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 15 毫秒
1.
A beam of state-selected NO molecules (J = Ω = 32) has been produced by an electrostatic hexapole and has been collided with O3 molecules in a scattering chamber. The E-field dependence of the chemiluminescent cross section, σhr, has been investigated and resulted in the determination of the M-dependence of σhr: σhr (M)/σ0 = 1.192±0.009, 0.0848±0.015, 1.177±0.015, 0.783±0.009 for M = 32, 12, ?12 and ?32, respectively. Application of the Legendre expansion technique and the density matrix formalism provided a deconvoluted σhr(γ), for a single angle of attack γ of the NO axis, expressed in simple model functions with adjustable parameters. From this analysis it is concluded that chemiluminescence only occurs when cos γ ≈ 1, the “end-on-head” orientation of NO yielding ≈ 30% of all collected light, and when cos γ ≈ ?0.275, the “broad-side-tail” orientation of NO yielding the remaining 70%. The steric factors belonging to these reactive orientations have been estimated and are S1 = 0.25±0.07 and S2 = 0.40±0.09, respectively. The observed dependence of σhr has been confronted with the rules of Woodward and Hoffman. Although there are indeed two symmetries (bpl and cpl) correlating the electron orbitals of the reactants and the products, these rules do not lead to an explanation of the steric effects of the NO+O3 reaction.  相似文献   

2.
The rate constant for the reaction of ozone with nitrogen dioxide has been measured over the temperature range 259 to 362°K, using a stopped-flow system coupled to a beam sampling mass spectrometer. A fit of the data to the Arrhenius equation gave: k = (9.44 ± 2.46) × 1010 exp[(?2509 ± 76)/T] cm3 mol?1 sec?1.  相似文献   

3.
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   

4.
The decay of NH2 radicals, from 193 nm photolysis of NH3, was monitored by 597.7 nm laser-induced fluorescence. Room-temperature rate constants of (1.21 ± 0.14) × 10?10, (1.81 ± 0.12) × 10?11, and (2.11 ± 0.18) × 10?11 cm3 molecule?1 s?1 were obtained for the reactions of NH2 with N, NO and NO2, respectively. The production of NH in the reaction of NH2 with N was observed by laser-induced fluorescence at 336.1 nm.  相似文献   

5.
CS radicals have been produced by photodissociation of CS2 at 193 nm and their disappearance monitored by LIF. The vibrationally excited CS radicals rapidly relax to CS(ν = 0). At 298 K, the rate coefficients for CS(ν = 0) reactions with O2, O3 and NO2 are (2.9 ± 0.4) × 10?19, (3.0 ± 0.4) × 10?16 and (7.6 ± 1.1) × 10?17 cm3 molecule?1 s?1 respectively. The quenching of CS(A 1II)ν=0 by He has a rate coefficient of (1.3 ± 0.2) × 10?12 cm3 molecule?1 s?1.  相似文献   

6.
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   

7.
HF laser emission was observed in the flash photolysis (λ ? 165 nm) of mixtures of O2 and CHFCl2. A total of 15 transitions ranging from Δυ = 3 → 2 to 1 → o were identified. The laser intensity was found to increase linearly with flash energy. The effects of temperature, reactant concentration and buffer gas pressure have been examined. The stimulated emission is concluded to result primary from the following reactions:
No CO2 laser emission was detected. A simple gain calculation revealed that only a small fraction of reaction energy (Eint ≈ 180 kcal/mole) was channeled into the HF product. Since the addition of D2 generated weak DF laser emission (Δυ = 3 → 2 to 1 → o) at only a minor expense (< 10%) to the HF emission, F atoms are believed to be formed as one of the minor decomposition products of FCOOH.  相似文献   

8.
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   

9.
Two different types of emission from excited NO2 were observed using pulsed ruby laser light at 6943 Å. The first type of fluorescence was seen in the near-IR and results from the single photon excitation of NO2 from the ground (2A1) state. By observing the emission as a function of time an unexpected behavior was observed in the near IR and could be explained by a consecutive deactivation mechanism, wherein a secondary species is preferentially detected. A second type of emission recently observed in the blue spectral region is weaker and is due to a multiphoton process. The intensity of the blue emission is a function of the cube of the laser intensity at low pressures and approaches the square at high pressures. We attribute this variation to simultaneous deactivation of the NO2* intermediate by collision (square) and by anti-Stokes Raman scattering off of the NO2* (cube).  相似文献   

