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.
Adsorption of C2H4 and C3H6 on copper in oxidized samples of CuZSM-5 is found to increase with the copper concentration; simultaneously, olefin adsorption on the Br?nsted acid sites decreases. The Cu2+ cations in the square-planar coordination exhibit higher reactivity in olefin adsorption than copper cations in the square-pyramidal coordination. Thermal treatment of CuZSM-5 with hydrogen results in regeneration of the Br?nsted acid sites for olefin adsorption and the disappearance of Cu2+ cations, the active sites of adsorption, due to the reduction of Cu2+ to Cu+ and Cu0. Desorption peaks appear in the TPD spectra upon the interaction between the adsorbed hydrocarbons and NO2. These peaks are not observed upon separate adsorption of the reactants, and they are likely due to decomposition of NO2-hydrocarbon complexes over both the Br?nsted and copper-containing sites of the zeolite  相似文献   

2.
The reduction of the concentration of strongly acidic Brönsted sites of HZSM-5 by neutralization due to added boehmite is reported. Strongly acidic Brönsted sites are neutralized by [AlO·(Al2O3)x]+ isopolyoxocations proceeding during the calcination treatment.  相似文献   

3.
The effect of steam-treatment to HZSM-5 zeolite and Mo/HZSM-5 with a steaming time range of 0.5-1 h on the catalytic performance of methane dehydro-aromatization (MDA) over Mo/HZSM-5 catalyst prepared with impregnation has been studied in detail in combination with the characterization of 1H MAS NMR technique. Both the deactivation rate constant (kd) and the Brnsted acid sites per unit cell were calculated to quantitatively evaluate the stability of Mo/HZSM-5 catalysts treated with steam at 813 K before and after impregnation of molybdenum species, and the corresponding variation of their Brnsted acid sites. The results reveal that a V-shape relationship between kd and the number of B1 acid sites per unit cell is presented on Mo/HZSM-5 catalyst under the tested steam-treatment and reaction conditions.  相似文献   

4.
It has been observed by FT-IR spectroscopy that both kinds of Brönsted acid protons present in HZSM-5 zeolite may be involved in the adsorption of methane at low temperature (173 K) and exchange with CH4 or CD4 at high temperature (>500°C). The sites which can adsorb methane at low temperature are the active sites for methane conversion at high temperature. Over HZSM-5 zeolite, the activation of methane is suggested to occur via a heterolytic cleavage of C–H bond with the assistance of protons.  相似文献   

5.
Catalytic properties of HZSM-5s with three different Na+ ion-exchange levels and SiO2 /Al2O3 ratios used in tert-butylation of DHB (1,2-dihydroxybenzene) are interpreted through pyridine adsorbed FT-IR and XPS study. The DHB conversion decreases as increment of degree of Na+ ion-exchange level and of Si content in HZSM-5. Catalytic properties with respect to Na amount in ZSM-5 are more sensitive than those of HZSM-5s with different SiO2/Al2O3 ratios. But selectivity for 4-TBC (4-t-butylcatechol) is not changed significantly. Acidic properties, i.e. acid strength and acid density are characterized by pyridine adsorbed FT-IR and XPS study. Based on FTIR and XPS analyses, DHB conversion and selectivities for DTBC (3,5-di-t-butylcatechol) and 3-TBC (3-t-butylcatechol) depend on type and strength of acid sites, with the result that strong Brønsted acid rather than weak Brønsted or Lewis acid sites are more closely related to the conversion. Furthermore, t-butyl alcohol is selectively adsorbed on the Brønsted acid site of FT-IR band at 3612 cm-1, which signifies that the Brønsted acid site is the active site. The mechanism for t-butylation of DHB is suggested based on the FT-IR results of adsorption/desorption of reactants.  相似文献   

