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.
Infrared fluorescence has been observed from the ν1, ν6, 2ν9, ν8 and ν4 levels of CH2F2 following excitation by a 9.6 μ Q-switch CO2 laser. All the observed states exhibit a single exponential decay rate of approximately 44 msec?1 torr?1. The rare gas dependence of this rate has also been measured and found to be up to 20 times slower than the rate for the pure gas. Measurements of the risetimes of the observed fluorescence signals yielded an upper limit of 5 μsec at 1 torr for the ν1, ν6 and ν8 levels. The 2ν9 and ν4 risetimes were effectively instantaneous under the experimental conditions that prevailed. The relative magnitudes of the measured rate are discussed in terms of existing V-T/R theories and collisional energy transfer processes.  相似文献   

2.
Previous works have reported vibration-vibration and vibration-translation transfer rates in CH3F and CH3FX mixtures. In this letter we report the study of the fast VV transfer rate populating the 3ν3, ν1 and ν4 states of CH3F. Gaseous CH3F was initially excited to the ν3 state by a TEA CO2 laser operating on the P(20 9.6 μ line and collisional pumping to the 3ν3, ν1 and ν4 states was measured by monitoring the rise time of the fluorescence at 300 cm−1. The rate constant was found to be 2.2 × 105 sec−1 torr−1.  相似文献   

3.
A liquid-nitrogen-cooled, active-N2-pumped CO laser has produced cw oscillation from ν = 2 → ν = 1 (4.8 μm) to ν = 35 → ν = 34 (8.0 μm). An analysis of the output spectrum gives the VT relaxation rate of CO (ν ≈30) by He at 150°K as 2 × 103 sec−1 torr−1.  相似文献   

4.
A pulsed CO2 laser was used to irradiate a rapidly flowing mixture of NO, O3, and SF6. When the laser was tuned to an SF6 absorption line, an increase in the visible NO*2 emission was observed. The laser-induced signal has two unusual features. First, the rise time is much longer than is observed when O3 is excited directly, and, second, the signal decays to a value above the original baseline. The rise rate is attributed to VV energy transfer from SF26 to O3, while the baseline shift is attributed to a temperature jump resulting from rapid non-resonant VV relaxation within the SF6 molecule. Both the size of the T-jump and the fraction of vibrationally excited ozone molecules vary inversely with NO pressure.  相似文献   

5.
Quenching of O(1D2) by COF2 has been investigated by time-resolved resonance fluorescence monitoring of the product O(3PJ) following 248 nm pulsed laser photolysis of O3. The rate constant for total removal of O(1D2) by COF2 is (7.4 ± 1.2) × 10?11 cm3 molecule?1 s?1. 71 ± 7% of the quenching interactions result in formation of O(3PJ).  相似文献   

6.
CW dye laser induced fluorescence emission and thermal emission spectra of YO-molecules in a 1 atm H2O2Ar flame of 2430 K were recorded simultaneously. Narrow band laser excitation was applied to four rotational lines in the (1, 1) Q-branch of the A2Π32X2Σ+ transition and broadband excitation was applied to several separate Q-branches of the A2Π12,32X2Σ+ transitions. From the differences between the fluorescence emission spectra and thermal emission spectra, we conclude that collisional de-excitation of an excited vibronic level takes place by vibrational relaxation, decay to the electronic ground state and by intermultiplet transfer in order of increasing probability.  相似文献   

7.
Inefficient vibrational energy exchange between the lowest vibrational mode and the higher lying vibrational modes of CH2Cl2, CD2Cl2, CH2ClBr and CH2Br2 was investigated by ultrasonic absorption experiments. Breathing sphere theory is used to interpret the data available for VV and VR, T transfer in methylene halides.  相似文献   

8.
Ba2V2O7 is triclinic with a = 13.571(3), b = 7.320(2), c = 7.306(2) Å, α = 90.09(1), β = 99.48(1), β = 99.48(1), γ = 87.32(1)°, V = 7.15.1 Å3, Z = 4, and space group P1. The crystal structure was solved by Patterson and Fourier methods and refined by full-matrix least-squares analysis to a Rw of 0.034 (R = 0.034) using 2484 reflections measured on a Syntex P1 automatic four-circle diffractometer. The structure has two unique divanadate groups that are repeated by the b and c lattice translations to form sheets of divanadate groups parallel to (100). These sheets are linked by four unique Ba atoms that lie between these sheets. Ba(1) and Ba(3) are coordinated by eight oxygens arranged in a distorted biaugmented triangular prism and a distorted cubic antiprism, respectively. Ba(2) is coordinated by 10 oxygens arranged in a distorted gyroelongated square dipyramid and Ba(4) is coordinated by nine oxygens arranged in a distorted triaugmented triangular prism. These coordination numbers are substantiated by a bond strength analysis of the structure, and the variation in 〈BaO〉 distances is compatible with the assigned cation and anion coordination numbers. Both divanadate groups are in the eclipsed configuraton with 〈VO(br)〉 bond lengths of 1.821(4) and 1.824(4) Å and VO(br)V angles of 125.6(3) and 123.7(3)°, respectively. Examination of the divanadate groups in a series of structures allows certain generalizations to be made. Longer 〈VO(br)〉 bond lengths are generally associated with smaller VO(br)V angles. When VO(br)V < 140°, the divanadate group is generally in an eclipsed configuration; when VO(br)V > 140°, the divanadate group is generally in a staggered configuration. Nontetrahedral cations with large coordination numbers require more oxygens with which to bond, and hence O(br) is more likely to be three coordinate, with the divanadate group in the eclipsed configuration. In the eclipsed configuration, decrease in VO(br)V promotes bonding between O(br) and nontetrahedral cations, and hence smaller nontetrahedral cations are generally associated with smaller VO(br)V angles.  相似文献   

