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.
Quantum chemistry calculations were carried out, using ONIOM2 methodology, in order to investigate the thiophene interaction with gold supported on silicoaluminophospates molecular sieves (Au/SAPO-11) catalysts. Two models were studied, one containing one Au atom per site, and the other with two Au atoms per site. Thiophene adsorption was found to be η1 type. This adsorption presents a ΔH of ?13.2 and ?9.7 kcal/mol, for the models with one Au atom (Au/SAPO-11), and two Au atoms (Au2/SAPO-11), respectively. The partial hydrogenation of the thiophene–Au/SAPO-11 and thiophene–Au2/SAPO-11 complexes gives 2,5-dihydrothiophene (DHT), with a ΔH of ?23.0 and ?36.8 kcal/mol, respectively. 2-Butene production was found in both models with further hydrogenation. Likewise the direct butadiene elimination is achieved, but only with the separated Au dimer (ΔH = ?17.5 kcal/mol).  相似文献   

2.
Based on van der Waals corrected density functional theory, we show that Na atoms acting as decoration metals are not inclined to form clusters due to a large binding energy of 3.31 eV, indicating a promising good reversible hydrogen storage. Both the polarization mechanism and the orbital hybridizations contribute to the adsorption of hydrogen molecules (storage capacity of 4.4 wt%) with optimal adsorption energy of 0.25 eV/H2. Additionally, the dimerization of these isolated B36 does not remarkably affect the number of adsorbed H2 per Na atom. Our results may serve as a guide in the design of new hydrogen storage materials based on low-dimension boron clusters.  相似文献   

3.
Five new triphenyltin(IV) sulfanylcarboxylates of the general formula [(SnPh3)2L] (L = pspa, tspa, fspa, p-mpspa or cpa, where p = 3-(2-phenyl)-, t = 3-(2-thienyl)-, f = 3-(2-furyl)-, p-mp = 3-(4-methoxyphenyl)-, spa = 2-sulfanylpropenoato and cpa = 2-cyclopentilyden-2-sulfanylacetate) have been synthesized by reacting triphenyltin(IV) hydroxide with the corresponding acid in ethanol/acetone. The complexes have been characterized by elemental analysis and mass spectrometry and by vibrational and NMR (1H, 13C, 119Sn) spectroscopies. In the case of [(SnPh3)2(p-mpspa)] and [(SnPh3)2(cpa)], X-ray structural studies showed that in both compounds each Sn atom is coordinated to three phenyl C atoms and to one S or O atom of the bridge ligand L. All five complexes are active against strains of Staphylococcus aureus, but are inactive against Escherichia coli and Pseudomonas aeruginosa. From a solution of [(SnPh3)2(tspa)] in DMSO-d6 the new complex [(SnPh3)2(tspa)(DMSO)] was isolated. The single-crystal X-ray diffractometric study of this complex is also reported, showing that both Sn atoms are bridged by the tspa ligand, whereas the molecule of DMSO is coordinated to one of the tin atoms via the oxygen atom.  相似文献   

4.
Extreme ultraviolet (EUV) spectroscopy was recorded on microwave discharges of helium with 2% hydrogen. Novel emission lines were observed with energies of q·13.6 eV where q=1,2,3,4,6,7,8,9, or 11 or these lines inelastically scattered by helium atoms wherein 21.2 eV was absorbed in the excitation of He (1s2) to He (1s12p1). These lines were identified as hydrogen transitions to electronic energy levels below the ‘ground’ state corresponding to fractional quantum numbers. Significant line broadening corresponding to an average hydrogen atom temperature of 33–38 eV was observed for helium–hydrogen discharge plasmas; whereas pure hydrogen showed no excessive broadening corresponding to an average hydrogen atom temperature of ≈3 eV. Since a significant increase in H temperature was observed with helium–hydrogen discharge plasmas, and energetic hydrino lines were observed at short wavelengths in the corresponding microwave plasmas that required a very significant reaction rate due to low photon detection efficiency in this region, the power balance was measured on the helium–hydrogen microwave plasmas. With a microwave input power of 30 W, the thermal output power was measured to be at least 300 W corresponding to a reactor temperature rise from room temperature to 900 °C within 90 s, a power density of 30 MW/m3, and an energy balance of about −4×105 kJ/mol H2 compared to the enthalpy of combustion of hydrogen of −241.8 kJ/mol H2.  相似文献   