10.
During the reduction of NO2 by C3H6 in O2 over alumina-supported Au, Rh and Pt it was found that three parallel reactions take place,i.e., reduction of NO2 to N2 and N2O, partial decomposition of NO2 to NO and oxidation of C3H6 to CO and CO2. In the absence of C3H6, the NO2→NO+O2 reaction reaches a fast equilibrium on Rh and Pt but not on Au and γ-Al2O3. Addition of C3H6 to the NO2+O2 mixture leads to the formation of NO above equilibrium conversion levels.  相似文献   

11.
The absolute rate constant of the reaction of NH2 with NO2 has been measured using a flash-photolysis laser resonance-fluorescence technique. The value obtained at room temperature is k1 = 2.3 (± 0.2) × 10?11 cm3 molecule ?1 s?1. A negative temperature coefficient has been found between 298 and 505 K for this reaction, k1 = 3.8 × 10?8 × T?1.30 cm3 molecule?1 s?1. It is thought that this is the major reaction of NH2 in the troposphere.  相似文献   

12.
A Bayard-Alpert (BA) gauge was used to determine apparent relative sensitivites Srel,X for O2, N2O, NO, NO2, NH3, CClF3 and CH3OH from gauge calibration measurements in the range 1.3×10–1 Pap1.3·10–3Pa. Nitrogen was used as a calibration standard.  相似文献   

13.
In this article, we report our detailed mechanistic study on the reactions of cyclic-N3 with NO, NO2 at the G3B3//B3LYP/6-311+G(d) and CCSD(T)/aug-cc-pVTZ//QCISD/6-311+G(d)+ZPVE levels; the reactions of cyclic-N3 with Cl2 was studied at the G3B3//B3LYP/6-311+G(d) and CCSD(T)/aug-cc-pVTZ//QCISD/6-31+G(d)+ZPVE levels. Both of the singlet and triplet potential-energy surfaces (PESs) of cyclic-N3 + NO, cyclic-N3 + NO2 and the PES of cyclic-N3 + Cl2 have been depicted. The results indicate that on singlet PESs cyclic-N3 can undergo the barrierless addition–elimination mechanism with NO and NO2 forming the respective dominant products N2 + 1cyclic-NON and 1NNO(O) + N2. Yet the two reactions on triplet PESs are much less likely to take place under room temperature due to the high barriers. For the cyclic-N3 + Cl2 reaction, a Cl-abstraction mechanism was revealed that results in the product cyclic-N3Cl + Cl with an overall barrier as high as 14.7 kcal/mol at CCSD(T)/aug-cc-pVTZ//QCISD/6-31+G(d)+ZPVE level. So the cyclic-N3 radical could be stable against Cl2 at low temperatures in gas phase. The present results can be useful for future experimental investigation on the title reactions.  相似文献   

14.
High-resolution spectra of the NO2 continuum emission produced from the reaction NO + O3 → NO2 + O2 have been investigated to detect any possible emission from O2(1Δg) at 1270 nm or O2(1Σ+g) at 762 nm. The photolysis of O3/O2 mixtures at 253.7 nm, which produces both states of O2 with known quantum efficiency, has been used as an internal standard. From the results it is concluded that less than 1/300 and 1/200 of the NO + O3 reactive collissions result in production of O2(1Δg) or O2(1Σ+g), respectively, at room temperature.  相似文献   

