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.
The generation of metastable O2(1Σg+) and O2(1Δg) in the H + O2 system of reactions was studied by the flow discharge chemiluminescence detection method. In addition to the O2(1Σg+) and O2(1Δg) emissions, strong OH(v = 2) → OH(v = 0), OH(v = 3) → OH(v = 1), HO2(2A000) → HO2(2A000), HO2(2A001) → HO2(2A000), and H O2(2A200) → HO2(2A000) emissions were detected in the H + O2 system. The rate constants for the quenching of O2(1Σg+) by H and H2 were determined to be (5.1 ± 1.4) × 10?13 and (7.1 ± 0.1) × 10?13 cm3 s?1, respectively. An upper limit for the branching ratio to produce O2(1Σg+) by the H + HO2 reaction was calculated to be 2.1%. The contributions from other reactions producing singlet oxygen were investigated.  相似文献   

2.
The reaction of O2(1Δg) with HO2(X?) was studied in an isothermal flow reactor in the pressure range 7?p? 10.7 mbar at temperatures between 299?T? 423 K. H-atom production was observed in the reaction O2(1Δg) + HO22A′) - H(2S)+ 2O2 (3Σg?). The rate of this reaction (k1) is estimated to be k1 = (1 ± 0.5) × 1014 CM3 Mol?1 s?1. The implications of this reaction to recent determinations of the rate of the reaction H + O2(1Δg) are discussed.  相似文献   

3.
Measurements of the quantum yield of self-sensitized 1,3-diphenylisobenzofuran peroxidation as a function of dissolved oxygen of added azulene concentrations indicate that oxygen quenching of the sensitizer singlet state produces both triplet and ground states of the sensitizer in addition to O2(1Δg) and O2(3Σ?g). This partitioning of quenching products may be due to the competitive relaxation of the initially formed complex (oxciplex), or to sequential relaxation of different oxciplex states in which symmetry and spin barriers are negotiated by complex dissociation and re-encounter of the solute pair in the required configuration. The latter interpretation provides re-encounter probabilities for the processes M(T1) + O2(1Δg) → M(T1) + O2(3Σ?g) and M(T1) + O2(3Σ?g) → M(So) + O2(1Δg) from which estimated rate constants are compatible with theoretical expectation.  相似文献   

4.
A laser flash photolysis-resonance fluorescence technique has been employed to study the kinetics of the reaction of O(3P) with CF3NO (k2) as a function of temperature. Our results are described by the Arrhenius expression k2(T) = (4.54 ± 0.70) × 10?12 exp[(?560± 46)/T] cm3molecule?1 s?1 (243 K ? T ? 424 K); errors are 2σ and represent precision only. The O(3P) + CF3NO reaction is sufficiently rapid that CF3NO cannot be employed as a selective quencher for O2(a1Δg) in laboratory systems where O(3P) and O2(a1Δg) coexist, and where O(3P) kinetics are being investigated. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
Strongly enhanced N2 first positive emission N2(B 3Πg → A 3Σ+u) has been observed on addition of N atoms into a flowing mixture of Cl and HN3. The dependence of the emission intensity on N atom concentration gave a rate constant for the reaction N + N3 → N2(B 3Πg) + N2(X 1Σ+g) of i(1.6 ± 1.1) × 10?11 cm3 molecule?1 s?1. That for the reaction Cl + HN3 → HCl + N3 is (8.9 ± 1.0) × 10?13 cm3 molecule?1 s?1 from the decay of the emission. Comparison of the emission intensity in ClHN3 with that in ClHN3N gave the rate constant of the reaction N3 + N3 → N2(B 3Πg) + 2N2(X 1Σ+g) as 1.4 × 10?12 cm3 molecule?1 s?1 on the assumption that N + N3 yields only N2(B 3Πg) + N2(X 1Σ+g).  相似文献   

6.
The reactions of electrically dicharged nitrogen and hydrogen with O2(1Δg) is probabnly slower than with ground state O2. ON the other hand, the reaction of H-atoms with O2(1Δg) was found to occur with a rate constant k=(2,5±0.5)× 10?14 cm3 molecule?1 sec?1, although it was not posible to establish whether the reaction produced OH radicals or simply represented physical quenching.  相似文献   

