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 324 毫秒
1.
Seeded supersonic NO beams were used to study the kinetic energy dependence of both the electronic (NO2*) and vibrational (NO23) chemiluminescence of the NO + O3 reaction. In addition the electronic CL is found to be enhanced by raising the NO internal temperature. This is shown to be due to enhanced reactivity of the NO(2Π,32) fine structure component. By difference NO(2Π12) is concluded to yield predominantly groundstate NO23. The excitation function for NO2* formation from NO(2Π32) is of the form σ32(E) = C(E/E0 - 1)n over the 3–6 kcal energy range where n = 2.4 ± 0.15, C = 0.163 Å2 and E0 = 3.2 ± 0.3 kcal/mole. Vibrational IR emission from NO23 has an energy dependence different from electronic NO2* emission, confirming that emitters are formed predominantly in distinct reaction channels rather than via a common precursor (either NO2* or NO23). The short wavelength cutoff of the CL spectra recorded at elevated collision energies E ? 15 kcal/mole corresponds to the total available energy. These and literature results are discussed in the light of general properties of the (generally unknown) ONO3 potential energy surfaces. The formation of electronically excited NO2* rather than energetically preferred O2 (1 Δg) (Gauthier and Snelling) can be rationalized in terms of surface hopping near a known intersection of potential energy surfaces more easily than by vibronic interaction in the asymptotic NO2 product.  相似文献   

2.
K3Sb3P2O14 crystallizes in the rhombohedral system, space group R3m with a = 7.147(1) Å, c = 30.936(6) Å, Z = 3. The structure was determined from 701 reflections collected on a Nonius CAD4 automatic diffractometer with MoKα radiation. The final R index and the weighted Rw index are 0.033 and 0.042, respectively. The structure is built up from layers of SbO6 octahedra and PO4 tetrahedra sharing corners. The potassium ions are situated between the (Sb3P2O14)3? covalent layers.  相似文献   

3.
The LiPO3CeP3O9 and NaPO3CeP3O9 systems have been investigated for the first time by DTA, X-ray diffraction, and infrared spectroscopy. Each system forms a single 1:1 compound. LiCe(PO3)4 melts in a peritectic reaction at 980°C. NaCe(PO3)4 melts incongruently, too, at 865°C. These compounds have a monoclinic unit cell with the parameters: a = 16.415(6), b = 7,042(6), c = 9.772(7)Å; β = 126.03(5)°; Z = 4; space group C2c for LiCe (PO3)4; and a = 9.981(4), b = 13.129(6), c = 7.226(5) Å, β = 89.93(4)°, Z = 4, space group P21n for NaCe(PO3)4. It is established that both compounds are mixed polyphosphates with chain structure of the type |MIIMIIIII (PO3)4|MII: alkali metal, MIIIII: rare earth.  相似文献   

4.
Use of Nd3+, Eu3+, and Gd3+ as local structural probes allows the determination of the rare earth positions in the NaxSr3?2xLnx(PO4)2 (Ln = La to Tb) and KCaLn(PO4)2 phases (Ln = rare earth). Moreover, a common feature of both series is a particularly high splitting of the excitation 6P72 and 6P52 levels of the Gd3+ ions.  相似文献   

5.
Perovskites of the type A2+3B2+M5+2O9, where A2+ = Ba, Sr; B2+ = Mn, Co, Ni, Zn; M5+ = Nb, Ta, show order-disorder phenomena. At lower temperatures a thermodynamically unstable disordered cubic perovskite is formed (13 formula unit—AB13M23O3—in the cell), which transforms irreversibly into a 1: 2 ordered high-temperature form with 3L structure (sequence (c)3). For A2+ = Ba this lattice is hexagonal (space group P3m1; one formula unit in the cell); with A2+ = Sr a triclinic distortion is observed. For Ba3CoNb2O9 a second transformation into a cubic disordered perovskite takes place at 1500°C. This transition is reversible and of the order-disorder type. The vibrational and diffuse reflectance spectra are discussed.  相似文献   

6.
7.
Crystals of RbEu3F10 are cubic with a = 11.844 Å, Z = 8, and three possible space groups, Fm3m, F43m, and F432. The structure has been resolved from three-dimensional X-ray data and refined by the least-squares method. The final R values are, respectively, 0.065 and 0.067 for Fm3m and F43m, and are not significantly different. The rare earth symmetry is C4v for Fm3m and C2v for F43m. Consequently, we have used the Eu3+ ion in the RbEu3F10 phase as a structural probe in order to state precisely the symmetry of the lanthanide site and distinguish among the possible space groups.  相似文献   

8.
TlMnI3 and TlFeI3 are isostructural with NH4CdCl3. TlMnI3 has a spiral structure which can be described with an incommensurable vector k, in the direction of the b1 axis of length 0.3614(5)b1. The spins lie in the (0 0 1) plane. TlMnI3 exhibits antiferromagnetic behavior with a Néel temperature of 6.0(2) K. The exchange interaction was calculated to be zJk = ?1.6 K, z being the number of nearest neighbors. Discontinuities in the magnetization are found for both the [1 0 0] and [0 1 0] directions at fields HaSF = 30.1(2) kOe and HbSF = 14.1(2) kOe. The magnetic structure of TlFeI3 consists of puckered ferromagnetic (1 0 0) planes, which are coupled antiferromagnetically. The magnetic moments are parallel to the b axis. The Néel temperature is 21.5(3) K. zJk was found to be ?10(1) K with g = 2.68 and s = 2. The magnetic structures found for TlMnI3 and TlFeI3 are derived taking into account inter- and intra-double-chain interactions via two I? ions.  相似文献   

