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 influence of the sulfation parameters (the source and concentration of sulfate ions) and the calcination temperature on the acidic and catalytic properties of sulfated alumina in the alkylation of isobutane with butylenes and n-pentane isomerization was studied. IR spectroscopy of adsorbed probe molecules and temperature-programmed desorption of ammonia were used to characterize the acidic properties of the catalysts. An increase in the content of sulfate groups to the value corresponding to a formal value of the monolayer capacity increases the activity of alkylation and the concentration of strong Brönsted sites. The dependence of the stability of activity in alkylation on the sulfate group concentration is extreme with a maximum at the concentration close to the monolayer capacity. It was concluded from the IR spectroscopic data that the decrease in the stability of activity with the further increase in the content of sulfate groups is due to an increase in the concentration of strong Lewis sites and/or an increase in the surface density of strong Brönsted sites. The absence of the correlation between the catalytic behavior of sulfated alumina samples in n-pentane isomerization and acidity indicates that paraffin activation on these samples occurs via the non-acidic mechanism.  相似文献   

2.
A detailed FTIR study of the effects of steaming and acid leaching on protonated Y faujasite (FAU) and EMT zeolites is provided and the results are thoroughly analysed. In particular, emphasis is placed on the Brønsted acidic evolution and acidic strength measurements for a large series of as‐modified zeolites using CO as a sensitive probe to distinguish various protonic sites. While an increase of acidity for framework OH groups is observed during the strengthening of dealumination for both FAU and EMT series, the steaming process also generates a large variety of additional Brønsted acidic groups. Regarding acidic strength, these heterogeneous OH groups are sensitive to post‐treatments and their existence strongly depends on the initial composition of the zeolites. The presence of residual Na+ cations in the starting materials induces dramatic Brønsted acidic changes after steaming. As a result, steamed zeolites that initially contain traces of sodium possess unusual acidic Brønsted groups with low acidity. This result contradicts the trend generally observed with framework OH groups, for which steaming results in an increase of Brønsted acidic strength. The study reveals that the situation is indeed more complex, as some compositions and post‐treatments strongly influence the Brønsted acidity of as‐steamed zeolites both in their nature and their corresponding acidic strength. By linking these IR‐compiled features to the as‐exposed modifications, a large acidity scale better suited to characterizing catalysts having Brønsted acidity expanding from lowest to highest strength is proposed.  相似文献   

3.
In this review, a method for the temperature‐programmed desorption (TPD) of ammonia experiment for the characterization of zeolite acidity and its improvement by simultaneous IR measurement and DFT calculation are described. First, various methods of ammonia TPD are explained, since the measurements have been conducted under the concepts of kinetics, equilibrium, or diffusion control. It is however emphasized that the ubiquitous TPD experiment is governed by the equilibrium between ammonia molecules in the gas phase and on the surface. Therefore, a method to measure quantitatively the strength of the acid site (?H upon ammonia desorption) under equilibrium‐controlled conditions is elucidated. Then, a quantitative relationship between ?H and H0 function is proposed, based on which the acid strength ?H can be converted into the H0 function. The identification of the desorption peaks and the quantitative measurement of the number of acid sites are then explained. In order to overcome a serious disadvantage of the method (i.e., no information is provided about the structure of acid sites), the simultaneous measurement of IR spectroscopy with ammonia TPD, named IRMS‐TPD (infrared spectroscopy/mass spectrometry–temperature‐programmed desorption), is proposed. Based on this improved measurement, Brønsted and Lewis acid sites were differentiated and the distribution of Brønsted OH was revealed. The acidity characterized by IRMS‐TPD was further supported by the theoretical DFT calculation. Thus, the advanced study of zeolite acidity at the molecular level was made possible. Advantages and disadvantages of the ammonia TPD experiment are discussed, and understanding of the catalytic cracking activity based on the derived acidic profile is explained.  相似文献   

