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 171 毫秒
1.
Ab initio MO calculations have been performed for neutral and cationic C2H2F2 structures. Olefinic and carbene structures are investigated for the neutral isomers, while olefinic, carbene, and fluoronium-type cations are found. Stability orders and rotational barriers are discussed in terms of orbital and Coulomb interaction. Contrary to previous studies, the higher stability of the geminal isomers is interpreted to be caused by Coulomb attraction.  相似文献   

2.
IR and Raman spectra of [Mo2O2S2(S2)2]2- were reported. The resonance Raman spectra and the depolarization ratios in CH3CN solution were measured. By using the data of crystal structure, the simplified normal coordinate calculation of the stretching vibrations for anion [Mo2O2S2(S2)2]2- was performed. The results obtained are useful to assign the vibrational bands of some Mo-Fe-S clusters.  相似文献   

3.
Summary The ability of [MoS4]2–, anions to be used as ligands for transition metal ions has been widely demonstrated, especially with Fe2+. The present study has been restricted to linear complexes such as (NEt4)2 [Cl2FeS2MoS2] and (NEt4)2[Cl2FeS2MoS2FeCl2]. Their electrochemical properties are described: upon electrochemical reduction, these compounds yield MoS2, as a black precipitate, and an iron complex in solution, assumed to be [SFeCl2]2–. The electrochemical reduction goes through two electron transfers, coupled with the breakdown of the molecular skeleton: a DISPl and an ECE mechanism. Depending on the solvent, the following equilibrium may be observed: [Cl4Fe2MoS4]2–[Cl2FeMoS4]2–+FeCl2. The equilibrium constant, KD, was evaluated by differential pulse polarography. KD is tightly related to the donor number of the solvent.  相似文献   

4.
Hf2Ni2In, Hf2Ni2Sn, Hf2Cu2In, and Hf2Pd2In were synthesized by reacting the elements in an arc-melting furnace under argon and subsequent annealing at 970 K. They crystallize with an ordered Zr3Al2 type structure, space group P42/mnm which was refined from single crystal X-ray data for Hf2Ni2In (a = 713.9(1) pm, c = 660.4(2) pm, wR2 = 0.0665, 513 F2 values) and Hf2Ni2Sn (a = 703.1(1) pm, c = 676.1(2) pm, wR2 = 0.0423, 507 F2 values) with 18 parameters for each refinement. The lattice constants for Hf2Cu2In and Hf2Pd2In are a = 715.5(1) pm, c = 677.0(1) pm and a = 742.6(1) pm, c = 679.4(2) pm, respectively. The structures may be considered as an intergrowth of distorted CsCl- and AlB2-like slabs. Magnetic susceptibility measurements indicate Pauli paramagnetism for Hf2Ni2In and Hf2Ni2Sn, which is consistent with the metallic conductivity observed for Hf2Ni2In. 119Sn Mössbauer spectroscopy of Hf2Ni2Sn shows one signal with an isomer shift of δ = 1.59(1) mm/s subjected to quadrupole splitting of δEq = 0.81(1) mm/s.  相似文献   

5.
The crystal structure of [Nb2(S1.72Se2.28)(S2CNEt2)4], which is a solid solution based on [Nb22, η2-X2)2(S2CNEt2)4], where X2=S2, SSe, and Se2, is determined. The compound was obtained by the reaction of NaS2CNEt2 in CH3CN with the product of the reaction of the NbSe2Cl2-KNCS fusion cake with an aqueous solution of Bu4NBr (yield 30%). The crystals are monoclinic, a=21.319(11), b=7.008(1), c=16.673(8) Å, β=133.99(2)°, Vcell=1792(1) Å3, space group C2/m, Z=2, dcalc=1.879 g/cm3 for C20H40N4Nb2S9.72Se2.28, Syntex P21, λCukα, Nmeas/corr=2423/1191, R(F)=0.0569 and wR(F2)=0.1282 for 889 Fhkl>4σ(F). The compound is isostructural to the thio analog Nb2S4(S2CNEt2)4 studied earlier; in both cases, the molecule is disordered in crystals over two positions. The composition was refined by diffraction data. The results are in good agreement with the Raman and FMB (fast molecule bombardment) mass spectrometry data. The structure of Nb2S4(S2CNEt2)4 was refined for the second time using a new model of disorder; as a result of the refinement, a normal value of S?S bond length in the S2 ligand [2.027(5) Å] was obtained.  相似文献   