6.
The formation of 2-aminoacetamide from ammonia and glycine and N-glycylglycine from two glycine molecules with and without Mg2+, Cu2+, and Zn2+ cations as catalysts have been studied as model reactions for peptide bond formation using the B3LYP functional with 6–311+G(d,p) and 6–31G(d) basis sets. The B3LYP method was also used to characterize the nine gas–phase complexes of neutral glycine, its amide (2-aminoacetamide), and N-glycylglycine with Lewis acids Mg2+, Cu2+, and Zn2+, respectively. Further, the gas-phase hydration of metal-coordinated complexes of glycine, 2-aminoacetamide, and N-glycylglycine was also investigated. Finally, the effect of water on the structure and reactivity of the metal coordinated complexes was determined. Enthalpies and Gibbs energies for the stationary points of each reaction have been calculated to determine the thermodynamics of the reactions investigated. A substantial decrease in reaction enthalpies and Gibbs energies was found for glycine–ammonia and glycine–glycine reactions coordinated by Mg2+, Cu2+, and Zn2+ ions compared to those of the uncoordinated 2-aminoacetamide bond formation. The formation of a dipeptide is a more exothermic process than the creation of simple 2-aminoacetamide from glycine. The energetic effect of the transition metal ions Cu2+ and Zn2+ is of similar strength and more pronounced than that of the Mg2+ cation. The basicity order of the amides investigated shows the order: NH2CH2CO2H < NH2CH2CONH2 < NH2CH2CONHCH2CO2H. Interaction enthalpies and Gibbs energies of metal ion–amide complexes increase as Mg2+2+2+. In both reactant (glycine) and reaction products (2-aminoacetamide, N-glycylglycine) dihydration caused considerable reduction (about 200–500 kJ-mol–1) of the strength of the bifurcated metal–amide bonds. Solvent effects also reduce the reaction enthalpy and Gibbs energy of reactions under study.  相似文献   

7.
The mordenite samples loaded with divalent nitrates of Mn, Co, Ni, Cu, Zn and Cd were investigated using FTIR and scanning electron microscopy (SEM) methods. It was found from FTIR spectra that in 3000-4000 cm−1 region of mordenite samples with similar water concentration ions, Mn2+, Co2+, Cu2+, and Zn2+ tend to break hydrogen bonds formed between water molecules and zeolite framework, whereas Ni2+ and Cd2+ accommodate to hydrogen bonds. From SEM results it was concluded, that ions Mn2+, Co2+, Zn2+ form innersphere complexes with oxygens from Brönsted acid sites, whereas Ni2+ and Cd2+ associate with Brönsted acid sites without exchange of protons.  相似文献   

8.
Effects of concentrations of ammonia (0.3–5.8 M) and supporting electrolytes (NaF, NaClO4; 0.1–0.5 M) on the kinetics of electroreduction of ammonia complexes of cobalt(II) at a dropping mercury electrode are studied. Most experiments are performed with low concentrations of cobalt(II) complexes (1 × 10–5 to 2 × 10–5 M) in the absence of a polarographic maximum. The dependence of the half-wave potential of the reversible cathodic wave pertaining to the reduction of ammonia complexes of cobalt(II) on the concentration of ammonia molecules is obtained. It is found from the dependence that, at ammonia concentrations of 0.5–2.6 M, the slow electrochemical stage involves predominantly complexes Co(NH3)2 2+. At higher ammonia concentrations, the stage involves complexes Co(NH3) k 2+ (k > 2), which form in preceding chemical stages from complexes Co(NH3) i 2+ (i = 3–6) that are predominant in solution. Values of the diffusion coefficients for complexes Co(NH3) i 2+, apparent transfer coefficients, and rate constant of the process of electroreduction of ammonia complexes of cobalt(II) are determined. The reasons for the complicating effect the insoluble products of reduction of cobalt(II) complexes have on the shape of polarographic waves are discussed.  相似文献   

9.
We have studied 18 reactions, including four identity reactions, involving transfer of a dimethylcarbamoyl group with N-acylpyridinium bonds to pyridine and its 4-substituted derivatives in acetonitrile solutions at 298 K. The rate constants k ij varied within the range 0.4 to 10–6 L/mol·s; the equilibrium constants K ij varied from 107 to 10–5. The rate and equilibrium for exchange of carbamoyl groups are described satisfactorily by the Brönsted equation. We have shown that all the reactions occur according to a forced concerted S N2 mechanism. The structure of the transition state and its position on the reaction coordinate for identity transfer are considered using a More O'Ferrall-Jencks diagram.  相似文献   

