.  相似文献   

9.
Gorse species (Ulex sp.) are ubiquitous in the shrublands of NW Spain and have the potential to become key players in an integral biofuel/biochar program in NW Spain. Here we present molecular characterization (using pyrolysis–GC/MS) of a biochar “thermosequence” obtained by laboratory heating of Ulex europaeus wood in a muffle furnace between 200 and 600 °C (TCHAR). Low temperature chars (TCHAR  350 °C) produced significant amounts of pyrolysis products of which the precursor biopolymer could be recognized, while high-temperature chars (TCHAR  400 °C) produced mainly phenols and monocyclic and polycyclic aromatic hydrocarbons, which are not specific for any biopolymer. Carbohydrate could hardly be recognized at TCHAR  350 °C. The thermal rearrangement of polyphenols, mainly lignin, was reflected in more detail (1) C3-side chain shortening and probably depolymerization (TCHAR 200–350 °C), (2) demethoxylation of syringyl and probably also some guaiacyl lignin (TCHAR 300–400 °C), (3) elimination of virtually all remaining methoxyl groups (TCHAR 350–400 °C), through dehydroxylation and demethoxylation, (4) almost complete dehydroxylation of lignin and other biopolymers (TCHAR 400–500 °C), (5) progressive condensation into polyaromatic structures (TCHAR 300–500 °C) and (6) partial elimination of alkyl bridges between (poly)aromatic moieties (TCHAR 450–500 °C). These results were supported by Fourier transform infrared spectroscopy (FTIR) of the same samples. We conclude that pyrolysis–GC/MS can be used as a rapid molecular screening method of gorse-derived biochar. Molecular properties elucidation is an essential part of predicting the stability and agronomical behavior of gorse-derived biochar after future implementation in soils.  相似文献   

10.
A silicon carbide-based membrane was formed in the macropores of an α-alumina support tube by chemical vapor deposition of triisopropylsilane at 700–800°C with a forced cross-flow through the porous wall. The membrane permeated gases except H2O mainly by the Knudsen diffusion mechanism at permeation temperatures of 50–400°C. The H2/H2O selectivity was near or below unity because of the hydrophilic nature of the membrane. After a heat-treatment in Ar at 1000°C for 1 h, however, the membrane formed at a final evacuation pressure of 1 kPa exhibited a H2/H2O selectivity of 3–5, for a mixed feed of H2–H2O–HBr system, associated in a thermochemical water-splitting process. The H2 permeance was (5–6)×10−7 mol m−2 s−1 Pa−1 at 50–400°C. The membrane maintained the H2/H2O selectivity for more than 100 h in the H2–H2O–HBr mixture at 400°C.  相似文献   

11.
The gas release properties and char structural evolution during the pyrolysis of cotton stalk were investigated. The evolution characteristics of volatile products were examined by pyrolysis–Fourier transform infrared spectroscopy (FTIR)/thermal conductivity detection (TCD) analysis (Py–FTIR/TCD). The char chemical structure and physical characteristics were investigated by means of FTIR and N2 physisorption techniques. Evolution characteristics of the main volatile products were given. The evolution of CO2 was approximately 26 °C earlier than that of CO. CH4 evolution covered over a wider temperature range of 300–600 °C, with a maximum at 394 °C. The amount of hydroxyl, aliphatic CH and olefinic CC bonds in the char decreased significantly above 250 °C. The aromatization process started at ≈350 °C and continued to higher temperatures, leaving the char enriched with condensed aromatic ring systems. The BET surface area increased continually with increasing temperature to reach a maximum value of 4.68 m2/g at 500 °C and decreased at higher temperatures. The micropore volume showed a similar behavior to the surface area, while the mesopore volume and total pore volume always increased.  相似文献   

12.
This paper emphasises the electrochemical and catalytic properties of a Ni–10% GDC (10% gadolinium-doped ceria) cermet anode of a single-chamber solid oxide fuel cell (SC-SOFC). Innovative coupling of electrochemical impedance spectroscopy with gas chromatography measurements was carried out to characterise the anode material using an operando approach. The experiments were conducted in a symmetric anode/electrolyte/anode cell prepared by slurry coating resulting in 100 μm-thick anode layers. The electrochemical performance was assessed using a two-electrode arrangement between 400 °C and 650 °C, in a methane-rich atmosphere containing CH4, O2 and H2O in a 14:2:6 volumetric ratio. The insertion of a Pt–CeO2 based catalyst with high specific surface area inside the cermet layer was found to promote hydrogen production from the Water Gas Shift reaction and consequently to improve the electrochemical performances. Indeed, a promising polarisation resistance value of 12 Ω cm2 was achieved at 600 °C with a catalytic loading of only 15 wt.%.  相似文献   