7.
A laser flash photolysis-resonance fluorescence technique has been employed to study the kinetics of the important stratospheric reactions Cl(2PJ) + O3 → ClO + O2 and Br(2P3/2) + O3 → BrO + O2 as a function of temperature. The temperature dependence observed for the Cl(2PJ) + O3 reaction is nonArrhenius, but can be adequately described by the following two Arrhenius expressions (units are cm3 molecule?1 s?1, errors are 2σ and represent precision only): ??1(T) = (1.19 ± 0.21) × 10?11 exp [(?33 ± 37)/T] for T = 189–269K and ??1(T) = (2.49 ± 0.38) × 10?11 exp[(?233 ± 46)/T] for T = 269–385 K. At temperatures below 230 K, the rate coefficients determined in this study are faster than any reported previously. Incorporation of our values for ??1(T) into stratospheric models would increase calculated ClO levels and decrease calculated HCl levels; hence the calculated efficiency of ClOx catalyzed ozone destruction would increase. The temperature dependence observed for the (2P3/2) + O3 reaction is adequately described by the following Arrhenius expression (units are cm3 molecule?1 s?1, errors are 2σ and represent precision only): ??2(T) = (1.50 ± 0.16) × 10?1 exp[(?775 ± 30)/T] for T = 195–392 K. While not in quantitative agreement with Arrhenius parameters reported in most previous studies, our results almost exactly reproduce the average of all earlier studies and, therefore, will not affect the choice of ??2(T) for use in modeling stratospheric BrOx chemistry.  相似文献   

8.
The photodegradation of the herbicide clomazone in the presence of S2O82? or of humic substances of different origin was investigated. A value of (9.4 ± 0.4) × 108 m ?1 s?1 was measured for the bimolecular rate constant for the reaction of sulfate radicals with clomazone in flash‐photolysis experiments. Steady state photolysis of peroxydisulfate, leading to the formation of the sulfate radicals, in the presence of clomazone was shown to be an efficient photodegradation method of the herbicide. This is a relevant result regarding the in situ chemical oxidation procedures involving peroxydisulfate as the oxidant. The main reaction products are 2‐chlorobenzylalcohol and 2‐chlorobenzaldehyde. The degradation kinetics of clomazone was also studied under steady state conditions induced by photolysis of Aldrich humic acid or a vermicompost extract (VCE). The results indicate that singlet oxygen is the main species responsible for clomazone degradation. The quantum yield of O2(a1Δg) generation (λ = 400 nm) for the VCE in D2O, ΦΔ = (1.3 ± 0.1) × 10?3, was determined by measuring the O2(a1Δg) phosphorescence at 1270 nm. The value of the overall quenching constant of O2(a1Δg) by clomazone was found to be (5.7 ± 0.3) × 107 m ?1 s?1 in D2O. The bimolecular rate constant for the reaction of clomazone with singlet oxygen was kr = (5.4 ± 0.1) × 107 m ?1 s?1, which means that the quenching process is mainly reactive.  相似文献   

9.
The chemiluminescence of the Cypridina luciferin analog, 2-methyl-6-(p-methoxyphenyl)-3,7-dihydroirnidazo[1,2-a]pyrazin-3-one (MCLA), with O2 (1Δg) generated by the retro-Diels-Alder reaction of 3-(4′-methyI-l'-naphthyl)-propionic acid endoperoxide was studied in an aqueous solution with pH 7.12 at 37°C. The retro-Diels-Alder reaction occurs with a first-order rate constant of (4.16 ± 0.13) × 10?4/s to quantitatively yield O2 (1Δg) and 3-(4′-methyl-l'-naphthyl (-propionic acid. MCLA consumed equimolar amounts of O2 (1Δg) with a second-order rate constant (6.96 ± 0.27) × 108/M/s to emit light in an aqueous solution with pH 7.12 at 37°C. The chemiluminescence spectrum was identified as the fluorescence spectrum of 2-acetylamino-5-(p-methoxyphenyl)pyrazine (OMCLA), a major chemiluminescence reaction. Chemiluminescence spectra and product yields for MCLA reactions with O21Δg, with O2 (3Σ?g) and with superoxide anion radicals are identical, suggesting that all of these reactions occur via a common MCLA-2-hydrope-roxide intermediate formed by a combination of MCLA radicals and superoxide anion radicals. We have established practical use of NEPO as an O2 (1Δg) source and MCLA as a biological probe for detecting O2 (1Δg).  相似文献   