4.
The acidic properties of the H-forms of zeolites ZSM-5, Beta, Y, and mordenite are studied by diffuse reflectance IR spectroscopy using n-pentane as a probe molecule. The decreasing order of Brønsted acid site strengths is constructed. The isopentane selectivity in n-pentane isomerization under supercritical conditions (260°C, 130 atm) increases in the order H-ZSM-5 < H-Beta < H-mordenite(11) ≈ H-Y with decreasing strength of Brønsted sites. Catalytic data are analyzed jointly with the results of physicochemical studies of H-mordenite (temperature-programmed ammonia desorption, benzene adsorption, and IR spectroscopy). Under the supercritical conditions, the conversion of n-pentane on mordenite is determined by the total acidity of the zeolite and also by the accessibility of the acid sites inside the zeolite channels to the reactant.  相似文献   

5.
UL-ZSM-5 materials have been prepared by templated solid-state crystallization of zeolites starting from the amorphous mesostructured aluminosilicate Al-Meso. Microcalorimetry and FTIR have been employed to characterize their surface acidity. In good agreement with 27Al MAS NMR data, UL-ZSM-5 displayed an improved density and strength of Brönsted acid sites, as compared to Al-Meso, owing to the incorporation of aluminium in a tetrahedral environment similar to that of zeolite ZSM-5. Moreover, they showed an enhanced Brönsted/Lewis relative acid ratio. However, Al-Meso showed the highest concentration of strong Lewis acid sites due to its largest amount of aluminium in extraframework positions.  相似文献   

6.
A time‐resolved in situ micro‐spectroscopic approach has been used to investigate the Brønsted acidic properties of fluid‐catalytic‐cracking (FCC) catalysts at the single particle level by applying the acid‐catalysed styrene oligomerisation probe reaction. The reactivity of individual FCC components (zeolite, clay, alumina and silica) was monitored by UV/Vis micro‐spectroscopy and showed that only clay and zeolites (Y and ZSM‐5) contain Brønsted acid sites that are strong enough to catalyse the conversion of 4‐fluorostyrene into carbocationic species. By applying the same approach to complete FCC catalyst particles, it has been found that the fingerprint of the zeolitic UV/Vis spectra is clearly recognisable. This almost exclusive zeolitic activity is confirmed by the fact that hardly any reactivity is observed for FCC particles that contain no zeolite. Confocal fluorescence microscopy images of FCC catalyst particles reveal inhomogeneously distributed micron‐sized zeolite domains with a highly fluorescent signal upon reaction. By examining laboratory deactivated FCC catalyst particles in a statistical approach, a clear trend of decreasing fluorescence intensity, and thus Brønsted acidity, of the zeolite domains is observed with increasing severity of the deactivation method. By comparing the average fluorescence intensities obtained with two styrenes that differ in reactivity, it has been found that the Brønsted acid site strength within FCC catalyst particles containing ZSM‐5 is more uniform than within those containing zeolite Y, as confirmed with temperature‐programmed desorption of ammonia.  相似文献   

7.
With P(CH3)3 as the probe molecule adsorbed on titanium silicalite (TS-1) zeolite, the special and important role of T12 site in MFI-type zeolite was clearly elucidated. There are altogether three active sites present in TS-1 zeolite with Ti at the T12 site. Owing to the preferential adsorption of probe molecules on the first Brönsted acidic site, the Ti12 center will probably fail to show Lewis acidity. The ionic [HP(CH3)3]+ species can be stabilized by the first or second Brönsted acidic site, with the former energetically favored. The latter was formed through the transfer of the ionic [HP(CH3)3]+ species from the first to the second Brönsted acidic site.  相似文献   

8.
The effect of protonation on the surface acidic properties of titanate nanowires (TiONWs) was investigated. Nanowires were synthesized by the alkali hydrothermal method which resulted in one dimensional nanostructures of large external surface area and well-defined lamellar interlayer structure. The Na+/H+ ratio in the structure can be tuned by ion-exchange. Our aim was to characterize the morphology of the as-synthesized nanostructures by HRTEM and SEM measurements and assess their surface acidity using in situ infrared spectroscopic measurements and temperature programmed desorption. It was found that the numbers of Lewis and Brönsted acidic sites in the Na-form and the H-form of the TiONWs is different. The ratio and the nature of acidic sites can be tuned by the ion exchange process. The wire-like morphology and the tunable acidity are features of titanate nanowires that may render them a promising material in various heterogeneous catalytic applications.  相似文献   

