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 46 毫秒
1.
V—V energy transfer from a large molecule excited to vibrational energies of chemical interest has been demonstrated by detection of ≈ 1.5% yield of CO2(001) due to energy transfer from azulene (Evib ≈ 30600 cm?1. Also, the average enery lost per collision by azulene was measured as a function of Evib, and the rate constant for CO2(001) deactivation by azulene was determined.  相似文献   

2.
Emission spectra resulting from reaction of “clean” N2(A3 Σu+) with copper atoms were studied using a flowing afterglow apparatus. The population distribution of the Cu states was calculated from the spectrum; it indicates that Cu atoms are excited by nearly resonant energy transfer processes. N2(A,v') + Cu(2S12) → N2(X, v) + Cu* , and that the transfer is most efficient for N2(A,v') → N2(X,v) transitions with large Franck-Condon factors. The preferential energy transfer results in population inversion between some of the Cu states.  相似文献   

3.
4.
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   

5.
The crystal structure of the new Bi∼3Cd∼3.72Co∼1.28O5(PO4)3 has been refined from single crystal XRD data, R1=5.37%, space group Abmm, a=11.5322(28) Å, b=5.4760(13) Å, c=23.2446(56) Å, Z=4. Compared to Bi∼1.2M∼1.2O1.5(PO4) and Bi∼6.2Cu∼6.2O8(PO4)5, this compound is an additional example of disordered Bi3+/M2+ oxyphosphate and is well described from the arrangement of double [Bi4Cd4O6]8+ (=D) and triple [Bi2Cd3.44Co0.56O4]6+ (=T) polycationic ribbons formed of edge-sharing O(Bi,M)4 tetrahedra surrounded by PO4 groups. According to the nomenclature defined in this work, the sequence is TT/DtDt, where t stands for the tunnels created by PO4 between two subsequent double ribbons and occupied by Co2+. The HREM study allows a clear visualization of the announced sequence by comparison with the refined crystal structure. The Bi3+/M2+ statistic disorder at the edges of T and D entities is responsible for the PO4 multi-configuration disorder around a central P atom. Infrared spectroscopy and neutron diffraction of similar compounds (without the highly absorbing Cadmium) even suggests the long range ordering loss for phosphates. Therefore, electron diffraction shows the existence of a modulation vector q*=1/2a*+(1/3+ε)b* which pictures cationic ordering in the (001) plane, at the crystallite scale. This ordering is largely lost at the single crystal scale. The existence of mixed Bi3+/M2+ positions also enables a partial filling of the tunnels by Co2+ and yields a composition range checked by solid state reaction. The title compound can be prepared as a single phase and also the M=Zn2+ term can be obtained in a biphasic mixture. For M=Cu2+, a monoclinic distortion has been evidenced from XRD and HREM patterns but surprisingly, the orthorhombic ideal form can also be obtained in similar conditions.  相似文献   

6.
By measurement of infrared chemiluminescence we have obtained for the branching ratio of the room temperature reaction H + Br2 (1), k*1/k1 = 0.015 ± 0.004 and for H + HBr (2), k*2/k2 ? 0.013. For H + Br2 → HBr(υ· ? 6) + Br (1), the detailed rate constant k* = 6) = 0.014 ± 0.003 relative to k· = 4) = 100.  相似文献   

7.
The ultraviolet luminescence from the Hg-photosensitised reaction of ammonia was investigated at pressures up to 10 atmospheres. From a variation of the wavelength distribution on the [NH3], it was concluded that Hg(63P0) can attach clusters of NH3 molecules to form Hg(NH3)n* with n up to at least 5. The emission profiles of the stabilized complexes with n = 1–4 were determined, and also the profile from unstabilised HgNH3* formed in a bimolecular encounter of Hg(63P0) with NH3. Dissociation constants for complexes with n = 2, 3 and 4 were measured.  相似文献   

8.
The 00 band maximum of the transition T3(π, π*) ← T1 (π, π*) of acridine occurs at ≈ 10200 ± 20 cm?1 in inert (n-hexane, benzene, CCl4), at 10220 ± 20 cm?1 in polar (acetonitrile) and at 10170 ± 50 cm?1 in hydrogen-bonding (methanol, 2-propanol and alkaline water) solvents. Based on the solvent-independent energy of T1 (π, π*), the T3(π, π*) state of acridine is estimated at 26050 ± 50 cm?1 in all the solvents.  相似文献   

9.
The diffusion coefficient of O*2(1Δg) in O2(3Σ?g) has been measured as a function of pressure, D* = 0.201 ± 0.005 cm2 s?1 at 1 atmosphere and 298 K.  相似文献   

10.
The Balmer α and β lines produced in e-NH3 collisions have been measured precisely with the use of a Fabry-Perot interferometer. These lines are not polarized. The translational energy distributions of (H*(n = 3,4) were determined from analysis of Doppler lineshapes and have five components; their peaks lie at 1, 3, 2, 4–5 and 8–12 eV. The excitation function [H*(n = 4)] has five thresholds at 22.5, 29.0, 33.3, 38.9 and 41.7 eV, and indicates that five dissociation processes contribute to the formation of H*. Excitation to the Rydberg states converging to the (2a1)?1 state of NH3+ is a major process for the formation of the first and the second components. Doubly excited Rydberg states play important roles in the dissociative excitation of NH3.  相似文献   

