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 24 毫秒
1.
The reaction of (CH3)2AsJ and AgN3 yields (CH3)2AsN3; a colourless liquid (b. p. 136°C) which dissolves as a monomeric in benzene. (CH3)2BiN3 is precipitated in form of colourless needles (dec. temp. 150°C) from an etherical solution of Bi(CH3)3 and HN3. According to its vibrational and mass spectra the molecules are not associated although the (CH3)2BiN3 is not soluble; dipole association of this polar molecules is assumed for the crystal structure. (CH3)2TlN3 can be obtained from TI(CH3)3 and ClN3 as well as from (CH3)2TlOH and HN3 in form of colourless needles and leaves (dec. temp. 245°C). According to its vibrational spectra it has an ionic structure, (CH3? Tl? CH3)+N?3.  相似文献   

2.
New Organometallic Indium Nitrogen Compounds. Synthesis and Crystal Structures of [{Cp(CO)3Mo}2InN(SiMe3)2] and [{Cp(CO)3Mo}In{N(SiMe3)2}2] The reaction of [{Cp(CO)3Mo}2InCl] with LiN · (SiMe3)2 leads to the formation of [{Cp(CO)3Mo}2InN · (SiMe3)2] ( 1 ). 1 is monomeric and it contains an indium atom which is coordinated in a trigonal planar manner by two {Cp(CO)3Mo} fragments and a N(SiMe3)2 group. The corresponding bis-amide [{Cp(CO)3Mo}In{N(SiMe3)2}2] ( 2 ) is prepared by the reaction of [{Cp(CO)3Mo}InCl2] with two equivalents of LiN(SiMe3)2. In analogy to 1, 2 is monomeric and it contains an indium atom in a trigonal planar coordination.  相似文献   

3.
Synthesis, Properties, and Structure of cis-Dihydrido Azido Tris(triphenylphosphino) Iridium cis-IrH2(N3)(PPh3)3 is formed from IrCl3(PPh3)3 and NaN3 in alcohol. The solvent transfers the hydride ions whereby aldehyde is formed. The creme-colored cis-IrH2(N3)(PPh3)3 is a diamagnetic low-spin complex. Exposure to light yields H2 and Ir(N3)(PPh3)3 in a reversible process. The cis position of the hydrido ligands is confirmed by the i.r. and H-N.M.R. spectra. Cis-IrH2(N3)(PPh3)3 crystallizes in the monoclinic system with the space group P21/c. The crystal structure exhibits isolated octahedral complexes, in which one hydrido ligand is located trans to the azido group. The other one being trans to one of the phosphine ligands.  相似文献   

4.
Ag3PO4 is widely used in the field of photocatalysis because of its unique activity. However, photocorrosion limits its practical application. Therefore, it is very urgent to find a solution to improve the light corrosion resistance of Ag3PO4. Herein, the Z-scheme WO3(H2O)0.333/Ag3PO4 composites are successfully prepared through microwave hydrothermal and simple stirring. The WO3(H2O)0.333/Ag3PO4 composites are characterized by X-ray diffraction, scanning electron microscopy, X-ray photoelectron spectroscopy and UV-Vis spectroscopy. In the degradation of organic pollutants, WO3(H2O)0.333/Ag3PO4 composites exhibit excellent performance under visible light. This is mainly attributed to the synergy of WO3(H2O)0.333 and Ag3PO4. Especially, the photocatalytic activity of 15%WO3(H2O)0.333/Ag3PO4 is the highest, and the methylene blue can be completely degraded in 4 min. In addition, the stability of the composites is also greatly enhanced. After five cycles of testing, the photocatalytic activity of 15%WO3(H2O)0.333/Ag3PO4 is not obviously decreased. However, the degradation efficiency of Ag3PO4 was only 20.2%. This indicates that adding WO3(H2O)0.333 can significantly improve the photoetching resistance of Ag3PO4. Finally, Z-scheme photocatalytic mechanism is investigated.  相似文献   