9.
Zou  H.  Li  M.  Shen  J.  Auroux  A. 《Journal of Thermal Analysis and Calorimetry》2003,72(1):209-221
The surface acidity of SiO2, γ-Al2O3 and TiO2 supported vanadia catalysts has been studied by the microcalorimetry and infrared spectroscopy using ammonia as the probe molecule. The acidity in terms of nature, number and strength was correlated with surface structures of vanadia species in the catalysts, characterized by X-ray diffraction and UV-Vis spectroscopy. It was found that the dispersion and surface structure of vanadia species depend on the nature of supports and loading and affect strongly the surface acidity. On SiO2, vanadium species is usually in the form of polycrystalline V2O5 even for the catalyst with low loading (3%) and these V2O5 crystallites exhibit similar amount of Brönsted and Lewis acid sites. The 25%V2O5/SiO2 catalyst possesses substantial amount of V2O5 crystallites on the surface with the initial heat of 105 kJ mol-1 and coverage of about 600 mmol g-1 for ammonia adsorption. Vanadia can be well dispersed on g-Al2O3and TiO2 to form isolated tetrahedral species and polymeric two-dimensional network. Addition of vanadia on γ-Al2O3 results in the change of acidity from that associated with g-Al2O3 (mainly Lewis sites) to that associated with vanadia (mainly Brönsted sites) and leads to the decreased acid strength. The 3%V2O5/TiO2 catalyst may have the vanadia structure of incomplete polymeric two-dimensional network that possesses the Ti-O-V-OH groups at edges showing strong Brönsted acidity with the initial heat of about 140 kJ mol-1 for ammonia adsorption. On the other hand, the 10%V2O5/TiO2 catalyst may have well defined polymeric two-dimensional vanadia network, possessing V-O-V-OH groups that exhibit rather weak Brönsted acidity with the heat of 90 kJ mol-1 for NH3 adsorption. V2O5 crystallites are formed on the 25%V2O5/TiO2 catalyst, which exhibit the acid properties similar to those for 25%V2O5 on SiO2 and γ-Al2O3.  相似文献   

10.
Quantitative in situ infrared spectroscopy in combination with kinetic analysis is utilized to derive mechanistic aspects for the reaction of methanol with ammonia on Brønsted acidic mordenite. Under non-reactive conditions, a coadsorption complex between methanol and ammonia is found, in which only ammonia is in direct interaction with the Brønsted acid sites of the zeolite. This complex is proposed to be the precursor for the formation of protonated methylamines in the zeolite pores which are formed in sequential order up to tetramethylammonium ions. These methylamines are unable to desorb under reaction conditions in the absence of ammonia. They leave the surface either by ammonia adsorption assisted desorption or by scavenging of methyl groups from protonated methylamines by ammonia. Both steps are concluded to be potentially rate determining.  相似文献   

11.
《Microporous Materials》1995,3(4-5):457-465
The acidic properties of dealuminated Y-type zeolites were characterized by infrared (IR) spectroscopy, microcalorimetry, 29Si magic-angle spinning nuclear magnetic resonance (MAS NMR) spectroscopy and temperature-programmed desorption (TPD). Microcalorimetric measurements exhibited a uniform heat of adsorption (140 kJ/mol) of ammonia on the strong Brönsted acid sites. The differences in the acid site concentrations measured by adsorption of ammonia from the gas phase and by decomposition of ammonium-exchanged zeolites are discussed. The results indicate that parts of the extra lattice material consisting of cationic aluminium oxide species or silica alumina species are removed by ion exchange with aqueous solutions of ammonium hydroxide. Based on this, a method for the controlled removal of extra lattice material was developed.  相似文献   