9.
The multiphoton ionization (MPI) spectrum of toluene arising from the 1B2 (1Lb) valence state has been investigated. The state participates as a two-photon resonance. A total of nine excited state fundamentals have been characterized, including three non-totally symmetric vibrations. The toluene MPI spectrum shows a strong resemblance to the two-photon fluorescence excitation spectrum with the strongest transitions taking place to the origin and excited state modes ν1(a1), ν12(a1) and ν14(b)2). The intensities of the observed fundamentals are rationalized in terms of Franck-Condon and vibronic coupling effects. A major conclusion is, that the primary mechanism for the activity of ν12 is vibronic coupling.  相似文献   

10.
Two anomalous emission bands in the fluorescence spectrum of 3,4-benzpyrene, dissolved in 2-methylpentane, have been studied as a function of temperature. These emissions originate from the second excited singlet state S2, and from a vibrationally excited S1 (S*1) respectively. From the temperature dependence of the relative yield and the decay time of the S*1 emission it can be concluded that the vibrational relaxation of this state is hampered. The rate constant for this relaxation process is smaller that 4 > 62;x 107 sec?1.  相似文献   

11.
The analysis is presented for the frequencies of stretching modes ν(GeH) in the IR spectra of organogermanium compounds R2XGeH, RX2GeH, RXYGeH, X2YGeH and XYGeH2 (where R is a substituent which does not make a dπpπ bond with germanium, and X and Y are groups capable of dπpπ interaction with germanium). It is shown that ν(GeH) in these compounds is dependent on both the I effect of R, X and Y, and the dπpπ interaction in GeX and GeY bonds. If only one substituent capable of dπpπ interaction with germanium is present, the value of such an effect is determined by itsσoRconstant. However, when germanium is bound to several substituents capable of dπpπ interaction its magnitude depends on the effective charge at germanium which is determined by the inductive and mesomeric effects of X and Y. The data obtained are compared to the dependences observed in the IR spectra of similar organosilicon compounds.  相似文献   

12.
The vibrational widths of the ν1 and ν3 Raman bands of N2O were determined at pressures ranging from 8 bar to 2 kbar and temperatures varied from 25 to 150°C. The different dephasing theories including motional narrowing collisional models and resonant vibrational energy transfer theory were tested. A comparison of the theoretical predictions with the experimental data indicates the resonance VV transfer represents the dominant broadening mechanism. The observed frequency shifts between isotropic and anisotropic components of the bands were interpreted in terms of dipole-dipole interactions in dense N2O.  相似文献   

13.
Vibrational relaxation times in SO2, SO3Ar (11%, 20% and 54% SO2) and SO2He (9.5% SO2) were measured behind incident shock waves using the laser schlieren technique in the temperature ranges 550–1200 K, 700–2100 K and 700–1600 K respectively for the three systems. An analysis of the density gradient signals revealed a double exponential behaviour consistent with earlier studies. The fast relaxation rates were not quantitatively studied and the slow relaxation rates were found to fit the following equations:
where Pτ is the relaxation time in atm μs and T is in °K. The collision numbers corresponding to the slower rates were found to agree well with a recent theoretical calculation using SSH-Tanczos theory.  相似文献   

14.
Tetragonal PbSnF4 was prepared by precipitation method with Pb(NO3)2 and SnF2 aqueous solutions. The product was characterized using X-ray diffraction (XRD), X-ray fluorescence spectroscopy (XFS), and the other chemical analyses. Tetragonal PbSnF4 exhibited the highest electric conductivity of 3.2 Sm−1 at 473 K in air as a fluoride ion conductor. We have investigated the possibility of COF2 formation using CO2 and F2 in an electrochemical cell with PbSnF4 as a solid electrolyte. At same time, we tried to produce an electric power from an electrochemical cell. This CO2/F2 electrochemical cell was constructed with a tetragonal PbSnF4 disk having Au electrodes. The electromotive force was about 0.9 V at room temperature for 0.1 MPa CO2/(0.01 MPa F2 + 0.09 MPa Ar). However, the short circuit current density was 0.24 A m−2, which was quite small. This current density was so small that no fluorocarbon compound was detected after 3 h discharge using FT-IR.  相似文献   