5.
The maximum monolayer dispersion (the threshold) for WO3 on γ-Al2O3 calcined at 500°, 550°, 600°, and 640°C has been determined quantitatively by XRD (amount of crystalline phase) and XPS (intensity ratios Iw4f/IAl2). The results show that if the amount of WO3 loaded is lower than the maximum monolayer dispersion, WO3 will react with γ-Al2O3 to form surface compound due to mutual ionic interaction, and will be dispersed on γ-Al2O3 surface as monolayer then. In case the amount is higher than this value, the residual crystalline WO3 will remain. The maximum monolayer dispersion (threshold) is 0.21 g and 0.20 g WO3/100 m2 γ-Al3O3 by XRD and XPS respectively. It agrees with the value (0.189 g WO3/100 m2 or 4.90 × 10?18 W atoms/m2) calculated from the model on assumption that the WO3 is dispersed as a closed-packed monolayer on γ-Al2O3 surface. Inasmuch as WO3/γ-Al2O3 system is stable up to higher temperature, e.g. 700°C, than MoO3/γ-Al2O3 system, WO3 seems unfavorable to form new bulk compound with γ-Al2O3 at that temperature. However, Al2(MoO4)3 forms perceptibly in MoO3/γ-Al2O3 system at 500°C. Besides, the size of residual crystalline WO3 in WO3/γ-Al2O3 is much smaller than that of MoO3 in MoO3/γ-Al2O3. It might be the reason that WO3/γ-Al2O3 catalyst is superior to MoO3/γ-Al2O3 in hydrodesulfurization (HDS) or hydrodenitrogenation (HDN) in some cases.  相似文献   

6.
The preparation is described of [CoCO3(NH3)5]ClO4 · H2O, trans-[CoCO3(NH3)4(15NH3)]ClO4, and trans-[CoCO3(NH3)4(NH2CH3)]ClO4. The transformation reactions of these complexes, in which a chelate carbonate ligand is formed and one NH3 is eliminated, were studied in solution and in the solid state. 1H NMR spectroscopy is used for the identification of the products. It is shown that the transformation reactions are not stereospecific.  相似文献   

7.
The reaction mechanism of (CH3)3CO. radical with NO is theoretically investigated at the B3LYP/6-31G* level. The results show that the reaction is multi-channel in the single state and triplet state. The potential energy surfaces of reaction paths in the single state are lower than that in the triple state. The balance reaction: (CH3)3CONO⇔(CH3)3CO.+NO, whose potential energy surface is the lowest in all the reaction paths, makes the probability of measuring (CH3)3CO. radical increase. So NO may be considered as a stabilizing reagent for the (CH3)3CO. radical.  相似文献   

8.
In this paper we report on the influence of light and oxygen on the stability of CH3NH3PbI3 perovskite‐based photoactive layers. When exposed to both light and dry air the mp‐Al2O3/CH3NH3PbI3 photoactive layers rapidly decompose yielding methylamine, PbI2, and I2 as products. We show that this degradation is initiated by the reaction of superoxide (O2?) with the methylammonium moiety of the perovskite absorber. Fluorescent molecular probe studies indicate that the O2? species is generated by the reaction of photoexcited electrons in the perovskite and molecular oxygen. We show that the yield of O2? generation is significantly reduced when the mp‐Al2O3 film is replaced with an mp‐TiO2 electron extraction and transport layer. The present findings suggest that replacing the methylammonium component in CH3NH3PbI3 to a species without acid protons could improve tolerance to oxygen and enhance stability.  相似文献   

9.
Nitromethane is the only presently known organic solvent for highly reactive selenium trioxide (SeO3)4. The stability of the solutions is limited and the beginning of the reaction between both components depends significantly on concentration and temperature. The nitromethane solvate of cyclic triselenium heptoxide Se3O7 · CH3NO2 is the major solid product at the temperature 20–30°C and concentration range 3–20% SeO3. Crystal and molecular structure of this compound was determined by X-ray structure analysis and vibrational spectroscopy. The solvating molecule CH3NO2 is removable from Se3O7 · CH3NO2 in vacuo. If reaction temperature does not exceed 10°C, selenium pentoxide (Se2O5)n is formed instead of Se3O7 · CH3NO2. Dinitrosyl triselenate (NO)2Se3O10, nitrosyl hydrogendiselenate NOHSe2O7, nitrosyl hydrogenselenate NOHSeO4, nitrosyl hydrogenselenatoselenite NOHSe2O6 and selenium dioxide (SeO2)n were further identified in the solid reaction products. The selenic and/or oligoselenic acids remains in the nitromethane solution. CO2 and N2O3 were found as gaseous products.  相似文献   