15.
Studies are made of the visible chemiluminescence resulting from the reaction of an atomic beam of samarium or europium with O3, N2O, NO2 and F2 under single-collision conditions (~10?4 torr). The spectra obtained for SmO, EuO, SmF, and EuF are considerably more extensive than previously observed. The variation of the chemiluminescent intensity with metal flux and with oxidant flux is investigated, and it's concluded that the reactions are bimolecular. From the short wavelength curoff of the chemiluminescent spectra, the following lower bounds to the ground state dissociation energies are obtained: D00(SmO) > 135.5 +- 0.7 kcal/mole, D00(EuO) > 131.4 ± 0.7 kcal/mole, D00(SmF) > 123.6 ± 2.1 kcal/mole, and D00(EuF) > 129.6 ± 2.1 kcal/mole. Using the Clausius-Clapeyron equation, the latent heats of sublimation are found to be ΔH1052 (Eu) = 42.3 ± 0.7 kcal/mole for europium and ΔH1084(Sm) = 47.9 ± 0.7 kcal/mole for samarium. Total phenomena- logical cross sections are determined for metal atom removal. Relative photon yields per product molecule are calculated from the integrated chemiluminescent spectra and it is found that Sm + F2 → SmF* + F is the brightest reaction. The comparison of the photon yields under single-collision conditions with those at several torr shows that energy transfer collisons play an important role in the mechanism for chemiluminescence at the higher pressures. A simple model is presented which explains the larger photon yields of the Sm reactions compared to the Eu reactions in terms of the greater number of electronic states correlating with the reactants in the case of samarium.  相似文献   

16.
A laser pulse-and-probe method has been used to determine the nascent vibrational populations in NO(v=0–4) and O2(v=6–11) formed in the thermal reaction: O(3P) + NO2 → O2(v) + NO(v). A frequency-tripled Nd: YAG laser is used to photolyse NO2, diluted tenfold in Ar, and laser-induced fluorescence spectroscopy in the NO A 2Σ+-X 2Π and O2 B 3Σu -X 3Σg electronic band system is used both to follow the kinetics of individual vibrational states and to determine the nascent vibrational distributions. The majority of the NO product is formed in v = 0 and the average vibrational yield is ≈ 4.6%. The O2 populations fall monotonically from v = 6 to 11 in a distribution close to what is expected on prior grounds. Based on a surprisal analysis, the average vibrational energy yield in O2 is ≈ 26%. The nature of the reaction dynamics is discussed.  相似文献   

17.
An oscillating time profile was observed in the visible emission of NO*2 produced by an infrared photosensitized reaction in NO2 + SF6. The origin of the modulation was found to be the periodical heating of NO2 due to the sound wave generated by the heat released from SF6 upon infrared multiphoton excitation.  相似文献   

18.
19.
The rate coefficient for the reaction of CF3O2 with NO has been measured at 295 K in helium using a flow tube sampled by a mass spectrometer. The value obtained for this rate coefficient was (17.8 ± 3.6) × 10?12 cm?3 s?1 and found to be independent of [He] over the range (6.3 ? 16.8) × 1016 cm?3. This value is approximately a factor of 2 higher than earlier measurements of the rate coefficients for CH3O2 and C2H5O2 with NO and indicates that further measurements are required for this important class of reactions.  相似文献   

20.
An experimental study on the conversion of NO in the NO/N2, NO/O2/N2, NO/C2H4/N2 and NO/C2H4/O2/N2 systems has been carried out using dielectric barrier discharge (DBD) plasmas at atmospheric pressure. In the NO/N2 system, NO decomposition to N2 and O2 is the dominating reaction; NO conversion to NO2 is less significant. O2 produced from NO decomposition was detected by an on-line mass spectrometer. With the increase of NO initial concentration, the concentration of O2 produced decreases at 298 K, but slightly increases at 523 K. In the NO/O2/N2 system, NO is mainly oxidized to NO2, but NO conversion becomes very low at 523 K and over 1.6% of O2. In the NO/C2H4/N2 system, NO is reduced to N2 with about the same NO conversion as that in the NO/N2 system but without NO2 formation. In the NO/C2H4/O2/N2 system, the oxidation of NO to NO2 is dramatically promoted. At 523 K, with the increase of the energy density, NO conversion increases rapidly first, and then almost stabilizes at 93–91% of NO conversion with 61–55% of NO2 selectivity in the energy density range of 317–550 J L−1. It finally decreases gradually at high energy density. A negligible amount of N2O is formed in the above four systems. Of the four systems studied, NO conversion and NO2 selectivity of the NO/C2H4/O2/N2 system are the highest, and NO/O2/C2H4/N2 system has the lowest electrical energy consumption per NO molecule converted.  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