10.
Summary Uranium (VI) forms anionic complexes with aliphatic dicarboxylic acids namely oxalic, succinic and adipic acids. These complexes are stable in solutions of pH up to 5 and retained quantitatively on Lewatit MN (Cl). The break-through capacities of the ion exchange column used are given for the different uranyl complexes. Li, Cu2+, Ni, Co, Cr3+, Fe3+, Al, Mn2+, Th, Zr, Ce4+ andPO4 3– which may interfere with uranium are tested under the prescribed conditions in order not to affect uranium determination.Thorium and zirconium form anionic oxalate complexes in 1% oxalic acid-ammonium hydroxide solution of pH range 2.5–5 and are retained quantitatively on Lewatit MN (Cl). 0.4 N hydrochloric acid solution is used for eluting uranium or zirconium while thorium is eluted with 4–5 N hydrochloric acid solution.
Zusammenfassung Uran(VI) bildet mit aliphatischen Dicarbonsäuren (Oxal-, Bernstein, Adipinsäure) anionische Komplexe, die in Lösung bis zu pH 5 beständig sind und von Lewatit MN (Cl) quantitativ zurückgehalten werden. Die Durchbruchskapazitäten der verwendeten Austauschersäule für die einzelnen Urankomplexe werden angegeben. Störungen durch Li, Cu2+, Ni, Co, Cr3+, Fe3+, Al, Mn2+, Th, Zr, Ce4+ und PO4 3– werden untersucht. Thorium and Zirkonium bilden anionische Oxalatkomplexe in 1%iger Oxalsäurelösung, die mit Ammoniak auf pH 2,5 bzw. 5 eingestellt ist; diese Komplexe werden von Lewatit MN (Cl) quantitativ zurückgehalten. Uran oder Zirkonium werden mit 0,4 n Salzsäure, Thorium mit 4–5 n Salzsäure eluiert.
  相似文献   

11.
Zusammenfassung Eine empfindliche und selektive spektralphotometrische Kobaltbestimmung (<0,9 g ml–1) ist mit 2-(5-Brom-2-pyridylazo)-5-diethylaminophenol (5-Brom-PADAP) auf Grund des kinetisch stabilen Co(III)L 2 -Chelates möglich, das nach seiner Bildung bei pH 7 (Ammoniumacetat) in Gegenwart von 0,1% Triton X-100, 5,10–3 M Ammoniumperoxodisulfat und 10% (v/v) Dimethylformamid (auch nach Zugabe von 1,25 M H2SO4, 1,80 M HNO3 oder auch 0,1–0,01 M EDTA) stabil ist. Eine Reihe von 5-Brom-PADAP-Chelaten störender Ionen werden dabei zersetzt. V(V), Hg2+, Ni2+, Cu2+, Pd2+ können stören. Die Kobaltbestimmung in Cyanocobalamin und auch Trinkwasser ist auf diesem Wege möglich. Die Fehler lagen im Bereich von — 3 bis — 7 % (1,8–9 g Co/l).Herrn Prof. Dr. F. Umland, Westfälische Universität, Münster/Westfalen, sind wir für das 5-Brom-PADAP mit herzlichem Dank verpflichtet.  相似文献   

12.
采用两步水热晶化法,通过在合成体系中加入硼酸、氟化铵、氟硼酸铵,合成出了硼和氟改性的ZSM-5分子筛。利用X射线衍射、氮气吸附-脱附、29Si固体核磁共振波谱、傅里叶变换红外光谱、扫描电子显微镜以及NH3程序升温脱附等测试手段对样品进行了表征。结果表明:硼和氟掺杂条件下可以合成具有较高结晶度的ZSM-5分子筛,杂原子掺杂提高了分子筛的硅铝比;硼和氟掺杂可以显著降低ZSM-5分子筛的Lewis酸量,但提高了Brønsted酸量;硼和氟共同作用可以降低ZSM-5分子筛的颗粒尺寸。甲醇制丙烯评价结果显示:较低的Lewis酸量和适宜的Brønsted酸性有利于提高丙烯选择性和催化剂寿命;NH4BF4改性的ZSM-5分子筛(Z5-BF2)表现出较高的丙烯选择性和较长的催化剂寿命。  相似文献   