11.
The jet-cooled fluorescence spectra of perylene excited to the S1 state with Evib = 0–1600 cm?1 are recorded and analyzed. For Evib <800 cm?1 only the resonant fluorescence was detected. Ground- and excited-state frequencies of 14 low-frequency normal modes are determined. A drastic change in frequency of the “butterfly” modes upon electronic excitation shows that perylene slightly deviates from planarity in its ground state and is more rigid in the excited singlet state. For a number of levels in the Evib = 800–1600 cm?1 range, the fluorescence is composed of the resonant emission and of non-resonant (“‘relaxed’”) bands. It is shown that apparently single bands in the fluorescence-excitation spectrum correspond to ovelapping bands pumping different molecular eigenstates resulting from the intrastate coupling. The relative role of the anharmonicity and of the Coriolis interaction are discussed. The data are treated in terms of a selective coupling between doorway and hallway states with the coupling constant rapidly decreasing with the difference in the overall vibrational quantum number between initial and final state.  相似文献   

12.
The experimental activation energies (E *) of dehydration of Cu(NH3)4(H2O)SO4, Cu(en)2(H2O)X2 (X=Cl?, Br?), Cu(en)(H2O)2SO4, Cu(py)2(H2O)2SO4, CuCl2 · 2H2O and M 2 I CuCl4 · 2H2O (M I =NH4, K, Rb) were obtained from their non-isothermal thermogravimetric curves using the Coats-Redfern method. TheseE * values were compared with known data on the structures of the Cu(II) coordination polyhedra in the above complexes. No dependence of theE * values was found on either the central atom — released ligand bond length, or the number and lengths of the hydrogen bonds formed by the released water molecules. However, it was found that it is justified to seek some relationship between theE * values and the anisotropic temperature factors of the donor atoms of the ligands split off.  相似文献   

13.
The chemiluminescence produced by the Ba + Cl2 reaction was recorded as a function of He and N2 pressure. A modified Stern-Volmer treatment of competitive electronic quenching of BaCl* and BaCl*2 emission yielded upper limits to the half pressures p12(He) ? 9.0 ± 3 mtorr and p12 (N2) ? 1.1 ± 0.2 mtorr for quenching of BaCl*2 by helium and nitrogen, respectively. A lower limit of the BaCl*2 radiative lifetime is placed at τR ? 100 μ.  相似文献   

14.
EPR measurements between 98 and 298 K on single crystal of Ni(ClO4)2·6H2O have indicated the appearance of a rhombic component in the axial crystal field at Ni2+ sites at Tc = 224 K, confirming a phase transition first reported by Chaudhuri from magnetic susceptibility measurements. Temperature variations of g, D and E parameters were determined. IR spectra at room and liquid nitrogen temperatures are consistent with our EPR results.  相似文献   

15.
The chemiluminescent spectra of C*2, d 3Πg-a 3Πu, Δv = O sequence from the reaction Na + CCl4 have been obtained. The C*2, d 3Πg,v' = 6 level is formed preferentially. The quenching and vibrational relaxation rates of the C*2, d 3Πg state in Ar are 1.9 × 106 and 2.2 × 106 Torr?1 s?1, respectively. Na is one of the most efficient species for deactivation of C*2.  相似文献   

16.
The rate of the reaction
has been investigated at 40–65°C with [HClO4] varying from 0.04 to 0.6 M (μ = 0.6 M, NaClO4). The observed rate law has the form: -d[Cr(NH3)5(NCO)2+]/dt = kobs[Cr(NH3)5(NCO)2+] where kobs = a[H+]2{1 + b[H+]2} and ?1 at 55.0°C, a = 0.36 M?1 s?2 and b = 6.9 × 10?3 M?1 s?1. The rate of loss of Cr(NH3)5(NCO)2+ increases with increasing acidity to a limiting value (at [H+] ~ 0.5 M) but the yield of Cr(NH3)63+ decreases with increasing [H+] and increases with increasing temperature. In the kinetic studies the maximum yield of Cr(NH3)63+ was 35% but a synthetic procedure has been developed to give a 60% yield.  相似文献   

17.
Electrical resistivity of some binary compounds Cr2X3 and Cr3X4 (X = S, Se or Te) is studied on polycrystalline samples with the four point probe method, at temperatures between 4.2 and about 330 K. A metallic behavior is observed on Cr2Te3, Cr3Te4, Cr3S4 and the 3c′ form of Cr2Se3. Some other compounds are semiconductors: Cr3Se4 (E300K ≈ 0.07 eV; E4.2K = 2.07 × 10?4 eV), the 2c′ form of Cr2+εSe3 (E300K ≈ 0.074 eV; E4.2K = 2.76 × 10?4 eV) and the 3c′ form of Cr2S3 (E275K ≈ 0.55 eV). The observed results seem to be closely related to the nature of the octahedral neighborhood of the cations.  相似文献   

18.
The 57Fe Mössbauer effect in the solid compounds [Fe(P4)X]BPh4 (X = Br, I; P4 = hexaphenyl-1,4,7,10-tetraphosphadecane) has been studied between 4.2 and 298 K. A single temperature dependent doublet is observed. For the iodo complex, ΔEQ = 2.25 mm s?1, δIS = +0.13 mm s?1, Vtzz 0, η ≈ 0.8 at 4.2 K, whereas ΔEQ = 1.81 mm s?1, δIS = +0.20 mm s?1 at 298 K. The results are consistent with a singlet-triplet transition with k > 109 s?1.  相似文献   

19.
20.
Ab initio ΔESCF calculations of the low-lying n-π* and π-π* transitions in pyrazines are reported with D2h and C2v symmetry adapted molecular orbitals. The use of broken orbital symmetry is essential for interpreting the emission properties of pyrazine at the computational level used. The energy of the Nis orbital is calculated using these two symmetry constraints with C2v orbitals leading to results in better agreement with experiment.  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