15.
A fluorescence excitation spectrum of (CH3)2CHO (isopropoxy radical) is reported following photolysis of isopropyl nitrite at 355 nm. Rate constants for the reaction of isopropoxy with NO, NO2, and O2 have been measured as a function of pressure (1–50 Torr) and temperature (25–110°C) by monitoring isopropoxy radical concentrations using laser-induced fluorescence. We have obtained the following Arrhenius expressions for the reaction of isopropoxy with NO and O2 respectively: (1.22±0.28)×10?11 exp[(+0.62±0.14 kcal)/RT]cm2/s and (1.51±0.70)×10?14 exp[(?0.39±0.28)kcal/RT]cm3/s where the uncertainties represent 2σ. The results with NO2 are more complex, but indicate that reaction with NO2 proceeds more rapidly than with NO contrary to previous reports. The pressure dependence of the thermal decomposition of the isopropoxy radical was studied at 104 and 133°C over a 300 Torr range using nitrogen as a buffer gas. The reaction is in the fall-off region over the entire range. Upper limits for the reaction of isopropoxy with acetaldehyde, isobutane, ethylene, and trimethyl ethylene are reported.We have performed the first LIF study of the isopropoxy radical. Arrhenius parameters were measured for the reaction of i-PrO with O2, NO, NO2, using direct radical measurement techniques. All reactions are in their high-pressure limits at a few Torr of pressure. The rate constant for the reactions of i-PrO with NO and NO2 reactions exhibit a small negative activation energy. Studies of the i-PrO + NO2 reaction produce data which indicate that O(3P) reacts rapidly with i-PrO. Unimolecular decomposition studies of i-PrO indicate that the reaction is in the fall-off region between 1 and 300 Torr of N2 and the high-pressure limit is above 1 atmosphere of N2.  相似文献   

16.
The luminescence of Li2ZrO3Ti is described. The emission and excitation spectra are reported. The temperature dependence of the decay time is discussed and points to a complicated energy-level scheme. Further, the decay time of the luminescence of ZrO2Ti has been measured. It appears to be very long (~1 msec).  相似文献   

17.
Phase relations in the system NiAl2O4Ni2SiO4 were studied in the pressure range 1.5 ~ 13.0 GPa and in the temperature range 800 ~ 1450°C. Two new phases, IV and V, were found in regions of pressure higher than 4 GPa. Phase V disproportionates into a mixture of Ni2SiO4-spinel, NiO, and Al2O3 at approximately 9.5 GPa and 1100°C. Phases III, IV, and V form a solid solution in some compositional range: phases IV and V have a composition around NiAl2O4·Ni2SiO4, whereas phase III spreads from NiAl2O4·Ni2SiO4 to the NiAl2O4-rich side. All the phases I ~ V are structurally considered to be spinel derivatives, “spinelloids,” with three kinds of tetrahedral groups; isolated tetrahedra TO4, linked ones T2O7, and triply linked ones T3O10. The ratios of isolated tetrahedra to linked ones are large in the higher-pressure phases and small in the lower-pressure phases. The difference of compositional range of phase III from that of phases IV and V is possibly explained by the avoidance of linked tetrahedra such as O3AlOAlO3.  相似文献   

18.
Dissociation rates of SO2 in SO2 + Ar mixtures at 6%, 11%, 15% and 20% of SO2 were measured behind incident shock waves over a temperature range 4000–6000 K at initial pressures 1.0 to 2.5 Torr. The recorded laser schlieren signals exhibited two exponentials, the faster one due to vibrational relaxation and the slower one due to dissociation. The initial dissociation rate was calculated from the value of the density gradient at the point of intersection of the two exponentials. A least-squares analysis of the experimental data yielded the following empirical relations: kSO2Ar = 3.34 × 1015 exp(?107.6 kcal mole?1/RT) cm3/mole s, kSO2SO2 = 5.02 × 1014 exp(?66.6 kcal mole?1 kcal mole?1/RT) cm3/mole s.  相似文献   

19.
The multiple-photon dissociation of N2H4 and CH3NH2 by pulsed CO2 laser light to produce NH2(X?2BI has been studied using the laser-induced fluorescence detection method. The relative NH2 yield, represented by the fluorescence signal, has been measured as a function of the fluence from the threshold at about 0.1 J/cm2 to about 100 J/cm2, at different CO2-laser lines and at pressures down to 10?4 Torr.  相似文献   

20.
An electron diffraction analysis of the molecular structures of 1,1,1,3,3,3-hexachloro-1,3-disilapropane and octachloro-1,3-disilapropane has been carried out. Deviations from the staggered conformation are indicated. The data may be approximated by models with C2 symmetry and a small tilt of the SiCl3 groups. The main bond lengths (rg) and bond angles obtained for (SiCl3)2 CH2 are: SiCl, 202.7(4); SiC, 186.6(6); CH, 109.8(24) pm, ClSiCl, 107.9(1); SiCSi, 118.3(7)°; and for (SiCl3)2CCl2: SiCl, 202.0(4); SiC, 190.2(9); CCl, 179.6(9) pm; ClSiCl, 109.5(1); SiCSi, 120.6(9); ClCCl, 110.9(16); SiCCl, 106.3(3)°.  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