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.
An improved dealumination method for adjusting the acidity of HZSM-5 was developed by utilizing the self-adsorbed water in HZSM-5. The parent and treated HZSM-5 were characterized by XRD, FT-IR, 27Al MAS NMR, N2 adsorption, XRF and NH3-TPD. The results showed that HZSM-5 can be dealuminated when treated at above 400 °C. About 80% acidic amount was removed from parent HZSM-5 with 8% adsorbed water and more Lewis acid sites were produced after treated at 500 °C. It is thought that the dealumination was mainly caused by its self-adsorbed water. The comparisons of the acidity in dealuminated HZSM-5 by traditionally steaming and improved methods indicated that the latter was more effective in decreasing the acidity and weakening acid strength, and more environmentally benign and timesaving. The method is also applicable to adjust acidity of other zeolites, such as HY, Hβ.  相似文献   

2.
Palladium-based catalysts were prepared by the impregnation (I) and ion-exchange method (E) with ZSM-5 and γ-Al2O3 as support respectively. The high activity of Pd/ZSM-5(I) and Pd-ZSM-5(E) catalysts for methane combustion was observed. The order of activity is consistent with Brønsted acidity of catalysts: Pd/ZSM-5(I) > Pd-ZSM-5(E) > Pd/Al2O3. It is shown by FT-IR that methane adsorbs on acidic bridging hydroxyl groups of ZSM-5-supported Pd catalysts. Symmetric v1 C–H stretching vibrations of methane shift to low frequency due to the interaction between methane molecules and Brønsted acid sites or Pd2+, indicating that methane molecules can be activated.  相似文献   

3.
Fe/ZSM-5 catalysts with various morphologies and sizes were prepared and the catalytic properties in NH3-SCR were also investigated. The different ZSM-5 morphologies and sizes indeed influence the dispersion of Fe species. The Fe/ZSM-5 catalyst, which was cauliflower-like morphology of ZSM-5 support aggregated by small nano-crystal zeolite with crystallite size of about 50 nm, exhibited the best NH3-SCR activity (T 90% = 280–650 °C). This specific morphology and size of ZSM-5 support were considered to benefit the distribution of isolated Fe3 + species, which was proved to be the main active sites in SCR reaction.  相似文献   

4.
A simple method for evaluating the surface acidity of different cation-exchanged montmorillonite (mont) clay catalysts, Mn+-mont (Mn+=Al3+, Fe3+, Cr3+, Zn2+, Ni2+, Cu2+, and H+), involving treatment with pyridine is described. After treating with pyridine, the samples were heated at 120 °C and the FT-IR spectra were directly recorded in the region 1650 and 1350 cm−1. The data obtained show the presence of both Lewis and Brønsted acid sites. The activities of the catalysts to bring about Brønsted acid catalysed esterification of succinic acid with iso-butanol to yield di-(iso-butyl) succinate have been studied. The Brønsted acidity data obtained for Mn+-mont correlated well with activity in the esterification reaction. The activities of the catalysts were found to decrease in the order of exchange ions Al3+ > Fe3+ > Cr3+ > Zn2+ > Ni2+ > Cu2+ > Na+-mont. They also correlated well with the charge to radius ratio of the cations. The catalysts exchanged with trivalent cations showed stronger absorption bands attributed to Brønsted acidity (1540 cm−1) whereas those exchanged with divalent cations showed an increased Lewis acidity (1450 cm−1) and reduced Brønsted acidity along with charge to radius ratio. Zn2+-, Cu2+- and Ni2+-exchanged clays showed an additional peak around 1605 cm−1 which is attributed to the pyridine adsorption on surface sites through its π electrons. The method suggested here to evaluate the acidity is suitable for active sites which are thermally unstable such as water molecules in the hydration shell of a cation in exchanged clay.  相似文献   