10.
Azido Derivatives of the Pentamethylcyclopentadienyl Vanadium(IV)-Fragment. Molecular Structures of the Binuclear Complexes [Cp*VCl(N3)(μ-N3)]2 and [Cp*V(N3)2(μ-N3)]2 The stepwise reaction of Cp*VCl3 with excess trimethylsilyl azide (Me3Si–N3) in solution leads to the paramagnetic, azido-bridged complexes [Cp*VCl2(μ-N3)]2 ( 3 ), [Cp*VCl(N3)(μ-N3)]2 ( 4 ) and [Cp*V(N3)2(μ-N3)]2 ( 5 ) which were characterized by their IR and mass spectra. The azide-rich binuclear complex 5 is also formed if a pentane solution of Cp*V(CO)4 is stirred in the presence of excess Me3Si–N3 in an open vessel. According to the X-ray structure analyses both 4 and 5 are centrosymmetric molecules with a planar V(N)2V four-membered ring. In the absence of free trimethylsilyl azide, solutions of 3 – 5 lose dinitrogen slowly; in the presence of traces of air, 5 is thereby converted to the diamagnetic, oxo-bridged complex [Cp*V(O)(N3)]2(μ-O) ( 6 ).  相似文献   

11.
Structural parameters of the cation [C(NPCl3)3]+ in vacuum and in acetonitrile are calculated by the methods RHF/6-311G(3d6), RHF/6-31(3d5) and B3LYP/6-311(3d5f7). Formation energy of the free adduct (MeCN) 2[C(NPCl3)3]+ is calculated and nonspecific character of interaction of the cation with liquid acetonitrile is established. Vibration spectrum of the cation is calculated and theoretical interpretation of IR and Raman spectra of the salt [C(NPCl3)3]+[SbCl6]? is refined.  相似文献   

12.
The potential of several alkylcobalt complexes as catalysts for hydrogenation and isomerization of alkenes has been investigated. The complexes CH3Co(CO)2(Pom-Pom) (Pom-Pom = 1,2 bis(dimethoxyphosphino)ethane), CH3Co(CO)3P(OMe)3 and C6H5CH2Co(CO)3PPh 3 are compared to CH3Co(CO)2(P(OMe)3)2, for their ability to function in catalytic cycles. Each is active for hydrogenation and isomerization of alkenes under conditions where the carbonylation-decarbonylation equilibrium is readily established. The lifetime for the complexes is much shorter than for CH3Co(CO)2 (P(OMe)3)2 suggesting that two phosphorus donors in trans positions in an intermediate is a requirement for catalyst stability in these alkylcobalt complexes.  相似文献   

13.
Phosphoraneiminato Complexes of Titanium(IV). Crystal Structures of [TiCl3(NPEt3)]2, [TiCl3(NPEt3)(THF)2], and [TiCl4{Me2Si(NPEt3)2}] [TiCl3(NPEt3)]2 ( 1 ) is formed from titanium(IV) chloride and the silylated phosphaneimine Me3SiNPEt3 in dichloromethane as reddish-brown, moisture-sensitive crystals. According to the crystal structure analysis these crystals show centrosymmetric Ti2N2 four-membered rings with Ti–N distances of 184.7 and 210.3 pm. With tetrahydrofurane 1 forms yellow, moisture sensitive crystals of the solvate [TiCl3(NPEt3)(THF)2] ( 2 ), in which the titanium atom is octahedrally coordinated. The THF molecule which is in trans position to the phosporaneiminato ligand realizes but a very weak Ti–O bond of 238.0 pm, the cis THF molecule shows a Ti–O distance of 213.7 pm. With 173.4 pm along with a TiNP bond angle of 160.0° the TiN distance is very short. The bis(phosphaneimine) complex [TiCl4{Me2Si(NPEt3)2}] ( 3 ) is formed as colourless crystals in low yield in the reaction of titanium(IV) chloride with Me3SiNPEt3 and trimethylcyclopentadienylsilane. In 3 the titanium atom is surrounded by four chlorine atoms in a distorted octahedral fashion and by the two N atoms of the Me2Si(NPEt3)2 molecule with TiN distances of 205.6 pm.  相似文献   