10.
The quenching rate constants of O2(1Δg) with n-butylamine, diethylamine, dipropylamine, dibutylamine, and tripropylamine have been determined in a discharge flow system. The rate constants are found to be (1.6 ± 0.2) × 103, (8.5 ± 0.6) × 104, (9.8 ± 0.5) × 104, (2.1 ± 0.1) × 105, and (8.6 ± 0.5) × 105 1 mol?1 s?1, respectively. The rate constants are found to increase in the order, tertiary amine → secondary amine → primary amine. The “inductive effect” of alkyl substitution is also found to increase the rate constant in a given series of amines.  相似文献   

11.
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   

12.
The heat of formation of benzophenone oxide, Ph2CO2, was measured using photoacoustic calorimetry. The enthalpy of the reaction Ph2CN2 + O2 → Ph2CO2 + N2 was found to be ?48.0 ±0.8 kcal mol?1 and ΔHf(Ph2CN2) was determined by measuring the reaction enthalpy for Ph2CN2 + EtOH → Ph2CHOEt + N2 (?53.6 ±1.0 kcal mol?1). Taking ΔHf(PhCHOEt) = ?10.6 kcal mol?1 led to ΔHf(Ph2CN2) = 99.2 ± 1.5 kcal mol?1 and hence to ΔHf(Ph2CO2) = 51.1 ± 2.0 kcal mol?1. The results imply that the self-reaction of benzophenone oxide i.e., 2Ph2CO2 → 2Ph2CO + O2 is exothermic by ?76.0 ±4.0 kcal mol?1.  相似文献   

13.
A new electronic systems has been observed from excited Hg vapour, which is assigned to collisionally induced emission from the Hg2 O±g first excited states of the dimer: Hg2O±g + M → 2Hg(6 1S0) + M + hvmax 3950 A). For M = N2, the rate coefficient is 5.3(±0.7) × 10?19 cm3 molecule?1 at 298 K. From time resolved measurements of the luminescence in the afterglow following pulsed excitation, the decay rate of the green emission, in an excess of N2, is shown to be a linear function of [Hg][N2]. It is concluded that the reaction which controls the decay of the excitation is formation of an excited trimer in a termolecular reaction; the trimer is the carrier of the green emission: Hg2 O±g + Hg(6 1S0 + Hg(61S0 + N2 → Hg33Πu + N2. The rate coefficient is 1.10(±0.07) × 10?30 cm6 molecule?2 s?1 at 298 K.  相似文献   

14.
Oxygen atoms are detected by NO + O + M chemiluminescence as a secondary product of the reaction between Cl and O3. The mechanism Cl + O3 → ClO + O2(1Σ+g), O2(1Σ+g) + O3 → O2 + O2 + O is proposed to account for the oxygen atom formation. The branching ratio to the O2(1Σ+g) product in the reaction of Cl with O3 is estimated to be in the range (0.1–0.5) x 10?2.  相似文献   