6.
Ab initio calculations have been performed on the complexes of CF2Cl2 with NO and SO2, and a set of stable configurations for CF2Cl2–NO and CF2Cl2–SO2 were found with no imaginary frequencies by the MP2 method. In addition, the binding energy and the NBO analysis were used to evaluate the relative stability of the complexes. The calculated results indicate that the weak interactions in the CF2Cl2–NO and CF2Cl2–SO2 systems involved are enhanced with the increase of the number of non-covalent bonds. Further studies predict that the CF2Cl2–SO2 system may play a more important role than the CF2Cl2–NO system in environmental problem because the former offers a stronger interaction than the latter. Furthermore, the non-covalent binding interactions of Cl···N and Cl···O for CF2Cl2–NO system, Cl···O, Cl···S and F···S for CF2Cl2–SO2 system, are the dominant forces, which seem to be very significant as a driving force influencing the arrangement of molecules, especially in CF2Cl2–SO2 system.  相似文献   

7.
[(n‐Bu)2Sn(O2PPh2)2] ( 1 ), and [Ph2Sn(O2PPh2)2] ( 2 ) have been synthesized by the reactions of R2SnCl2 (R=n‐Bu, Ph) with HO2PPh2 in Methanol. From the reaction of Ph2SnCl2 with diphenylphosphinic acid a third product [PhClSn(O2PPh2)OMe]2 ( 3 ) could be isolated. X‐ray diffraction studies show 1 to crystallize in the monoclinic space group P21/c with a = 1303.7(1) pm, b = 2286.9(2) pm, c = 1063.1(1) pm, β = 94.383(6)°, and Z = 4. 2 crystallizes triclinic in the space group , the cell parameters being a = 1293.2(2) pm, b = 1478.5(4) pm, c = 1507.2(3) pm, α = 98.86(3)°, β = 109.63(2)°, γ = 114.88(2)°, and Z = 2. Both compounds form arrays of eight‐membered rings (SnOPO)2 linked at the tin atoms to form chains of infinite length. The dimer 3 consists of a like ring, in which the tin atoms are bridged by methoxo groups. It crystallizes triclinic in space group with a = 946.4(1) pm, b = 963.7(1) pm, c = 1174.2(1) pm, α = 82.495(6)°, β = 66.451(6)°, γ = 74.922(6)°, and Z = 1 for the dimer. The Raman spectra of 2 and 3 are given and discussed.  相似文献   

8.
We have performed large-scaleab initio calculations using second order Møller-Plesset perturbation theory (MP2) on the three van der Waals dimers formed from acetylene and carbon dioxide. Intermolecular geometrical parameters are reliably computed at this level of theory. Calculations of vibrational frequencies of the van der Waals modes, currently unobtainable by experimental means, give important information about the intermolecular potential and predict significant large-amplitude motion. Zero point energy contributions are shown to be vital in assessing the relative stability of conformations which are close in energy. Our studies suggest that the barrier to interconversion tunnelling in (CO2)2 is significantly smaller than previously inferred and is approximately the same as in (C2H2)2. The reason for the rigidity of (CO2)2 is the difference in monomer centre-of-mass separation between ground state and transition state. We also show that, in addition to the previously observedC 2v form, the collinear form of C2H2-CO2 is a local minimum on its potential energy surface.  相似文献   