9.
Trimethyl- and triphenyl-tin(IV) hydroxide act on triphenyl(2-carboxyethyl)-phosphonium hydrochloride, which is made from 3-chloropropionic acid and triphenylphosphine, to release water in the presence of dimethylformamide (DMF) as a catalyst. The water is azeotropically distilled to drive the reaction forward and produce triphenylphosphonopropionbetainetrimethyl- and triphenyl-tin(IV) chlorides in high yield. The latter product also results from the displacement of chloride from triphenyltin(IV) chloride by the phosphobetaine, (C6H5)3P(CH2)2CO2, which is made by treating the phosphonium hydrochloride with bicarbonate, and the compounds [(C6H5)3P(CH2)2CO2Sn(C6H5)3]+X? where X = Cl, Br, I, N3, NCS, NO3, B(C6H5)4 and Co(CO)4 are made in the same way. The acetate salt results from metathesis from the chloride and lead(II) acetate. A double salt, [(C6H5)3P(CH2)2CO2SnR3]+ [R3SnX2]?, is formed for X = Cl, Br and N3 by adding additional (C6H5)3SnX to the already-formed simple salts. Double salts are also obtained from the 11 reactions between the phosphobetaine and triphenyltin(IV) isocyanate and methyldiphenyltin(IV) chloride. The phosphonium chloride double salt could be converted to the thiophosphonium derivative by heating with elemental sulfur in ethanol. The products of these novel nucleophilic displacement reactions are high melting solids. Tin-119m Mössbauer data are consistent with five-coordinated, triorganotin(IV) formulations with the exception of the diphenyl(8-hydroxyquinolinato)tin(IV) chloride salt in which the tin atom is six-coordinated, and the diphenyltin system cis-oriented. The parameters otherwise do not change with the nature of the X group, which is the tetracarbonylcobaltate derivative is tetrahedral by infrared, establishing the ionicity of the products. The chloride exhibits a molar conductivity indicative of a 11 electrolyte in DMF. A bridging acetate structure in the solid is consistent with the lowered ν(CO2) frequencies. The Mössbauer spectra of the double salts give simple doublets of lowered isomer shift (IS) and raised quadrupole splitting (QS) which may arise from a cross-linking ion pairing of the polymer chains in the solid, and the NMR spectra of the two methyltin derivatives shows only a single resonance line and tin satellites which is rationalized by a dynamic exchange process. The products are formulated as associated in carboxylate polymers with dangling triphenylphosphonium cations.  相似文献   

10.
The reaction of IrH3(PPh3)2 with p-substituted aryldiazonium salts gives the compounds [IrH2(NHNC6H4R)(PPh3)2]+BF4- at low temperature (-10°C) and the o-metalated complexes [IrH(NHNC6H3R)(PPh3)2]+BF4- (R  F, OCH3) at 40–50°C. The reactions of the o-metalated complexes with CO, PPh3, NaI and HCl have been studied.  相似文献   

11.
The structure of two new oxides KCuTa3O9 and KCuNb3O9 has been solved from X-ray powder data and by electron microscopy. Both compounds are orthorhombic, space group Pnc2 with a ? 8.8 Å, b ? 10.1 Å, and c ? 7.6 Å. Their host lattice is built up from corner-sharing MO6 octahedra (M = Nb, Ta) forming pentagonal tunnels where the K+ ions are located. The copper ions are located in distorted perovskite CaCu3Mn4O12-type cages and exhibit a square planar coordination. The relationships between these oxides and the TTB, HTB, ITB, and Ba0.15WO3 structures are discussed.  相似文献   

12.
The reactions of the sulphite radical anion, SO3.?? (generated from the Ti3+-H2O2-Na2SO3 system at pH 9), in aqueous solutions with some nitroalkane compounds were investigated by using a rapid-mixing flow technique coupled with electron spin resonance (ESR) which can detect the radicals having a lifetime of 5–100 ms. The SO3.?? radical added to the double bond of CN in the nitroalkane aci-anions which are the main form of nitroalkanes in aqueous alkaline solutions. From the observed hyperfine splitting constants of the SO3.?? adducts of nitroalkane aci-anions, the preferred conformation of the adducts was deduced.  相似文献   

13.
Laser excitation of equilibrium vapor mixtures ErCl3(s)-ACl3(g) (A = Al, Ga, In) at 475–1100 K gives rise both to resonance fluorescence from the f → f Er3+ transitions of the Er-Cl-A vapor complexes, and to Raman scattering due to the vibrational modes of the ACl3 vapor. The laser-induced fluorescence from the 4F92, 4S32 and 2H112 states has been investigated at different temperatures and excitation.  相似文献   