5.
Substitution of the peripheral H atoms in the corannulene molecule as a carbon nanostructure by OH, CH3, NH2 and NO2 groups on the molecular hydrogen physisorption was evaluated at MP2/6-31G(d) level of theory. Two orientations of hydrogen were used on the concave and convex sides of corannulene. It was seen that binding to the concave face is favored relative to the convex face. The average binding energy was calculated and corrected for the basis set superposition error (BSSE) using the counterpoise method. Results showed that binding energy varies depending upon the site and side of absorption. The electronic density, charge transfer and spatial prohibition of the substituted groups affects the binding energy. The increment of the electronic density because of the substitution of electron donor groups facilitates hydrogen adsorption and leads to larger binding energies than when H atoms are substituted by electron acceptor groups. Substitution of H atoms with each of the considered groups leads to decreasing of the HOMO–LUMO energy gap and so decreasing of the kinetic stability and increasing of the reactivity. The energy gap and binding energy for corannulene derivatives decreases in the order of: CH3 > OH > NH2 > NO2.  相似文献   

6.
The adsorption of VB12 onto CMK-3 was studied as a function of temperature and initial VB12 concentration. The highest VB12 adsorption capacity was determined as 353.4 mg/g at 40℃. Adsorption data were well described by the Langmuir model, although they could be modelled by the Freundlich equation. The pseudo-first-order and pseudo-second-order kinetic models were applied to test the experimental data. The pseudo-second-oider kinetic model provided the best correlation of the experimental data compared to the pseudo-nrsi-order model.  相似文献   

7.
《Polyhedron》2005,24(3):435-441
The title compounds were synthesized by the reactions of [NH4]2[MS4] (M = Mo, W), AgI and Hmimt in acetone and characterized by IR, 1H NMR and UV–Vis spectroscopy. The polymeric structure of [WS4Ag2(Hmimt)2]n was determined by X-ray crystallography. In this compound, there are two distinctly different coordination modes for the silver atoms. One Ag atom has a pseudo-tetrahedral geometry with one terminal monodentate-S Hmimt, two μ2-S bridging Hmimt and one S atom of a monodentate WS4 unit. The other is surrounded by four sulfur atoms belonging in pairs to two WS4 fragments; the coordination geometry is distorted tetrahedral. The [WS4Ag2(Hmimt)2]n polymer represents the first example of tetrathiometalate anions [MS4]2− (M = Mo, W, or V) coordinated to another metal atom in a monodentate fashion. In both crystal structures determined the Hmimt ligands are present in the thione form, with coordination taking place via the sulfur atom only.  相似文献   

8.
《Progress in Surface Science》2006,81(8-9):337-366
Recent progress on desorption and adsorption dynamics of hydrogen (deuterium) on monohydride and dihydride Si(1 0 0) surfaces is reviewed and discussed. The dynamics experiments reveal that the desorption dynamics of hydrogen is well related to the adsorption dynamics via detailed balance. Dependence of time-of-flight (TOF) distributions of desorbed molecules on H(D) coverage is noticed to be important in understanding the kinetics mechanism of the adsorption/desorption reactions of hydrogen on the Si(1 0 0) surface. The desorption dynamics varies from the situation of strongly translational heating to the other situation of less translational heating with D coverage. This trend seems to be consistent with the 2H/3H/4H interdimer mechanism. However, despites by far the richest 4H configuration at high H coverage, the 2H desorption prevails over the 4H desorption already at 0.8 ML. To reconcile this unexpected desorption kinetics, a diffusion-promoted desorption mechanism is proposed. Height of the adsorption barriers for the 2H and 3H pathways could be reduced by the H-atom diffusion along the Si dimer rows, but that for the 4H pathway could not be the case because of no capability of diffusion on the H saturated surface. The desorption dynamics of hydrogen from the (3 × 1) dihydride surface is also reviewed and compared with the case on the monohydride surface. The sticking coefficients of hydrogen molecules onto the monohydride surfaces are evaluated from the TOF curves and found to be strongly activated by the kinetic energy. Not only the degrees of freedom of the molecules but also the vibrational degrees of freedom of substrate Si atoms determine the barrier height for adsorption. The desorption dynamics of hydrogen from the monohydride and dihydride surfaces appears to be quite similar, but the dynamics of substrate Si atoms is expected to be quite dissimilar between the two desorption pathways.  相似文献   