9.
Preparation and Crystal Structure of the Pnictide Oxides Na2Ti2As2O and Na2Ti2Sb2O Na2Ti2As2O and Na2Ti2Sb2O were synthesized in form of very easily hydrolysed metallic-grey powders by reaction of Na2O and TiAs resp. TiSb in sealed tantalum tubes under argon. The tetrahedral bodycentered crystallizing compounds from a modified anti-K2NiF4 structure type [1] (also called Eu4As2O-type [2,3]), space group I4/mmm (no. 139), with the lattice constants for Na2Ti2As2O: a = 407.0(2) pm, c = 1528.8(4) pm and for Na2Ti2Sb2O: a = 414.4(0) pm, c = 1656.1(1) pm. Magnetic measurements of powder samples of Na2Ti2Sb2O show antiferromagnetic interaction within the Ti—O-layers. Superconductivity was not found by ac-shielding method down to 4 K.  相似文献   

10.
2-Aryl- and 2-alkyl-2-oxazolines have been polymerized to poly-(N-aroyl)aziridines and poly(N-acyl)aziridines, respectively, in the presence of boron trifluoride. The polymers obtained were glassy, light yellow resins with molecular weights ranging from 3500 to 7500 (35–50 oxazoline units per chain). The polymerization rates have been determined for several of these monomers. A polymerization mechanism is proposed.  相似文献   

11.
Conclusions A physicochemical investigation of the system Rb2CO3-H2O2-H2O in the region of high hydrogen peroxide concentrations at low temperatures revealed for the first time the formation of a peroxyhydrate with a high hydrogen peroxide content, of the composition Rb2CO3 · 6H2O2.  相似文献   

12.
The title structures NaGdS2 (sodium gadolinium sulfide), NaLuS2 (sodium lutetium sulfide) and NaYS2 (sodium yttrium sulfide) were redetermined in order to improve the structural information available for the family of group 1 and thallium rare earth sulfides, which are isostructural with the rhombohedral α‐NaFeO2 structure type. In particular, the present investigation has been directed at the rhombohedral sodium rare earth sulfides. The observed dependence of the fractional coordinate z(S2−) on the identity of the rare earth element in the newly determined structures is in agreement with the known structures of the potassium and rubidium analogues. Crystals of NaGdS2 and NaLuS2 display obverse–reverse twinning.  相似文献   

13.
Two glasses based on lithium disilicate (LS2), with and without fluorapatite (FA), were synthesised in the Li2O-SiO2-CaO-P2O5-CaF2 system with P2O5: CaO: CaF2 ratios corresponding to fluorapatite. Glass-ceramics have then been prepared by thermal treatment. The mechanism and kinetics of crystallization as functions of grain size and rate of heating were investigated using thermal analysis methods. The smaller particles crystallize preferentially by surface crystallization, which is replaced by volume crystallization at larger particle sizes. Inclusion of FA in the LS2 favours crystallization through the surface mechanism. The onset limit for volume crystallization replacing the surface mechanism is at about 0.3 mm for pure LS2 glass and 0.9 mm for glass containing FA. The calculated activation energies of the glasses (299 ± 1 kJ mol-1 for pure LS2 glass and 288 ± 7 kJ mol−1 for glass containing FA according to Kissinger, or 313 ± 1 kJ mol-1 for pure LS2 glass and 303 ± 8 kJ mol-1 for glass containing FA according to Ozawa) indicate that the tendency of the glasses to crystallize is supported by the FA presence. Bioactivity of all samples has been proved in vitro by the formation of new layers of apatite-like phases after soaking in SBF.   相似文献   

14.
Rate constants for the reactions of atomic oxygen (O3P) with C2H3F, C2H3Cl, C2H3Br, 1,1-C2H2F2, and 1,2-C2H2F2 have been measured at 307°K using a discharge-flow system coupled to a mass spectrometer. The rate constants for these reactions are (in units of 1011 cm3 mole?1 s?1) 2.63 ± 0.38, 5.22 ± 0.24, 4.90 ± 0.34, 2.19 ± 0.18, and 2.70 ± 0.34, respectively. For some of these reactions, the product carbonyl halides were identified.  相似文献   