13.
Summary Acid catalysed dissociation of the copper(II) and nickel(II) complexes (ML2+ of the quadridentate macrocyclic ligand 1, 5, 9, 13-tetraaza-2, 4, 4, 10, 12, 12-hexamethyl-cyclohexadecane-1, 9-diene (L) has been studied spectrophotometrically. Both complexes dissociate quite slowly with the observed pseudo-first order rate constants (kobs) showing acid dependence; for the nickel(II) complex (kobs)=kO+kH[H+], the ko path is however absent with the copper(II) complex. At 60°C (I=0.1M) the kH values areca 10–4 M–1 s–1 for both complexes; k H Cu /k H Ni =ca. 3.9, comparable to some other square-planar complexes of these metal ions. The rate difference is primarily due to H values [copper(II) complex, 29.4±0.5 kJ mol–1; nickel(II) complex, 35.6±1.5 kJ mol–1] with highly negative S values [for copper(II), –215.5 ±6.1 JK–1 mol–1 and for nickel(II), –208.1 ±5.6 JK–1 mol–1] which are much higher than the entropy of solvation of Ni2+ (ca. –160 JK–1 mol–1) and Cu2+ (ca. –99 JK–1 mol–1) ions; significant solvation of the released metal ions and the ligand is indicated.  相似文献   

14.
The local structures of various Brønsted and Lewis acid sites in H-Beta zeolite were resolved with the combined 31P MAS NMR, 31P–27Al TRAPDOR NMR experiments and theoretical calculations at different levels. In addition, the interacting mechanisms of these acid sites with probe molecules such as trimethylphosphine (TMP) and trimethylphosphine oxide (TMPO) were clarified, which greatly aids the understanding of acid catalysis. Owing to the narrow chemical shift range and close Brønsted acid strengths, only an average resonance at −4.5 ppm was observed in TMP adsorbed H-Beta zeolite, consistent with the calculated data of acidities (substitution energies and proton affinities), geometries, adsorption energies as well as 31P chemical shifts. However, two types of Brønsted acids were distinguished by TMPO, and the HF/DZVP2 (MP2/DZVP2) chemical shifts were calculated at 68.1 (69.5) and 69.7–72.1 (71.7–74.9) ppm, respectively. Two types of Lewis acids were identified at −32.0 and −47.0 ppm with the latter exhibiting strong 31P–27Al TRAPDOR effects. With theoretical calculations, these two peaks were attributed to the extra-lattice oxo-AlOH2+ species and the three-fold coordinated lattice-Al, extra-framework Al(OH)3, oxo-AlO+ species, respectively. The peak at −60.0 ppm was conventionally assigned to the TMP physisorption, but the calculations indicated that the EFAL monovalent Al(OH)2+ species coordinating with two lattice-O atoms near the framework Al atom can contribute to it as well.  相似文献   

15.
应用原位漫反射红外-质谱联用、程序升温和暂态响应技术研究了CuO/Al2O3催化剂表面酸性及其反应性能. 实验结果表明, CuO/Al2O3催化剂表面呈Lewis酸性, 硫化不仅可增强CuO/Al2O3催化剂的Lewis酸性, 而且可产生新的Brønsted酸性位; 吸附于Lewis酸性位的NH3具有选择性催化还原(SCR)活性. 而在硫化样Cu8(400S)中Lewis和Brønsted酸性位同时存在的情况下, 吸附于Lewis和Brønsted酸性位的氨均具有SCR活性, 且后者较前者弱; CuO/Al2O3催化剂上的SCR反应遵循Eley-Rideal机理, 即SCR反应发生于吸附态NH3与气相NO之间.  相似文献   

16.
Nitrogen adsorption on SiO2, -Al2O3, TiO2, and sulphated zircona (SO 2- 4 /ZrO2) is studied by Fourier transform IR spectroscopy. Integrated absorption coefficients for the bands due to the N—N vibrations in nitrogen complexes with Brønsted and Lewis acid sites are determined. A general correlation between integrated absorption coefficients and the positions of N—N bands of nitrogen interacting with the above sites in zeolites and oxides is discussed. The orientation of a nitrogen molecule relative to Brønsted and Lewis acid sites is calculated ab initio using a 6-31G** basis set.Translated from Kinetika i Kataliz, Vol. 46, No. 1, 2005, pp. 115–121.Original Russian Text Copyright © 2005 by Malyshev, Paukshtis, Malysheva.  相似文献   