5.
Nickel was introduced in ZSM-5 zeolite by two different methods: dry impregnation and ionic exchange. Different loadings of metal, ranging from 0.4 to 6 wt% were explored. These materials were thoroughly characterized by EXAFS, TPR, acidity measurements by H/D isotope exchange and ethane hydrogenolysis. Regardless of the metal introduction method, at 0.4 wt% Ni, the Ni-ZSM-5 catalysts present only nickel located inside the zeolite channels as compensation cations. In contrast, an increase to 1 wt% nickel (via impregnation) led to its presence both inside and outside the channels.The catalytic activity of these Ni-ZSM-5 zeolites was tested in n-hexane cracking. Depending on the way the metal was introduced, it was possible to modify the n-hexane cracking activity and the selectivity toward light alkenes. Hence, a proper design of metal and acid sites could be achieved.  相似文献   

6.
The hydrocracking and hydroisomerization of n-octane, 2,5-dimethylhexane and 2,2,4-trimethylpentane on Ni-containing ZSM-5, mordenite and beta catalysts was investigated at 20 bar hydrogen pressure and a temperature of 533 K. The activity decreased in the sequence ZSM-5≫BETA≈MOR for the n-octane conversion and increased in this sequence for the conversion of 2,5-dimethylhexane and 2,2,4-trimethylpentane. The selectivity for isomerization of n-octane and 2,5-dimethylhexane was the highest on NiHBETA and the lowest on NiHZSM-5. The trends in the activity and selectivity were explained by the accessibility of the acid sites, estimated by adsorption of probe molecules followed by IR spectroscopy, and by a simulation of the space available in the pores of these zeolites.  相似文献   

7.
《Fuel》2005,84(2-3):259-269
Kuwait atmospheric residue (KAR) was hydrodemetallized (HDM) over a commercial Mo/Al2O3 catalyst using a fixed-bed reactor unit. Initial coke deposition on the catalyst vs. time-on-stream (TOS) was studied by solid-state 13C NMR to derive structural information on the initial coke deposited in the first 240 h of TOS. 13C NMR spectral editing (cross-polarization technique in combination with the polarization inversion pulse sequence) was applied to distinguish on the one hand between quaternary and tertiary aromatic carbon and on the other hand between secondary and primary aliphatic carbon. Structural parameters, derived from those NMR measurements in combination with the aromaticity from single pulse excitation 13C NMR and the H/C ratio from elemental analysis, have been used to assemble the hypothetical structure of initial coke.The 1 h coke has nearly the same H/C ratio as the KAR and represents an intermediate between heavy coke generating molecules from KAR and real coke. In the following 12 h, aromatic carbon is accumulated and aromatic rings are dealkylated but the coke is still rich in hydrogen. As the carbon deposition slows down (after 50 h) further aromatic carbon is built-up and aliphatic carbon decreases. After 120 h the hydrogenation activity of the HDM catalyst is still high and prevents the coke from getting hydrogen depleted, merely the degree of alkyl-substitution has dropped compared with the start of run. With further increase of run time, carbon deposition reaches a steady state (240 h). As the analysis of a spent HDM catalyst (6500 h) from an industrial ARDS unit shows, throughout the following months of operation the porous hydrogen rich coke is increasingly converted into highly condensed polynuclear aromatic coke. THF-extraction of the used catalysts (TOS; 1, 12, 120, and 6500 h) has mainly removed small aromatic compounds with a high degree of alkyl-substitution.  相似文献   

8.
Copper or cobalt incorporated TiO2 supported ZSM-5 catalysts were prepared by a sol–gel method, and then were characterized by XRD, BET, XPS and UV–vis diffuse reflectance spectroscopy. Ti3 + was the main titanium specie in TiO2/ZSM-5 and Cu–TiO2/ZSM-5, which will be oxide to Ti4 + after Co was doped. With the deposition of Cu or Co, the efficiency of the CO2 conversion to CH3OH was increased under low energy irradiation. The peak production rate of CH3OH reached 50.05 and 35.12 μmol g 1 h 1, respectively. High photo energy efficiency (PEE) and quantum yield (φ) were also reached. The mechanism was discussed in our study.  相似文献   