15.
Constructing nanocomposites that combine the advantages of composite materials, nanomaterials, and interfaces has been regarded as an important strategy to improve the photocatalytic activity of TiO2. In this study, 2D-2D TiO2 nanosheet/layered WS2 (TNS/WS2) heterojunctions were prepared via a hydrothermal method. The structure and morphology of the photocatalysts were systematically characterized. Layered WS2 (~4 layers) was wrapped on the surface of TiO2 nanosheets with a plate-to-plate stacked structure and connected with each other by W=O bonds. The as-prepared TNS/WS2 heterojunctions showed higher photocatalytic activity for the degradation of RhB under visible-light irradiation, than pristine TiO2 nanosheets and layered WS2. The improvement of photocatalytic activity was primarily attributed to enhanced charge separation efficiency, which originated from the perfect 2D-2D nanointerfaces and intimate interfacial contacts between TiO2 nanosheets and layered WS2. Based on experimental results, a double-transfer photocatalytic mechanism for the TNS/WS2 heterojunctions was proposed and discussed. This work provides new insights for synthesizing highly efficient and environmentally stable photocatalysts by engineering the surface heterojunctions.  相似文献   

16.
Direct synthesis of H2O2 solutions by a fuel cell method was reviewed. The fuel cell reactor of [O2, gas-diffusion cathode electrolyte solutions Nafion membrane electrolyte solutions gas-diffusion anode, H2] is very effective for formation of H2O2. The three-phase boundary (O2(g)–electrode(s)–electrolyte(l)) in the gas-diffusion cathode is essential for efficient formation of H2O2. Fast diffusion processes of O2 to the active surface and of H2O2 to the bulk electrolyte solutions are essential for H2O2 accumulation. The maxima H2O2 concentrations of 1.2 M (3.5 wt%) and 2.4 M (7.0 wt%) were accomplished by the heat-treated Mn-OEP/AC electrocatalyst with H2SO4 electrolyte and by the VGCF electrocatalyst with NaOH electrolyte, respectively, under short circuit conditions.  相似文献   

17.
The kinetics of the thermal dehydration of various kinds of BaCl2 · 2H2O and of BaCl2 · H2O are investigated using a differential scanning calorimeter. The loss of H2O proceeds in two steps: BaCl2 · 2H2O→BaCl2 · H2O→BaCl2 and is therefore revealed by two endothermic peaks. In the experiments at varying temperature both steps follow a contracting-circle law, after an initial acceleratory stage according to a (n=2) power law. In the experiments at constant temperature, after an initial acceleratory stage according to a (n=2) power law, both steps (except BaCl2 · 2H2O single-crystals which follow a contracting-circle law) follow an Avrami-Erofeev law (withn=2) in the form used by Galwey and Jacobs. The activation energies for the various steps are compared and the different kinetic behaviour is discussed.  相似文献   