12.
The acidity of the surfaces of rhodium promoted zirconium oxide catalysts (Rh-Cr2O3/ZrO2 and Rh-CeO2/ZrO2) has been investigated by the thermally programmed desorption of ammonia (TPDA) method and IR spectroscopy. A correlation between the acid properties of the surfaces of the catalysts (strength and type of acid centers, presence of Brönsted acid centers) and their activity in the SCR process for the reduction of NO with a propane–butane mixture and propene has been established.  相似文献   

13.
Studies on Oxide Catalysts. XXV. Catalytic Activity and Aging Properties of Modified Mordenites in the Cracking of n-Octane MeH-mordenites (Me = Li, K, Mg, Ca, Ba) were prepared by ion exchange starting with H-mordenite (SiO2/Al2O3 mole ratio = 14). To characterize these samples the cracking of n-octane was used as catalytic test reaction. Surface OH groups and the adsorption of NH3 on these samples were investigated by i. r. spectroscopy. Unaffected by the kind of the exchanged cation the Brönsted acidity of the H-mordenite decreases monotonously with increasing content of the incorporated cation. The catalytic activity and (to a much higher degree) the rate of deactivation by coking during the reaction decrease as the Brönsted acidity decreases. The strong dependence of the Brönsted acidity on the deactivation rate points to a multi-site mechanism of the coking process.  相似文献   

14.
Using trimethylphosphine (TMP) and d5-pyridine(deuterated pyridine) as the basic probe molecules, the concentrations of Brönsted acid sites on both HY zeolite and dealuminated HY zeolite have been quantitatively determined using solid-state 1H and 31P magic-angle spinning (MAS) NMR. After adsorption of the probe molecules, the concentration of Brönsted acid sites on the dealuminated HY zeolite increases by about 25%, whereas that in the parent HY sample remains almost unchanged. The increase in the concentration of Brönsted acid sites is due to the appearance of base-induced Brönsted acid sites in the dealuminated HY zeolite. The terminal SiOH in the vicinity of the aluminum atom is “induced” to form a bridging hydroxyl group (SiOHAl) in the presence of the basic probe molecules. The mechanism of formation of the induced Brönsted acid sites has also been discussed.  相似文献   

15.
Novel magnesium fluorides have been prepared by a new fluorolytic sol–gel synthesis for fluoride materials based on aqueous HF. By changing the amount of water at constant stoichiometric amount of HF, it is possible to tune the surface acidity of the resulting partly hydroxylated magnesium fluorides. These materials possess medium‐strength Lewis acid sites and, by increasing the amount of water, Brønsted acid sites as well. Magnesium hydroxyl groups normally have a basic nature and only with this new synthetic route is it possible to create Brønsted acidic magnesium hydroxyl groups. XRD, MAS NMR, TEM, thermal analysis, and elemental analysis have been applied to study the structure, composition, and thermal behaviour of the bulk materials. XPS measurements, FTIR with probe molecules, and the determination of N2/Ar adsorption–desorption isotherms have been carried out to investigate the surface properties. Furthermore, activity data have indicated that the tuning of the acidic properties makes these materials versatile catalysts for different classes of reactions, such as the synthesis of (all‐rac)‐[α]‐tocopherol through the condensation of 2,3,6‐trimethylhydroquinone (TMHQ) with isophytol (IP).  相似文献   

16.

In this paper, an interesting Brönsted acidic dicationic ionic liquid based on piperazine was prepared via two-step synthesis. The structure of the dicationic ionic liquid (DIL) was confirmed with various techniques including Fourier transform infrared spectroscopy (FT-IR), Proton NMR (1H NMR), Carbon-13 NMR (13C NMR), mass spectroscopy (MS), acid–base titration as well as Hammett acidity function. Catalytic activity of this novel Brönsted acidic DIL was examined in the xanthene synthesis. High yield of product, short reaction time, easily recovery and reusing of the catalyst, and the absence of organic solvents are some of the merits of the developed procedure.