9.
《Comptes Rendus Chimie》2015,18(4):422-429
Compound 1-benzyl-5-amino-1H-tetrazole (BAT) was synthesized and characterized by 1H NMR, FT–IR, and UV–Vis spectroscopies and elemental (CHNS) analysis. The crystal structure was further elucidated by single-crystal X-ray diffraction. Density functional theory (DFT) calculations with B3LYP and PBE1PBE functionals of the BAT were performed to provide structural and spectroscopic information and guide spectral assignments. The compound crystallizes in monoclinic primitive system space group P2(1)/c with a = 14.91 Å, b = 5.12 Å, c = 11.19 Å, V = 852 Å3, Z = 4, R1 = 0.0428 at 298 K. The structure exhibits intermolecular hydrogen bonds of the type N–H(amino)···N(tetrazole). Simultaneous hydrogen bonds between amino···tetrazole and tetrazole···amino establish a dimeric intermolecular structure, whereas another hydrogen bond between the remaining H atom of the amino group and the other N atoms of the tetrazole ring extends the structure into another dimension. The crystal structure of BAT is properly reproduced by DFT calculations only when a dimeric or tetrameric model is employed in the modeling. Comparisons between experimental and calculated spectral properties suggest that the monomeric form of BAT is dominant in aprotic, polar, hydrogen-bonding solvents, such as DMSO and DMF.  相似文献   

10.
Strontium phosphate apatites with compositions Sr5(PO4)3Zn0.15O0.3(OH)0.7, Sr5(PO4)3Ni0.2O0.4(OH)0.6, and Sr5(PO4)3Co0.2O0.5(OH)0.4 were synthesized by solid state reaction at 1400 °C in air. The samples were characterized by powder X-ray diffraction, EDX analysis, magnetic measurements and IR spectroscopy. The crystal structures were refined by the Rietveld method in the space group P63/m with lattice constants a = 9.7499(1), 9.7722(1), 9.7507(1) Å and c = 7.3066(1), 7.2962(1), 7.2988(1) Å, respectively. The 3d-metal atoms were found randomly distributed in the hexagonal channels formally substituting hydrogen in the initial hydroxyapatite. Zn and Ni atoms were twofold coordinated by oxygen atoms such that the linear O–M–O groups formed in the channel separated by the OH groups. Co atom was shifted from the channel center giving the O–Co–O fragment distorted from a linear geometry probably due to the additional coordination by the oxygen atoms of the phosphate groups.  相似文献   

11.
Hydrogen adsorption isotherms, evaluated by combination of cyclic voltammetry and chronoamperometry, are reported on Pt(1 1 1) and Pt(1 0 0) surfaces in 0.1 M HClO4. We found that at E > 0.05 V Pt(1 1 1) and Pt(1 0 0) are only partially covered by the adsorbed hydrogen (Had). On both surfaces, a full monolayer of the adsorbed hydrogen is completed at −0.1 V, i.e. the adsorption of atomic hydrogen is observed in the hydrogen evolution potential region. We also found, that the activity of the hydrogen oxidation reaction is mirrored by the shape of the hydrogen adsorption isotherms, implying that Had is in fact a spectator in the HOR.  相似文献   

12.
The electrochemical reduction of CO2 is strongly influenced by both the applied potential and the surface adsorption status of the catalyst. In this work a gas diffusion electrode (GDE) coated with Pd nanoparticles/carbon black (Pd/XC72) was used to study the electrochemical reduction of CO2. Cyclic voltammetric (CV) analysis of Pd/XC72 between 1.5 V and − 0.6 V (vs. RHE) shows the formation of intermediates and the blocking of hydrogen absorption on the Pd nanoparticles (NPs) under a CO2 atmosphere. The relationships between the Faradaic efficiency/current density and the applied potential reveal that the onset potential of CO formation is around − 0.4 V. Moreover, the presence of adsorbed CO was confirmed through CV analysis of Pd/XC72 under CO2 and CO/He atmospheres. This demonstrates that H atoms and CO intermediates co-adsorb on the surface of the Pd NPs at an applied potential of around − 0.4 V. When the applied potential is more negative than − 0.6 V, adsorption of CO intermediates on the surface of the Pd NPs becomes dominant.  相似文献   