15.
Abstract— Tris (2,2'-bipyridyl)ruthenium(II)chloride hexahydrate (Ru[bpy]32+) free in solution and adsorbed onto antimony-doped SnO2 colloidal particles was used as a photosensitizer for a comparison of the O2(1Δg) and electron-transfer-mediated photooxidation of tryptophan (TRP), respectively. Quenching of excited Ru(bpy)32+ by O2(3σg?) in an aerated aqueous solution leads only to the formation of O2(1Δg) (φ4= 0.18) and this compound was used as a type II photosensitizer. Excitation of Ru(bpy)32+ adsorbed onto Sb/SnO2 results in a fast injection of an electron into the conduction band of the semiconductor and accordingly to the formation of Ru(bpy)32+ and was used for the sensitization of the electron-transfer-mediated photooxidation. The Ru(bpy)33+ is reduced by TRP with a bimolecular rate constant kQ= 5.9 × 108M?1 s?1, while O2(1Δg) is quenched by TRP with kt= 7.1 × 107M?1 s?1 (chemical + physical quenching). Relative rate constants for the photooxidation of TRP (kc) via both pathways were determined using fluorescence emission spectroscopy. With Np, the rate of photons absorbed, being constant for both pathways we obtained kc= (372/Np) M?1 s?1 for the O2(1Δg) pathway and kc≥ (25013/Np) M?1 s?1 for the electron-transfer pathway, respectively. Thus the photooxidation of Trp is more than two orders of magnitude more efficient when it is initiated by electron transfer than when initiated by O2(1Δg).  相似文献   

16.
The 300 K reactions of O2 with C2(X 1Σ+g), C2(a 3 Πu), C3(X? 1Σ+g) and CN(X 2Σ+), which are generated via IR multiple photon dissociation (MPD), are reported. From the spectrally resolved chemiluminescence produced via the IR MPD of C2H3CN in the presence of O2, CO molecules in the a 3Σ+, d 3Δi, and e 3Σ? states were identified, as well as CH(A 2Δ) and CN(B 2Σ+) radicals. Observation of time resolved chemiluminescence reveals that the electronically excited CO molecules are formed via the single-step reactions C2(X 1Σ+g, a 3Πu) + O2 → CO(X 1Σ+ + CO(T), where T denotes are electronically excited triplet state of CO. The rate coefficients for the removal of C2(X 1Σ+g) and C2(a 3Πu) by O2 were determined both from laser induced fluorescence of C2(X 1Σ+g) and C2(a 3Πu), and from the time resolved chemiluminescence from excited CO molecules, and are both (3.0 ± 0.2)10?12 cm3 molec?1 s?1. The rate coefficient of the reaction of C3 with O2, which was determined using the IR MPD of allene as the source of C3 molecules, is <2 × 10?14 cm3 molec?1 s?1. In addition, we find that rate coefficients for C3 reactions with N2, NO, CH4, and C3H6 are all < × 10?14 cm3 molec?1 s?1. Excited CH molecules are produced in a reaction which proceeds with a rate coefficient of (2.6 ± 0.2)10?11 cm3 molec?1 s?1. Possible reactions which may be the source of these radicals are discussed. The reaction of CN with O2 produces NCO in vibrationally excited states. Radiative lifetime of the ā 2Σ state of NCo and the ā 1Πu(000) state of C3 are reported.  相似文献   

17.
The rate constant for the reaction Cl + CHClO → HCl + CClO was determined from relative decay rates of CHClO and CH3Cl inthe photolysis of mixtures containing Cl2 (~1 torr), CH3Cl (~1 torr), and O2 (~0.1 torr) in 700 torr N2. In such mixtures CHClO was generated in situ as a principal product prior to complete consumption of O2. The value of k(Cl + CHClO)/k(Cl + CH3Cl) = 1.6 ± 0.2(3σ) combined with the literature value of k(Cl + CH3Cl) = 4.9 × 10?13 cm3/molecule sec gives k(Cl + CHClO) = 7.8 × 10?13 cm3/molecule sec at 298 ± 2 K, in excellent agreement with a previous value of (7.9 ± 1.5) × 10?13 cm3/molecule sec determined by Sanhueza and Heicklen [J. Phys. Chem., 79 , 7 (1975)]. Thus this reaction is approximately 100 times slower than the corresponding reactions of aldehydes and alkanes with comparable C? H bond energies (≤95 kcal/mol).  相似文献   