  相似文献   

17.
The overall view of the TPD of ammonia to measure the acidic property of zeolites is described. The desorption peaks were identified and the significance of readsorption of ammonia was pointed out for the first time. This part of the work was done using reference catalysts of the Catalysis Society of Japan. The theoretical equation for the TPD with free readsorption of ammonia was then derived. Two methods for determining the strength of zeolite acidity based on the derived equation were proposed. A curve fitting method was then proposed to determine the zeolite acidity; based on this method, not only the strength of acidity but also its distribution could be determined. This method was applied to mordenite and ZSM-5 zeolites with different contents of Al and Na cations, and a simple conclusion was reached; namely, the strength of the acidity was not influenced by the number of acid sites but by the structure of the zeolite. Finally, water vapor treatment to rub out the l-peak (lower temperature peak) was briefly mentioned. This method was applied to precisely determine the acidity of Y-zeolite. A case study about the beta zeolite as the catalyst for the amination of phenol was exemplified; the catalytic activity was discussed in terms of the measured acidity.  相似文献   

18.
Silica spheres with uniform size of 230–250 nm were functionalized with sulfonic groups and bi-functionalized with carboxylic and sulfonic groups via a co-condensation route, by adding the organosilanes (3-(triethoxysilyl)mercaptopropyl and 4-(triethoxysilyl)butyronitrile) to a pre-hydrolized TEOS solution. The conversion of mercapto and cyano groups to, respectively, sulfonic and carboxylic groups was carried out by treating both the samples with nitric acid solution. The presence of alkyl-SO3H and alkyl-COOH species at the silica surface in an approximate molar ratio of 1:1 was assessed by TG and NMR. FT-IR spectroscopy showed that both Brønsted acidic groups are accessible and give proton-transfer reaction to ammonia with the formation of ammonium ion. Sulfonic groups react irreversibly with ammonia at room temperature at variance with carboxylic groups which give a reversible proton-transfer, in agreement with the stronger Brønsted acidity of the former.  相似文献   

19.
HZSM‐5‐supported Brönsted and Lewis acidic ionic solid 1,3‐disulfoimidazolium chlorozincate materials ([dsim]2[ZnCl4]@HZSM‐5) were synthesized with various ratios (3, 6, 9, 17 and 50% w/w). The heterogeneous materials were characterized via a variety of spectroscopic techniques. Dual acidity of these materials was determined using specified techniques. These acidic solids were examined as stable heterogeneous catalysts for the Fischer indole reaction of equimolar amounts of phenylhydrazine hydrochloride and various aliphatic or aromatic ketones at 80–90°C in neat condition to produce substituted indole derivatives. The efficient 17% ionic salt‐loaded HZSM‐5 composite was easily reused for ten consecutive cycles with a slight loss of its activity. The recycled catalyst was further analysed using powder X‐ray diffraction and inductively coupled plasma optical emission spectrometric techniques.  相似文献   

20.
To date, investigations onto the regulation of reactants mass transfer has been paid much less attention in environmental catalysis. Herein, we demonstrated that by rationally designing the adsorption sites of multi-reactants, the pollutant destruction efficiency, product selectivity, reaction stability and secondary pollution have been all affected in the catalytic chlorobenzene oxidation (CBCO). Experimental results revealed that the co-adsorption of chlorobenzene (CB) and gaseous O2 at the oxygen vacancies of CeO2 led to remarkably high CO2 generation, owning to their short mass transfer distance on the catalyst surface, while their separated adsorptions at Brönsted HZSM-5 and CeO2 vacancies resulted in a much lower CO2 generation, and produced significant polychlorinated byproducts in the off-gas. However, this separated adsorption model yielded superior long-term stability for the CeO2/HZSM-5 catalyst, owning to the protection of CeO2 oxygen vacancies from Cl poisoning by the preferential adsorption of CB on the Brönsted acidic sites. This work unveils that design of environmental catalysts needs to consider both of the catalyst intrinsic property and reactant mass transfer; investigations of the latter could pave a new way for the development of highly efficient catalysts towards environmental pollution control.  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