13.
Dimethylsulfoxide (DMSO)–Br complexes were generated by pulse radiolysis of DMSO/bromomethane mixtures and the formation mechanism and spectral characteristics of the formed complexes were investigated in detail. The rate constant for the reaction of bromine atoms with DMSO and the extinction coefficient of the complex were obtained to be 4.6×109 M−1 s−1 and 6300 M−1 cm−1 at the absorption maximum of 430 nm. Rate constants for the reaction of bromine atoms with a series of alcohols were determined in CBrCl3 solutions applying a competitive kinetic method using the DMSO–Br complex as the reference system. The obtained rate constants were ∼108 M−1 s−1, one or two orders larger than those reported for highly polar solvents. Rate constants of DMSO–Br complexes with alcohols were determined to be ∼ 107 M−1 s−1. A comparison of the reactivities of Br atoms and DMSO–Br complexes with those of chlorine atoms and chlorine atom complexes which are ascribed to hydrogen abstracting reactants strongly indicates that hydrogen abstraction from alcohols is not the rate determining step in the case of Br atoms and DMSO–Br complexes.  相似文献   

14.
《Polyhedron》2005,24(16-17):2491-2496
The bis(dithiolene) oxovanadium complex, namely (n-Bu4N)2[(dcbdt)2VO] (dcbdt = 4,5-dicyanobenzene-1,2-dithiolato), was unprecedentedly obtained from the reaction of Na2dcbdt and vanadium trichloride. An X-ray structure analysis indicated that [(dcbdt)2VO] moieties are surrounded by n-Bu4N cations and there was no direct interaction among these moieties. Although there was no direct interaction among them, several S⋯H, C⋯H and N⋯H van der Waals contacts between n-Bu4N cations and [(dcbdt)2VO] moieties were observed. The ESR and SQUID measurement showed that the vanadium atom is in the state of V(IV) (S = 1/2) and these magnetic moments interact with each other very weakly antiferromagnetically (θ = -0.371 K).  相似文献   

15.
New bimetallic compound [CuxNi1?x(bpy)2(ox)]·4H2O (x = 0.05, ox = oxalato, bpy = 2,2′-bipyridine) was synthesized and chemically characterized. Its crystal structure is molecular. The octahedron around the metal central atom is deformed due to coordination by one bidentate oxalate anion and two bpy ligands. There are four uncoordinated water molecules in the asymmetric unit. The metal site is occupied by both Cu(II) and Ni(II) atoms in the 5:95 ratio. The complex molecules interact with water molecules through hydrogen bonds and, moreover, π–π interactions between aromatic rings lead to a 1D arrangement of molecules. The susceptibility data measured down to 2 K were analyzed using strong-coupling theory and the best agreement with the experimental data were found for g = 2.1, D/k = 5.6, E/k = 0.35, J/k = 0.2. The dehydration starts at 30 °C. As a final product of its thermal decomposition a solid solution of CuxNi1?xO was detected by X-ray powder diffraction.  相似文献   

16.
《Vibrational Spectroscopy》2007,43(2):405-414
The catena-poly[{aqua(η2-indole-3-propionato-O,O′)zinc}-η2-:-μ-indole-3-propionato-O,O′:-O], [Zn(I3PA)2(H2O)]n was prepared and characterized by infrared spectroscopy and X-ray structure determination. The crystals are monoclinic, space group Pc, with a = 21.380(2), b = 5.9076(7), c = 8.1215(9) Å, V = 1020.2(2) Å3 and Z = 2. The central zinc atom shows the coordination distorted from ideal octahedral. Each zinc centre is coordinated by two oxygen atoms of the bidentate chelating indole-3-propionato (I3PA), two oxygen atoms tridentate chelating-bridging I3PA, water molecule and one oxygen atom tridentate chelating-bridging I3PA from an adjacent [Zn(I3PA)2(H2O)] unit. The infrared spectrum of [Zn(I3PA)2(H2O)]n in the solid state is supported by X-ray analysis. The theoretical wavenumbers and infrared intensities have been calculated by the density functional methods (B3LYP and mPW1PW) with the 6-311++G(d,p)/LanL2DZ basis sets. The theoretical wavenumbers, infrared intensities show a good agreement with experimental data. Detailed band assignment has been made on the basis of the calculated potential energy distribution (PED).  相似文献   

17.
Crystal of a new neodymium oxyborate fluoride Nd6Li(BO3)3O4F2 was grown by the flux method. Its structure, determined by single crystal x-ray diffraction, belongs to the space group C2/c with cell parameters of a = 12.0629(2) Å, b = 6.94650(10) Å, c = 16.0528(3) Å, β = 104.5360(10)°. In the structure, Nd atoms coordinate to oxygen or fluorine atoms to yeild 7 or 8 coordinated Nd(O,F)n polyhedra. Those polyhedra are edge-shared to form a double layer of (Nd12O23F4)14? fluorite blocks. The blocks are linked by oxygen atoms of planar BO3 groups in the c direction into a 3-dimensional network. Another novel element in the structure is that Li coordinates to 6 oxygen atoms from three BO3 groups forming a propeller like arrangement, and theoretical calculation shows that such arrangement should give 3/4 that of BO3 contribution to second harmonic effect. The crystal shows deep violet color with typical Nd3+ optical absorption and a UV transmission cut-off of 260 nm.  相似文献   