13.
A commercial FCC catalyst based on a zeolite active phase has been used in the catalytic pyrolysis of HDPE. The experimental runs have been carried out in a conical spouted bed reactor provided with a feeding system for continuous operation. Different treatments have been applied to the catalyst to improve its behaviour. This paper deals with the optimization of catalyst steaming and pyrolysis temperature in order to maximize the production of diesel-oil fraction. The performance of the fresh catalyst has been firstly studied at 500 °C. This catalyst gives way to 52 wt% gas yield, 35 wt% light liquid fraction and a low yield of C10+ fraction (13 wt%). After mild steaming (5 h at 760 °C) the results show a significant improvement in product distribution. Thus, gas yield decreases to 22 wt%, the yield of light liquid is similar to that of the fresh one (38 wt%), whereas the yield of the desired C10+ fraction increases to 38 wt%. Nevertheless, the best results have been obtained when a severe steaming is applied to the catalyst (8 h at 816 °C) and pyrolysis temperature is reduced to 475 °C. There is a significant reduction in the gaseous fraction (8 wt%). The light liquid fraction has also been reduced to 22 wt%, but the yield of diesel fraction increases to 69 wt%. Moreover, the deactivation of the catalyst has also been studied under the optimum conditions.  相似文献   

14.
Electrochemical lithium intercalation within graphite from 1 mol dm 3 solution of LiClO4 in propylene carbonate (PC) was investigated at 25 and − 15 °C. Lithium ions were intercalated into and de-intercalated from graphite reversibly at − 15 °C despite the use of pure PC as the solvent. However, ceaseless solvent decomposition and intense exfoliation of graphene layers occurred at 25 °C. The results of the Raman spectroscopic analysis indicated that the interaction between PC molecules and lithium ions became weaker at − 15 °C by chemical exchange effects, which suggested that the thermodynamic stability of the solvated lithium ions was an important factor that determined the formation of a solid electrolyte interface (SEI) in PC-based solutions. Charge–discharge analysis revealed that the nature of the SEI formed at − 15 °C in 1 mol dm 3 of LiClO4 in PC was significantly different from that formed at 25 °C in 1 mol dm 3 of LiClO4 in PC containing vinylene carbonate, 3.27 mol kg 1 of LiClO4 in PC, and 1 mol dm 3 of LiClO4 in ethylene carbonate.  相似文献   

15.
A general and efficient procedure for the stereoselective synthesis of (E)-(1-propenyl)phenyl esters from readily accessible allylphenols has been developed. The process involves a two-step sequence consisting of the initial acylation of the allylphenols with an acid chloride, followed by catalytic CC bond isomerization in the resulting allylphenyl esters. The latter step was performed in methanol at 80 °C using catalytic amounts (0.5 mol %) of the commercially available bis(allyl)-ruthenium(IV) dimer [{RuCl(μ-Cl)(η33-C10H16)}2] (C10H16=2,7-dimethylocta-2,6-diene-1,8-diyl). Reactions proceeded in high yields (68–93%) and short times (4–9 h) with complete E-selectivity.  相似文献   

16.
The magnetic barium ferrite (BaFe12O19) hollow fibers with a high specific surface area about 22–38 m2 g?1, diameters around 1 μm and a ratio of the hollow diameter to the fiber diameter estimated about 1/2–2/3 have been prepared by the gel-precursor transformation process. The precursor and resulting ferrite hollow fibers were analyzed by thermo-gravimetric and differential scanning calorimetry, infrared spectroscopy, scanning electron microscopy and X-ray diffraction. The specific surface area was measured by the Brunauer–Emmett–Teller method. The gel formed at pH 5.5 has a good spinnability. A pure barium ferrite phase is formed after calcined at 750 °C for 2 h and fabricated of nanograins about 38 nm with a hexagonal plate-like morphology, which are increased to about 72 nm with the calcination temperature increased up to 1050 °C. The barium ferrite hollow fibers obtained at 750 °C for 2 h have a specific surface area 38.1 m2 g?1 and average pore size 6.5 nm and then the specific surface area and average pore size show a reduction tendency with the calcination temperature increasing from 750 to 1050 °C owing to the particle growth and fiber densification. These barium ferrite hollow fibers exhibit typical hard-magnetic materials characteristics and the formation mechanism for hollow structures is discussed.  相似文献   