9.
A series of Fe2O3-doped (1–5%) Pt/SO42 /ZrO2 were prepared by a co-precipitation method. The incorporation of small amounts of Fe2O3 into Pt/SO42 /ZrO2 results in an enhanced Brønsted acidity in the presence of H2, which makes the Fe2O3-doped catalysts much more active than the undoped one for n-heptane hydroisomerization. A maximal yield of C7 isomers appears at 3.5% Fe2O3.  相似文献   

10.
Photocatalytic reduction of Cr(VI) to Cr(III) in aqueous solution containing ZnO or ZSM-5 zeolite under ambient condition was studied by using oxalate as model organic compound in the natural environment. ZSM-5 zeolite was characterized by X-ray diffraction (XRD), and point of zero net proton charge (PZNPC) titration. The effect of illumination time, mass content of catalyst (m/V), Cr(VI) initial concentrations, pH, ionic strength, and oxalate concentrations on the photocatalytic reduction of Cr(VI) was determined. The results indicate that the PZNPC of ZSM-5 zeolite is at pH 3.6 ± 0.1. At C[Cr(VI)(initial)] = 2.00 × 10?4 mol/L, pH 7.5 ± 0.1 and after illumination time of 24 h, the reduction of Cr(VI) were 1.1 × 10?5 mol/L (no ZSM-5 zeolite, 4.0 × 10?3 mol/L oxalate) and 1.0 × 10?5 mol/L (0.4 g/L ZSM-5 zeolite, no oxalate), respectively; whereas the reduction of Cr(VI) achieved 1.0 × 10?4 mol/L in the presence of 0.4 g/L ZSM-5 zeolite and 4.0 × 10?3 mol/L oxalate. The removal of Cr(VI) from solution is dependent on pH value. The results are important for the application of zeolites in the treatment of Cr(VI) polluted solution in the natural environment.  相似文献   

11.
The catalytic cracking of n-dodecane over HZSM-5 zeolite catalyst was investigated at 400–450 °C under supercritical and subcritical pressures (0.1–4.0 MPa). The results show that both the activity of the catalyst and its stabilization towards deactivation decrease with increasing pressure, and the catalyst maintains substantially higher activity when feed rate exceeds 4.00 ml/min under supercritical conditions. A first-order Langmuir kinetic model with a novel decay function is developed for the supercritical catalytic cracking of hydrocarbon incorporating supercritical extraction effect on catalyst stability, which is satisfactory to describe the kinetic behaviors of catalytic cracking of supercritical n-dodecane. According to the estimated reaction rate and adsorption constant of n-dodecane on HZSM-5 at different temperature, the activation energy of 125.4 kJ/mol and adsorption heat 109.5 kJ/mol were calculated. An index of CRSE is proposed to define contribution ratio of supercritical extraction to the activity of the HZSM-5 catalyst in the developed kinetics model, and it is found that the CRSE increases with increasing hydrocarbon feed rates and decreasing catalytic activities, and reaches maximum value when the coke formation rate equals to the coke removal rate by supercritical hydrocarbon.  相似文献   

12.
A series of 2,2′-disubstituted-4,4′-dihydroxybiphenyl monomers was prepared from 3,4,5-tris(n-alkoxy)benzyl chlorides (n = 5, 6, 8, 10, 12) and tetramethylammonium salt of 4,4′-dihydroxydiphenic acid, which was synthesized from two different 5-step routes. 2,2′-Bis(trifluoromethyl)-4,4′-biphenyldicarboxylic acid was synthesized via 5-step route. A series of aromatic polyesters containing multiple alkyl side chains was prepared from the 2,2′-disubstituted-4,4′-dihydroxybiphenyl monomers and 2,2′-bis(trifluoromethyl)-4,4′-biphenyldicarboxylic acid using diisopropylcarbodiimide as a dehydrating agent and 4-(dimethylamino)pyridinium 4-toluenesulfonate as a catalyst at room temperature. Their thermal and solution properties were measured and compared with the polyester without multiple alkyl side chains. The polyesters displayed better solubility in common solvents such as chlorinated solvents and THF but lower thermal stability than the polyester without multiple alkyl side chains. The intrinsic viscosities of the polyesters ranged from 0.68 to 2.53 dL/g and their number-average molecular weights ranged from 19,300 to 61,400. Polyesters containing C5–10 side chains were amorphous while the two polyesters containing C12 side chains crystallized at ?27 and ?31 °C, respectively. The thermal stability of the polyesters decreased as a result of alkyl side chains. The films of polyesters were opaque, indicating that the aromatic backbones and aliphatic side chains underwent phase separation.  相似文献   