14.
We report about quantum chemical ab initio calculations at the MP2/6‐311+G(2d)//MP2/6‐31G(d) level and DFT calculations at BP86/TZP of the geometries and bond dissociation energies of the borane‐phosphane complexes X3B‐PY3 and the alane‐phosphane complexes X3Al‐PY3 (X = H, F, Cl; Y = F, Cl, Me, CN). The nature of the B‐P and Al‐P bonds is analyzed with a bond energy partitioning method. The calculated bond dissociation energies De of the borane adducts X3B‐PY3 show for the phosphane ligands the trend PMe3 > PCl3 ∼ PF3 > P(CN)3. A similar trend PMe3 > PCl3 > PF3 > P(CN)3 is predicted for the alane complexes X3Al‐PY3. The order of the Lewis acid strength of the boranes depends on the phosphane Lewis base. The boranes show with PMe3 and PCl3 the trend BH3 > BCl3 > BF3 but with PF3 and P(CN)3 the order is BH3 > BF3 > BCl3. The bond energies of the alane complexes show always the trend AlCl3 ≥ AlF3 > AlH3. The bonding analysis shows that it is generally not possible to correlate the trend of the bond energies with one single factor which determines the bond strength. The preparation energy which is necessary to deform the Lewis acid and Lewis base from the equilibrium form to the geometry in the complex may have a strong influence on the bond energies. The intrinsic interaction energies may have a different order than the bond dissociation energies. The trend of the interaction energies are sometimes determined by a single factor (Pauli repulsion, electrostatic attraction or covalent bonding) but sometimes all components are important. The higher Lewis acid strength of BCl3 compared with BF3 in strongly bonded complexes is not caused by the deformation energy of the fragments but it is rather caused by the intrinsic interaction energy. P(CN)3 is a weaker Lewis base than PF3, PCl3 and PMe3 mainly because of its weaker electrostatic attraction. The complex H3B‐P(CN)3 is predicted to have a bond dissociation energy Do = 14.8 kcal/mol which should be sufficient to synthesize the compound as the first adduct with the ligand P(CN)3. The calculated bond energies at the BP86 level are in most cases very similar to the MP2 results. In a few cases significantly different absolute values have been found which are caused by the method and not by the quality of the basis set.  相似文献   

15.
The reaction between liquid F3SiI and red HgS mainly yields the disilylsulfane (F3Si)2S and, in smaller amounts, hitherto unknown (F3SiS)2SiF2. These fluorosilylsulfanes have different thermal stabilities. (F3Si)2S is a versatile precursor for F3Si derivatives, and at ambient temperature it is stable, while (F3SiS)2SiF2 decomposes rapidly. F3SiSH has been obtained, along with F3SiBr, by selective cleavage of one of the Si–S bonds in (F3Si)2S with HBr in the liquid phase and was for the first time unambiguously characterized. Contrary to previous reports, (F3Si)2O rather than (F2SiS)2 is formed when SiF4 is passed over SiS2 at 1298 K in a quartz tube. Raman and infrared spectra of (F3Si)2S, F3SiSH and its deuterated derivative F3SiSD have been measured and assigned to vibrational fundamentals. Multinuclear NMR spectra have been recorded.  相似文献   

16.
Mixtures of CaCO3 and varying amounts of Na2CO3, K2CO3 and NaCl were subjected separately to thermal analysis. DTG, DTA, TG analyses indicate that the presence of alkali salts in CaCO3 influences its decomposition behaviour. A minimum DTA peak temperature of CaCO3 decomposition is noticed at low concentrations of alkali salts (K2CO3 and Na2CO3); an increase in concentration increases the DTA peak temperature. However, in the case of NaCl no appreciable lowering of the DTA peak temperature of CaCO3 decomposition is observed. Similarly, the minimum temperature at which decomposition completes is found to correspond to the concentration of 1 per cent salt (K2CO3 and Na2CO3) in CaCO3.  相似文献   

17.
Transition Metal Silyl Complexes, 44. — Preparation of the Binuclear Silyl Complexes (CO)3(R3Si)Fe(μ-PR′R′′)Pt(PPh3)2 by Oxidative Addition of (CO)3(R′R′′HP)Fe(H)SiR3 to (C2H4)Pt(PPh3)2 The complexes (CO)3(R′R′′HP)Fe(H)SiR3 ( 1 ) [PHR′R′′ = PHPh2, PH2Ph, PH2Cy; SiR3 = SiPh3, SiPh2Me, SiPhMe2, Si(OMe)3] react with Pt(C2H4)(PPh3)2 to give the dinuclear, silyl-substituted complexes (CO)3(R3Si)Fe(μ-PR′R′′)Pt(PPh3)2 ( 2 ) in high yields. Upon reaction of 2 (R = R′ R′′ = Ph) with CO, the PPh3 ligand at Pt being trans to the PPh2 bridge is exchanged, and (CO)3(Ph3Si)Fe(μ-PPh2)Pt(PPh3)CO ( 3 ) is formed. Complex 3 is characterized by an X-ray structure analysis. The rather short Fe — Si distance [233.9(2) pm] and the infrared spectrum of 3 indicate that the Fe — Pt bond is quite polar.  相似文献   