14.
Phosphorus pentafluoride was reported long ago to give adducts 2 PF5 ·5 NH3 (1) and nNH3·PF5 (n= 1 ? 4) (2). None of the compounds was characterised in detail. Repeating the reaction of PF5 and NH3 we found the adduct H3N·PF5, 1, in 8% yield besides (H2N) 2PF3 (3) and NH4PF6. However, HF and (F2P=N)3 gave 1 in 41% yield. The 1H, 19F, and 31P n.m.r. spectra of 1 exhibit 14NH, 14NPF(cis), and 14NP coupling. The x-ray structure determination shows almost perfect octahedral geometry at phosphorus with a PN bond length of 1.842 ā. Compound 1 is soluble in water without decomposition. Treatment with NH3 leads to the anion H2NPF5?. Upon heating 1 forms in good yield H2NPF4 and NH4PF6. Without a solvent 1 and NH3 react to give (H2N) 2PF3. A mechanism for the ammonolysis of PF5 is proposed.  相似文献   

15.
Proton NMR relaxation times (T2T1, and T1?) and absorption spectra are reported for the compounds H1.71MoO3 (red monoclinic) and H0.36MoO3 (blue orthorhombic) in the temperature range 77 K < T < 450 K. Rigid lattice dipolar spectra show that both compounds contain proton pairs, as OH2 groups coordinated to Mo atoms in H1.71MoO3 and as pairs of OH groups in H0.36MoO3. The room temperature lineshape for H1.71MoO3 shows that the average chemical shielding tensor has a total anisotropy of 20.1 ppm. The relaxation measurements confirm that hydrogen diffusion occurs and give EA = 22 kJ mole?1 and τ0C ? 10?13sec for H1.71MoO3 and EA = 11 kJ mole?1 and τ0C ? 3 × 10?8sec for H0.36MoO3.  相似文献   

16.
A new experimental system for atomic resonance spectrometry at λ < 105 nm in a discharge-flow system is described. The spectrum of a fluorine resonance lamp has been studied, and possible precursors for the 2p4 3s excited F atoms formed are suggested. Ground state (2p52P32) and J-excited 2P12 F atoms have been detected for the first time in resonance absorption and fluorescence using the first resonance transitions with wavelengths between 95.2 and 97.8 nm. Preliminary measurements (using both 4P-2P and 2P-2P lines) of the variations with concentration of absorption intensity by ground state F 2P32 and by J-excited F 2P12 atoms are reported; F atom concentrations were measured using a titration method based on the rapid reaction, F + Cl2 → FCl + Cl.  相似文献   

17.
The tracer diffusion coefficient, D1O, of oxide ions in LaFeO3 single crystal was determined over the temperature range of 900–1100°C by the gas-solid isotopic exchange technique using 18O as a tracer. For the determination of D1O, the depth profile of 18O was measured by means of a secondary ion mass spectrometer (SIMS). The surface exchange reaction was found to be slow and the surface exchange rate constant, k, was determined together with D1O. It was found that D1O at 950°C is proportional to P?0.58O2, where PO2 is an oxygen pressure. The vacancy mechanism was determined for the diffusion of oxide ions from the PO2 dependence. The vacancy diffusion coefficient, DV, for LaFeO3 was nearly the same as that for LaCoO3 at the same temperature. The activation energy for migration of oxide ion vacancies was 74 kJ · mole?1 for both oxides.  相似文献   

18.
The acid catalysed rearrangement of 8-methyl-pentacyclo(5.4.0.02,6.03,10.05,9)undecan-8-endo-o1 (8) to 3-methyl-(D3)-trishomocuban-4-o1 (9) provided the key step to the synthesis of the title compound (11).  相似文献   

19.
The formation of the XeF+ ion by ion-molecule reaction has been observed in an ionized mixture of Xe and NF3 by ion cyclotron resonance mass spectrometry. The excited 2P12 state of the xenon ion has unambiguously been identified as the major precursor by photoionization mass spectrometry. The NF+3 ion makes an additional minor contribution. Evidence suggests that the excited 2P12 xenon ion radiatively decays to the 2P32 ground state on the time scale of the experiment. The transition probability deduced for this dipole forbidden emission, 18 ± 4 s?1, is in good agreement with the theoretical value of 21 s?1 for the sum of the magnetic dipole and electric quadrupole transition rates.  相似文献   

20.
The excitation-transfer reaction in thermal energy collisions of state-selected metastable Ar*(3P2) and Ar*(3P0) atoms with ground state H atoms, giving excited H*(n = 2) atoms, has been studied with the stationary afterglow technique. The rate constant for the excitation of H atoms by Ar*(3P2) has been found to be more than one order of magnitude larger than in excitation by Ar*(3P0). This difference in the reactivity of two metastable species is explained to be a consequence of the attractive nature of the D(2II) and E(2Σ+) potentials that develop from the Ar*(3P2)+H entrance channel and which give curve crossing with the B(2II) and C(2Σ+ potentials, respectively, leading to the Ar+H*(n=2) exit channel, whereas only a repulsive 4II (Ω=12) potential develops from the Ar*(3P0+H entrance channel.  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