13.
Selective dealumination of mordenite by high-temperature steam treatment improved its stability in dimethyl ether carbonylation. Most of the framework Al species in the 12-membered ring channels of mordenite was removed while those in the 8-membered ring channels were retained. On the resulting mordenite catalyst, the formation rate of methyl acetate was only slightly decreased and maintained at 1.71 mol (mol H+) 1 h 1 at 15 h. Temperature-programmed oxidation analyses on the used sample evidenced that the deposition of hard coke was effectively suppressed in the selectively dealuminated catalyst because of the reduced number of acidic sites in the 12-membered ring channels.  相似文献   

14.
The complete parameters of montmorillonite solid acidity, namely amount, strength, and types of acidity, were determined and the properties of the acid sites after heating were proposed by combining the temperature-programmed desorption (TPD) and Fourier transform infrared spectroscopy (FTIR) based on the NH3 adsorption system. The total amount of montmorillonite acid sites was 1.15 mmol/g, which was higher than the value obtained by the Hammett indicator method because of the detection of solid acid sites in the montmorillonite interlayer space. These acid sites were composed of 1.00 mmol/g Brønsted and 0.15 mmol/g Lewis acid sites. The acidity of montmorillonite was primarily derived from the interlayer polarized water, Si–OH, H3O+ adsorbed by negatively charged tetrahedral AlO4, and unsaturated Al3 + ions, all of which were attributed to the Brønsted acid sites with the exception of the unsaturated Al3 + ions (Lewis acid sites). Heating led to an increase in the acid strength and the acid amount and altered the type of the partial acid sites. The interlayer polarized water provided more protons after heating at 120 °C and exhibited higher acid strength than that of raw montmorillonite. After heating at 400 °C, the interlayer polarized water acted as very strong acid sites. The H3O+ adsorbed by tetrahedral AlO4 was attributed to weak-strength acid sites and transformed into Si–O(H)–Al after dehydration, while displaying strong-strength acidity. The unsaturated Al3 + ions showed medium-strength Lewis acidity, although a portion of these ions adsorbed water molecules and exhibited weak Brønsted acidity. After dehydroxylation at 600 °C, an abundance of unsaturated Al3 + ions appeared and the amount of Lewis acid sites increased.  相似文献   

15.
In this study, thermal degradation of additive-free polypropylene powder over different type of zeolite catalysts was investigated. BEA, ZSM-5 and MOR with different surface areas, pore structures, acidities and Si/Al molar ratios were used as solid catalysts for degradation of polypropylene (PP). Degradation rate of the PP over zeolites was studied by thermogravimetric analysis (TGA) employing four different heating rates and apparent activation energies of the processes were determined by the Kissinger equation. The catalytic activity of zeolites decreases as BEA > ZSM-5a (Si/Al = 12.5) > ZSM-5b (Si/Al = 25) > MOR depending on pore size and acidity of the catalysts. On the other hand, initial degradation is relatively faster over MOR and BEA than that over both ZSM-5 catalysts depending on the apparent activation energy. It can be concluded that acidity of the catalyst is the most important parameter in determining the activity for polymer degradation process as well as other structural parameters, such as pore structure and size.  相似文献   

16.
Catalytic fast pyrolysis of Jatropha residue was performed over SiO2 catalysts with different pore sizes (SiO2-Q3, -Q10, -Q30, and -Q50 with pore diameters of 3, 16, 45, and 68 nm, respectively) at 500 °C using a pyrolyzer-gas chromatography/mass spectrometry system. SiO2-Q10, which combined weak acidity and medium porosity, was the most effective catalyst in removing oxygenated compounds such as acids, ketones, and aldehydes, which are the principal reason for the polymerization of hydrocarbons from bio-oil, and in inhibiting coke and polycyclic aromatic compound formation. SiO2-Q10 is also useful for stabilizing bio-oil and has potential for catalytic fast pyrolysis.  相似文献   