18.
The structures of all compounds were determined from three dimensional single crystal X-ray diffraction data and refined by least squares. Ba2CdS3 and Ba2CdSe3 are isostructural, Pnma, a = 8.9145(6)Å, b = 4.3356(2)Å, c = 17.2439(9)Å for the former compound and a = 9.2247Å, b = 4,4823(6)Å, c = 17.8706(11)Å for the latter, z = 4, R = 0.0751 and R = 0.0462, respectively. The compounds are isostructural with the previously reported Mn analogues and with K2AgI3. Cd ions are in tetrahedral environment and the tetrahedra form infinite linear chains by corner sharing. Ba ions are in 7-fold coordination in which 6 anions form a trigonal prism and 1 anion caps one of the rectangular faces. BaCdS2, Pnma, a = 7.2781(3)Å, b = 4.1670(1)Å, c = 13.9189(6)Å, z = 4, R = 0.0685. Cd ions can be considered to have a triangular planar coordination with CdS distances of 2.47 and 2.53 Å (twice). Two additional S ions are at 2.89 and 3.22 Å to complete a triangular bipyramidal configuration. Ba is in 7-fold coordination with the anions forming a trigonal prism which is capped on one rectangular face. The compound is isostructural with BaCdO2 and is related to the structure of BaMnS2. BaCdSe2 could not be prepared. BaCu2S2 and BaCu2Se2 are isostructural, Pnma, a = 9.3081(4)Å, b = 4.0612(3)Å, c = 10.4084(5)Å for the sulfide and a = 9.5944(6)Å, b = 4.2142(4)Å, c = 10.7748(8)Å for the selenide, z = 4, R = 0.0634 and 0.0373, respectively. Ba ions are in the usual 7-fold, capped hexagonal prism, coordination. However, 9 Cu ions also can be considered to form a trigonal prism with all rectangular faces capped, around Ba since the BaCu distances range from 3.24 to 3.54 Å for the sulfide and from 3.37 to 3.67 Å for the selenide. One of the Cu ions is in a very distorted tetrahedral environment and the second one is located in a more regular tetrahedral configuration of the anions. Two independent infinite chains of tetrahedra are present. They are formed by sharing of two adjacent edges of each tetrahedron and then these chains in turn are linked by corner sharing into a three-dimensional network of tetrahedra.  相似文献   

19.
We report quantitative calculations of stereomutation tunneling in the disulfane isotopomers H2S2, D2S2, and T2S2, which are chiral in their equilibrium geometry. The quasi‐adiabatic channel, quasi‐harmonic reaction path Hamiltonian approach used here treats stereomutation including all internal degrees of freedom. The torsional motion is handled as an anharmonic reaction coordinate in detail, whereas all the remaining degrees of freedom are taken into account approximately. We predict how stereomutation is catalyzed or inhibited by excitation of the various vibrational modes. The agreement of our theoretical results with spectroscopic data from the literature on H2S2 and D2S2 is excellent. We furthermore predict the influence of parity violation on stereomutation as characterized approximately by the ratio (ΔEpv/ΔE±) of the (local or vibrationally averaged) parity violating potential ΔEpv and the tunneling splittings ΔE± in the symmetrical case. This ratio is exceedingly small for the reference molecules H2O2 and D2O2, and still very small (2⋅10−6 cm−1) for H2S2, which, thus, all exhibit essentially parity conservation in the dynamics. However, for D2S2 it is ca. 0.002, and for T2S2 it is ca. 1, which seems to be the first case where such intermediate mixing through parity violation is quantitatively predicted for spectroscopically accessible molecules. The consequences for the spectroscopic detection of molecular parity violation are discussed briefly also in relation to other molecules.  相似文献   

20.
The elimination kinetics of the title compounds have been examined over the temperature range of 270–320°C and pressure range of 19–117 torr. The reactions, carried out in seasoned vessels, with the free-radical suppressor toluene always present, are homogeneous, unimolecular, and follow a first-order rate law. The products of 2-hydroxy-2-methylbutyric acid are 2-butanone, CO, and H2O; while of 2-ethyl-2-hydroxybutyric acid are 3-pentanone, CO, and H2O. The rate coefficient is expressed by the following Arrhenius equation: for 2-hydroxy-2-methylbutyric acid, log k1(s?1 = (12.87 ± 0.19) ? (171.2 ± 2.1) kJ mol?1 (2.303 RT)?1; and for 2-ethyl 2-hydroxybutyric acid, log k1s?1) = (12.13 ± 0.34) ? (159.4 ± 3.7) kJ mol?1 (2.303 RT)?1. Augmentation of alkyl bulkiness at the 2-position of the 2-hydroxycarboxylic acids showed an increase in the rate of dehydration. The electron release of alkyl groups, rather than steric acceleration, appears to enhance the pyrolysis decomposition of these substrates. These reactions are believed to proceed through a semi-polar five-membered cyclic transition type of mechanism. © 1995 John Wiley & Sons, Inc.  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