18.
《Vibrational Spectroscopy》2007,45(2):266-272
The solubility in the three-component system Rb2SO4–BeSO4–H2O at 25 °C was studied by the method of isothermal decrease of supersaturation. A new compound, Rb2Be(SO4)2·2H2O, is formed in a wide concentration range (from solutions containing 27.49 mass% beryllium sulfate and 20.16 mass% rubidium sulfate up to solutions containing 15.08 mass% beryllium sulfate and 39.07 mass% rubidium sulfate).Rb2Be(SO4)2·2H2O crystallizes in the monoclinic space group P21/c (a = 11.371(2) Å, b = 11.858(2) Å, c = 7.431(1) Å, β = 96.33(1), V = 996.0 Å3, Z = 4, R1 = 0.039 for 2672Fo > 4σ(Fo) and 153 variables). The crystal structure is characterized by three-membered chain fragments, composed of a central BeO2(H2O)2 polyhedron sharing corners with two SO4 tetrahedra. These bent [Be(SO4)2(H2O)2]2− units are linked by rubidium ions and hydrogen bonds to double layers and further to a three-dimensional framework structure. Rb2Be(SO4)2·2H2O is isotypic to the respective potassium sulfate and selenate compounds.The strengths of the hydrogen bonds in the title compound as deduced from the infrared wavenumbers of the uncoupled OD stretches of matrix-isolated HDO molecules (isotopically dilute sample) are discussed in terms of the Ow⋯O hydrogen bond distances, the different hydrogen bond acceptor capabilities of the sulfate oxygen atoms and the strong BeOH2 interactions (synergetic effect). The intramolecular OH bond lengths are derived from the νOD versus rOH correlation curve [H.D. Lutz, C. Jung, J. Mol. Struct. 404 (1997) 63].  相似文献   

19.
《Chemical physics》2005,309(1):33-39
The interaction of propyne with the Pt(1 1 1) and Pd(1 1 1) surfaces has been studied by means of the generalised gradient approach of density functional theory using periodic slab models. For both surfaces, the most stable adsorption mode of propyne is di-σ/π mode where the hydrocarbon is σ-bonded to two metal atoms with some additional π bonding to a third adjacent surface atom. The adsorption geometry is a highly distorted propyne with the C1 and C2 in a nearly sp2 hybridisation. Two equivalent surface structures have been found on Pt and Pd. These correspond to the adsorption on the fcc or hcp hollow sites. The adsorption energies on Pt(1 1 1) and Pd(1 1 1) are predicted to be ∼−197 and −161 kJ mol−1, respectively. The electronic factors that control the chemisorption have been analysed by means of the projected density of states.  相似文献   

20.
Layered material of zinc hydroxychlorides (Zn5(OH)8Cl2·nH2O: ZHC), which is one of the basic zinc salts (BZS), was synthesized from ZnO nano-particles aged with aqueous ZnCl2 solutions at different temperatures ranging from 6 to 140 °C for 48 h. X-ray diffraction (XRD) results indicated that the diffraction peaks of ZnO completely disappeared by aging at 6 °C and the ZHC peaks were developed. By increasing the aging temperature, crystallinity of the layered structure was improved. At 6 °C, the ZHC particles were thin hexagonal plate particles with sizes ranging from 1 to 3 μm. The particle size of ZHC was independent of aging temperature. The atomic Cl/Zn ratios of all the ZHC materials were almost 0.2 less than 0.4 of the theoretical ratio, indicating that the synthetic ZHC is Cl-deficient. It seemed that half of Cl atoms in the layer were replaced with HCO3 and/or OH. The specific surface areas of ZHC estimated from N2 adsorption isotherms were ca. 10 m2 g−1 and were independent of the aging temperature. However, the H2O monolayer adsorption capacity per unit surface area (nw) for all the samples was higher than that of ZnO particles, revealing the high affinity of ZHC to H2O molecules. The nw values were increased by reducing the crystallinity of ZHC. This enhancement of H2O adsorption selectivity was thought to be related with less-crystallized parts of the 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