17.
Decamethylosmocene, (Me5C5)2Os (1), is a pre-catalyst in a very efficient oxidation of alkanes with hydrogen peroxide in acetonitrile at 20–60 °C. The reaction proceeds with a substantial lag period that can be reduced by the addition of pyridine in a small concentration. The lag period can be removed if 1 is incubated with pyridine and/or H2O2 in MeCN prior to the alkane oxidation. Alkanes, RH, are oxidized primarily to the corresponding alkyl hydroperoxides, ROOH. Turnover numbers attain 51,000 in the case of cyclohexane (maximum turnover frequency was 6000 h?1) and 3600 in the case of ethane. The oxidation of benzene and styrene also occurs with a lag period to afford phenol and benzaldehyde, respectively. A kinetic study of cyclohexane oxidation and selectivity parameters (measured in the oxidation of n-heptane, methylcyclohexane, isooctane, cis- and trans-dimethylcyclohexanes) indicates that the oxidation of saturated, olefinic, and aromatic hydrocarbons proceeds with the participation of hydroxyl radicals. The 1/H2O2/py/MeCN system also oxidizes 1-phenylethanol to acetophenone.  相似文献   

18.
Hydroisomerization of n-pentane over platinum promoted acid zeolites was studied. The effect of structure and strength of acid sites of the support was investigated on the reaction carried out under atmospheric pressure at 300 °C. In order to accomplish the above purpose, some zeolitic catalytic systems were studied, which have the same structure but different acidity and with similar acidity but different structure.  相似文献   

19.
The performance of Pt–Sn-based catalyst, supported on ZSM-5 of different Si/Al ratios were investigated for simultaneous dehydrogenation and cracking of n-butane to produce light olefins. The catalysts were characterized by number of physio-chemical techniques including XRF, TEM, IR spectra, NH3-TPD and O2-pulse analysis. Increase in Si/Al ratio of zeolite support ZSM-5 significantly increased light olefin's selectivity, while feed conversion decreases due to lower acidity of support. The results indicated that both the n-butane cracking and dehydrogenation activity to light olefin's over Pt–Sn/ZSM-5 samples with increasing Si/Al ratios greatly enhanced catalytic performance. The catalysts were deactivated with time-on-stream due to the formation of carbon-containing deposits. A coke deposition was significantly related to catalyst activity, while at higher Si/Al ratio catalyst the coke precursors were depressed. These results suggested that the Pt–Sn/ZSM-5 catalyst of Si/Al ratio 300 is superior in achieving high total olefins selectivity (above 90 wt.%). The Pt–Sn/ZSM-5 also demonstrates resistance towards hydrothermal treatment, as analyzed through the three successive reaction-regeneration cycles.  相似文献   

20.
Using density functional calculations, we studied the conversion of methylcyclopentane to its ring-opening products: branched hexanes [2-methylpentane (2MP), 3-methylpentane (3MP)], as well as unbranched n-hexane (nHx). We employed flat Pt(1 1 1) and stepped Pt(2 1 1) to describe terrace-rich large and defect-rich small Pt particles, respectively. On Pt(1 1 1), the barriers of all elementary steps for the paths leading to branched hexanes lie below 90 kJ mol?1, while the formation of nHx features a barrier of 116 kJ mol?1 in its C–C bond scission step. This higher barrier impedes the formation of nHx on Pt(1 1 1) and thus rationalizes the experimental observations that terrace-rich large Pt particles selectively produce branched hexanes. However, on Pt(2 1 1), the barrier of C–C scission for the formation of nHx decreases to 94 kJ mol?1, thus implying enhanced formation of nHx over the defects, in agreement with the essentially statistical product distribution observed with defect-rich small Pt particles.  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