17.
Pyrolysis of lignocellulosic biomass leads to an array of useful solid, liquid and gaseous products. Staged degasification is a pyrolysis-based conversion route to generate value-added chemicals from biomass. Because of different thermal stabilities of the main biomass constituents hemicellulose, cellulose and lignin, different temperatures may be applied for a step-wise degradation into valuable chemicals. Staged degasification experiments were conducted with deciduous (beech, poplar), coniferous (spruce) and herbaceous (straw) biomass. Thermogravimetry was used to estimate appropriate temperatures for a two-stage degradation process that was subsequently evaluated on bench-scale by moving bed and bubbling fluidised bed pyrolysis experiments. Degasification in two consecutive stages at 250–300 °C and 350–400 °C leads to mixtures of degradation products that originate from the whole biomass. The mixtures that were generated at 250–300 °C, predominantly contain hemicellulose degradation products, while the composition of the mixtures that were obtained at 350–400 °C, is more representative for cellulose. Lignin-derived fragments are found in both mixtures. Yields up to 5 wt% of the dry feedstock are obtained for chemicals like acetic acid, furfural, acetol and levoglucosan. Certain groups of thermal degradation products like C2–C4 oxygenates and phenols are formed in yields up to 3 wt%. Highest yields have been obtained for beech wood. Staged degasification is a promising pyrolysis-based route to valorise lignocellulosic biomass. Clear opportunities exist to increase product yields and selectivities by optimisation of reactor conditions, application of catalysts and specific biomass pretreatments like demineralisation and pre-hydrolysis.  相似文献   

18.
《Comptes Rendus Chimie》2015,18(3):250-260
CuO–ZnO–Al2O3 catalysts were synthesized by two methods, sol–gel and co-precipitation syntheses. Al2O3 was then substituted with other supports, such as ZrO2, CeO2 and CeO2–ZrO2 in order to have a better understanding of the support's effect. These catalysts containing 30 wt% of Cu were then tested for CO2 hydrogenation into methanol. The effect of reaction temperature and GHSV on the catalytic behaviour was also investigated. The best results were obtained with a 30 CuO–ZnO–ZrO2 catalyst synthesized by co-precipitation and calcined at 400 °C. This catalyst presents a good CO2 conversion rate (23%) with 33% of methanol selectivity, leading to a methanol productivity of 331 gMeOH.kgcata−1·h−1 at 280 °C under 50 bar and a GHSV of 10,000 h−1.  相似文献   

19.
The effect of “permanent electrochemical promotion of catalysis” (P-EPOC) was studied for the first time in the catalytic deep oxidation of C3H8 over a thin (~ 150 nm) sputtered Pt film on YSZ, under excess of oxygen at 350 °C. Short positive potential application (+ 1 V) resulted in a 5.6-fold increase of the catalytic rate, where C3H8 conversion reached 33%, while the apparent Faradaic efficiency was ~ 330. After positive current interruption the catalytic rate remained in a highly active steady-state, determined by the total charge of the anodic polarization step. Restoration of the catalytic activity to the initial value occurred only by a similar negative potential imposition. This new stable steady-state after current interruption can be interpreted by storage of a non-reactive oxygen species upon anodic polarization at the proximity of the Pt/YSZ interface and subsequent enhanced migration of spillover Oδ? species from the electrolyte support to the Pt/gas interface under open-circuit conditions.  相似文献   

20.
Pyrolysis of corncob with and without catalyst was investigated using thermogravimetry analyzer coupled with Fourier transform infrared spectroscopy (TGA–FTIR). The effects of two completely different catalysts, acid catalyst (MCM-41) and base catalyst (CaO), on the formation characteristics and composition of pyrolysis vapor were studied. The results show that these two catalysts give different product distributions. For catalytic run with MCM-41, the molality of carbonyl compounds decreases 10.2%, while that of phenols, hydrocarbons and CH4 increases 15.32%, 4.29% and 10.16% compared with non-catalytic run, respectively. The increase of phenols exhibits in a wide temperature range from about 295 °C to 790 °C in the catalytic run with MCM-41 catalyst. However, the use of CaO in pyrolysis of corncob leads to a huge change of product distribution. The molality of acids decreases 75.88%, while the molality of hydrocarbons and CH4 increases 19.83% and 51.05% compared with non-catalytic run, respectively. CaO is very effective in deacidification and the conversion of acids promotes the formation of hydrocarbons and CH4. Catalytic pyrolysis of corncob with CaO shows two main weight-loss stages. The first stage is from 235 °C to 310 °C with a weight loss of 31%. The second stage is from 650 °C to 800 °C with a weight loss of 21%.  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