18.
We have used the single‐pulse shock tube technique with postshock GC/MS product analysis to investigate the mechanism and kinetics of the unimolecular decomposition of isopropanol, a potential biofuel, and of its reaction with H atoms at 918‐1212 K and 183‐484 kPa. Experiments employed dilute mixtures in argon of isopropanol, a radical scavenger, and, for H‐atom studies, two different thermal precursors of H. Without an added H source, isopropanol decomposes in our studies predominantly by molecular dehydration. Added H atoms significantly augment decomposition, mainly by abstraction of the tertiary and primary hydrogens, reactions that, respectively, lead to acetone and propene as stable organic products. Traces of acetaldehyde were observed in some experiments above ≈ 1100 K and establish branching limits for minor decomposition pathways. To quantitatively account for secondary chemistry and optimize rate constants of interest, we employed the method of uncertainty minimization using polynomial chaos expansions (MUM‐PCE) to carry out a unified analysis of all datasets using a chemical model–based originally on JetSurF 2.0. We find: k(isopropanol → propene + H2O) = 10(13.87 ± 0.69) exp(?(33 099 ± 979) K/ T) s?1 at 979‐1212 K and 286‐484 kPa, with a factor of two uncertainty (2σ), including systematic errors. For H atom reactions, optimization yields: k(H + isopropanol → H2 + p‐C3H6OH) = 10(6.25 ± 0.42) T2.54 exp(?(3993 ± 1028) K /T) cm3 mol?1 s?1 and k(H + isopropanol → H2 + t‐C3H6OH) = 10(5.83 ± 0.37) T2.40 exp(?(1507 ± 957) K /T) cm3 mol?1 s?1 at 918‐1142 K and 183‐323 kPa. We compare our measured rate constants with estimates used in current combustion models and discuss how hydrocarbon functionalization with an OH group affects H abstraction rates.  相似文献   

19.
The kinetics of the gas-phase reaction of Cl atoms with CF3I have been studied relative to the reaction of Cl atoms with CH4 over the temperature range 271–363 K. Using k(Cl + CH4) = 9.6 × 10?12 exp(?2680/RT) cm3 molecule?1 s?1, we derive k(Cl + CF3I) = 6.25 × 10?11 exp(?2970/RT) in which Ea has units of cal mol?1. CF3 radicals are produced from the reaction of Cl with CF3I in a yield which was indistinguishable from 100%. Other relative rate constant ratios measured at 296 K during these experiments were k(Cl + C2F5I)/k(Cl + CF3I) = 11.0 ± 0.6 and k(Cl + C2F5I)/k(Cl + C2H5Cl) = 0.49 ± 0.02. The reaction of CF3 radicals with Cl2 was studied relative to that with O2 at pressures from 4 to 700 torr of N2 diluent. By using the published absolute rate constants for k(CF3 + O2) at 1–10 torr to calibrate the pressure dependence of these relative rate constants, values of the low- and high-pressure limiting rate constants have been determined at 296 K using a Troe expression: k0(CF3 + O2) = (4.8 ± 1.2) × 10?29 cm6 molecule?2 s?1; k(CF3 + O2) = (3.95 ± 0.25) × 10?12 cm3 molecule?1 s?1; Fc = 0.46. The value of the rate constant k(CF3 + Cl2) was determined to be (3.5 ± 0.4) × 10?14 cm3 molecule?1 s?1 at 296 K. The reaction of Cl atoms with CF3I is a convenient way to prepare CF3 radicals for laboratory study. © 1995 John Wiley & Sons, Inc.  相似文献   

20.
The enthalpies of formation of 1.6-methano-[10] annulene (IV) (ΔHf298 (IV, g) = 75.2 ± 0.6 kcal mol?1), 1.6-imino-[10] annulene (V) (ΔHf298(V, g) = 87.8 ± 0.7 kcal mol?1) and of 1.6-oxido-[10] annulene (VI) (ΔHf298(VI, g) = 47.8 ± 1.2 kcal mol?1) have been determined by combustion calorimetry. The difficulties connected with an attempt to derive meaningfull «resonance energies» are discussed.  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