17.
Synthesis of an intercalated compound of montmorillonite and 6-polyamide   总被引:7,自引:0,他引:7  
Natural montmorillonite, fractionated from bentonite produced in Yamagata, Japan, was ion-exchanged for NH 3 + –(CH2)11–COOH, NH 3 + –(CH2)5–COOH, Al3+, Cu2+, Mg2+, Co2+, Li+, K+ and H+. The mixtures of the ion-exchanged montmorillonite and -caprolactam were heated at 263°C in glass ampoules for various periods. The intercalated compounds before and after the heating were examined by X-ray powder diffraction, DSC and GPC. Although -caprolactam was not polymerized without montmorillonite, it was polymerized at 263°C in the presence of montmorillonite. The polymerization rate varied with the interlayer cations in the order of NH 3 + –(CH2)11–COOH>Al3+>NH 3 + –(CH2)5–COOH>H+>Cu2+>Mg2+>Co2+>Li+>K+. After heating at 263°C for 5 h, the mean number-average molecular weight was about 1.5×104. Although the interlayer distance of NH 3 + –(CH2)11–COOH type montmorillonite/-caprolactam compound increased from 2.85 nm to 4.90 nm by heating at temperatures above the melting point of -caprolactam, those of other compounds were not changed. After heating at 263°C, an intercalated compound of montmorillonite and 6-polyamide, whose interlayer distance was more than 10 nm, was obtained. It is concluded that montmorillonite acts as a Brönsted acid and initiates the open ring polymerization of -caprolactam and that the driving force of swelling is the polymerization energy.Presented at the Fourth International Symposium on Inclusion Phenomena and the Third International Symposium on Cyclodextrins, Lancaster, U.K., 20–25 July 1986.  相似文献   

18.
The applicability of molecular nitrogen as a probe for the Brønsted and Lewis acid sites of HNaY and HZSM-5 zeolites was studied by Fourier transform IR spectroscopy. The integrated absorption coefficients of bands due to N—N vibrations in complexes with Brønsted and Lewis acid sites were determined. The correlation between the integrated absorption coefficients and the positions of bands due to N—N vibrations in nitrogen interacting with the acid sites of test samples is discussed. We propose using the low-temperature adsorption of nitrogen for express determination of the concentrations of strong Lewis and Brønsted acid sites in zeolites.Translated from Kinetika i Kataliz, Vol. 46, No. 1, 2005, pp. 108–114.Original Russian Text Copyright © 2005 by Malyshev, Paukshtis, Malysheva, Toktarev, Vostrikova.  相似文献   

19.
Conclusions The dissociation constnats of cyclic oxygen-containing phosphorus acids in absolute alcohol were measured by potentiometric titration in charge transfer circuits.For acyclic phosphorus acids ABPOOH, there is a Brönsted dependence in pK(EtOH)-pk(H2O) coordinates, which does not apply to 1,3-alkylenephosphoric acids, while in the pK(MeNO2)pK(EtOH) coordinates there is a uniform Brönsted dependence for all the acids.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 6, pp. 1244–1248, June, 1987.  相似文献   

20.
New picoline adducts with carbamic acid [(furan-2-yl)methylene]hydrazide–CuII (CFMH) (1); thiocarbamic acid [(furan-2-yl)methylene]hydrazide–CuII (TFMH) (2); carbamic acid [(furan-2-yl)ethylidene]hydrazide–CuII (CFEH) (3), thiocarbamic acid [(furan-2-yl)ethylidene]hydrazide–CuII (TFEH) (4); carbamic acid [(thiophene-2-yl) methylene]hydrazide–CuII (CTMH) (5), thiocarbamic acid [(thiophene-2-yl)methylene]hydrazide–CuII (TTMH) (6), carbamic acid [(thiophene-2-yl)ethylidene]hydrazide–CuII (CTEH) (7), thiocarbamic acid [(thiophene-2-yl)ethylidene]hydrazide–CuII (TTEH) (8) have been prepared and characterized by analytical, i.r., electronic, e.s.r. and c.v. spectral data. The electronic spectra suggest distorted octahedral geometry for all the picoline adducts. E.s.r. g values lie between 2.251–2.286 at l.n.t. All the adducts undergo a quasi-reversible one-electron reduction in the range +0.47 to +0.51 V versus s.c.e., attributable to the CuIII/CuII redox couple. The electron transfer is much faster in the semicarbazone complexes than in the thiosemicarbazone complexes. All adducts showed increased nuclease activity in the presence of oxidant; the nuclease activity is compared with that of the parent copper(II) complexes.  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