18.
The system 4CaO·3Al2O3·SO3-CaSO4·2H2O-Ca(OH)2 was hydrated in the presence of ten dopants, specifically soluble salts of heavy metals.When added in 10% amount, the effect of each salt is strongly evident at shorter curing times, the hydration kinetics being more favoured in the order Pb(NO3)22CrO4< Cd(NO3)2< Zn(NO3)2t~Mn(NO3)3=K2MoO43)23)23)33)3. At longer curing times the differences among the systems decrease significantly.The 28-day compressive strength is almost the same for all the systems except those containing Pb(NO3)2, K2MoO4 and K2CrO4.  相似文献   

19.
Quantum chemical calculations at various levels of theory (BP86, B3LYP, MP2, CCSD(T), CBS‐QB3) of the beryllium complexes [BeCl2(NHPH3)], [BeCl2(NHPH3)2], [BeCl3(py)]?, [BeCl2(NH3)], [BeCl2(NH3)2], [BeCl3(py)]? and [BeCl3(NH3)]? as well as the boron compounds [BCl3(py)] and [BCl3(NH3)] show that BeCl2 is a very strong Lewis acid. The theoretically predicted bond dissociation energy at CBS‐QB3 of Cl2Be‐NH3 (De = 32.5 kcal/mol)is even higher than that of Cl3B‐NH3 (De = 28.6 kcal/mol). Even the second ammonia molecule in [BeCl2(NH3)2] still has a strong bond with De = 24.2 kcal/mol. The theoretically predicted bond strengths for the phosphaneimine ligands in [BeCl2(NHPH3)2] are De = 46.7 kcal/mol for the first ligand and De = 29.8 kcal/mol for the second. The anion BeCl3? is a moderately strong Lewis acid which has bond energies of De = 14.1 kcal/mol in [BeCl3(py)]? and De = 14.2 kcal/mol in [BeCl3(NH3)]?. The higher bond energy of [BeCl2(NH3)] compared with [BCl3(NH3)] comes from less deformation energy for BeCl2 than for BCl3. The intrinsic attraction between BeCl2 and NH3 calculated with frozen geometries of the complex geometry is ~5 kcal/mol less than the attraction between BCl3 and NH3. The bonding analysis with the EDA method shows that the attractive energy of the beryllium complexes comes manly from electrostatic attraction. The larger contribution of the electrostatic term is the most significant difference between the nature of the donor‐acceptor bonds of the beryllium and boron complexes.  相似文献   

20.
The far IR (450–480 cm?1) and Raman (3200–3230 cm?1) spectra of (CH3)3 NGaCl3 have been recorded in the solid state and interpreted in detail on the basis of C3 molecular symmetry. A modified valence force field model is used to calculate the frequencies and potential energy distribution of the adduct. The calculated force constants of the adduct are compared with those previously reported for the free Lewis acid and the free Lewis base moieties, and the observed differences ascribed to geometrical changes of the uncomplexed species on adduct formation and explained on the basis of the VSEPR model and non-bond interactions. Extensive coupling is observed between the GaN stretching mode and the NC3 symmetric stretching and the NC3 symmetric deformational modes. Strong coupling interaction is also found between the GaCl3 antisymmetric stretch and the NC3 antisymmetric deformation. The calculated value of 2.50 mdyn Å?1 for the GaN stretching force constant in (CH3)3NGaCl3 is larger than any of those previously determined in complexes such as (CH3)3NGaH3 (2.43 mdyn Å?1), (CH3)3NGa(CH3)3 (1.61 mdyn Å?1), and H3NGa(CH3)3 (1.08 mdyn Å?1). The observed variations in the magnitudes of the stretching force constants of the donor—acceptor dative bond is found to be consistent with the estimated relative stabilities of this series of adducts.  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