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.
A steady-state system involving photolysis of HONO as a source of OH was used to investigate the reaction of OH with CS2 at 1 atm and 295 K. In the presence of O2 ( > 40 Torr) a rapid reaction of OH with CS2 occurs giving OCS. At lower O2 concentrations, OCS formation ceases. In air the overall rate constant for OH + CS2 → OCS was (1.7 ± 0.9) × 10?12 cm3 molecule?1 s?1.  相似文献   

2.
The photochemistry of a potential light-activated anticancer complex, trans,trans,trans-[PtIV(py)2(N3)2(OH)2], was explored by steady-state and laser flash photolysis. The photolysis was a multistage process with the formation of complexes trans-[PtIV(py)2(N3)(OH)2(H2O)]+ and/or trans-[PtIV(py)2(N3)(OH)3] due to chain photoaquation at the first stage.  相似文献   

3.
Unsaturated 1,6‐dicarbonyls like 2,4‐hexadienedial are ring opening products in the OH initiated photo‐oxidation of aromatic hydrocarbons. In the present study, the photolysis of E,Z‐ and E,E‐2,4‐hexadienedial has been investigated under natural sunlight conditions in a large volume outdoor reaction chamber. In the case of the E,Z‐isomer, an extremely rapid isomerization into the E,E‐form was observed. The photoisomerization frequency, relative to that of NO2, was found to be J(E,Z‐2,4‐hexadienedial)/J(NO2) = (0.148 ± 0.012). A more complex photolysis behavior was observed for E,E‐2,4‐hexadienedial. Here, a fast equilibrium preceded a comparably slow photolysis. For the equilibrium reaction, relative frequencies of J(E,E‐2,4‐hexadienedial → EQUI)/J(NO2) = (0.113 ± 0.009) and J(EQUI → E,E‐2,4‐hexadienedial)/J(NO2) = (0.192 ± 0.016) were obtained, giving an equilibrium constant of K = (0.59 ± 0.07). For the photolysis frequencies, ratios of J(E,E‐2,4‐hexadienedial → products)/J(NO2) = J(EQUI → products)/J(NO2) = (1.22 ± 0.45)·10−2 were determined. Qualitative aerosol measurements during the experiments showed that the photolysis of 2,4‐hexadienedials is a source of secondary organic aerosol. In addition to the photolysis study, OH radical reaction rate constants were determined, values of (7.4 ± 1.9)·10−11 and (7.6 ± 0.8)·10−11 cm3 s−1 were obtained for E,Z‐ and E,E‐2,4‐hexadienedial, respectively. The results indicate that the dominant fate of E,Z‐2,4‐hexadienedial in the atmosphere will be photoisomerization, while for E,E‐2,4‐hexadienedial, both photolysis and OH radical reaction will be important sinks. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 689–697, 1999  相似文献   

4.
A lower limit to the OH(X2Π) vibrational excitation produced by the reaction O(1D) + H2 has been observed using a low-pressure infrared chemiluminescence apparatus. The O(1D) was generated by laser photolysis of O3. The measured OH(v') vibrational distribution is inverted; it peaks at v' = 2.  相似文献   

5.
A flash photolysis–resonance fluorescence technique was used to investigate the kinetics of the OH(X2Π) radical and O(3P) atom‐initiated reactions with CHI3 and the kinetics of the O(3P) atom‐initiated reaction with C2H5I. The reactions of the O(3P) atom with CHI3 and C2H5I were studied over the temperature range of 296 to 373 K in 14 Torr of helium, and the reaction of the OH (X2Π) radical with CHI3 was studied at T = 298 K in 186 Torr of helium. The experiments involved time‐resolved resonance fluorescence detection of OH (A2Σ+ → X2Π transition at λ = 308 nm) and of O(3P) (λ = 130.2, 130.5, and 130.6 nm) following flash photolysis of the H2O/He, H2O/CHI3/He, O3/He, and O3/C2H5I/He mixtures. A xenon vacuum UV (VUV) flash lamp (λ > 120 nm) served as a photolysis light source. The OH radicals were produced by the VUV flash photolysis of water, and the O(3P) atoms were produced by the VUV flash photolysis of ozone. Decays of OH radicals and O(3P) atoms in the presence of CHI3 and C2H5I were observed to be exponential, and the decay rates were found to be linearly dependent on the CHI3 and C2H5I concentrations. Measured rate coefficients for the reaction of O(3P) atoms with CHI3 and C2H5I are described by the following Arrhenius expressions (units are cm3 s?1): kO+C2H5I(T) = (17.2 ± 7.4) × 10?12 exp[?(190 ± 140)K/T] and kO+CHI3(T) = (1.80 ± 2.70) × 10?12 exp[?(440 ± 500)K/T]; the 298 K rate coefficient for the reaction of the OH radical with CHI3 is kOH+CHI3(298 K) = (1.65 ± 0.06) × 10?11 cm3 s?1. The listed uncertainty values of the Arrhenius parameters are 2σ‐standard errors of the calculated slopes by linear regression.  相似文献   

6.
Previous studies have shown a significant OH yield from the reaction of RCO radicals (generated from the photolysis of corresponding ketone) with oxygen below total pressures of 200 Torr. The potential of these reactions as a source of OH radicals for flash photolytic kinetic studies is investigated. The viability of the method was tested by measuring rate coefficients for the reaction of OH with ethanol using both acetone/O2 mixtures and t‐butyl hydroperoxide photolysis. The results (with statistical errors at the 2σ level) are in excellent agreement with each other (kEtOH(acetone) = (5.87 ± 0.34) × 10?18 T2 exp((515 ± 21)K/T) cm3 molecule?1 s?1 and kEtOH (t‐butyl hydroperoxide) = (5.27 ± 0.34) × 10?18 T2 exp((557 ± 20)K/T) cm3 molecule?1 s?1) and with the IUPAC recommendation. The reaction of OH with methyl ethyl ketone (2‐butanone) has also been investigated using a similar technique. The results show a strong non‐Arrhenius temperature dependence, k = (3.84 ± 0.12) × 10?24× T4 × exp((1038 ± 11)/t). The merits of the ketone/oxygen OH source are contrasted with other established precursors. A major advantage of the technique is the ability to cleanly generate OD without the potential for isotopic scrambling prior to photolysis. © 2008 Wiley Periodicals, Inc. 40: 504–514, 2008  相似文献   

7.
Absolute rate coefficients for the reaction of OH with HCl (k1) have been measured as a function of temperature over the range 240–1055 K. OH was produced by flash photolysis of H2O at λ > 165 nm, 266 nm laser photolysis of O3/H2O mixtures, or 266 nm laser photolysis of H2O2. OH was monitored by time-resolved resonance fluorescenceor pulsed laser–induced fluorescence. In many experiments the HCl concentration was measured in situ in the slow flow reactor by UV photometry. Over the temperature range 240–363 K the following Arrhenius expression is an adequate representation of the data: k1 = (2.4 ± 0.2) × 10?12 exp[?(327 ± 28)/T]cm3 molecule?1 s?1. Over the wider temperature range 240–1055 K, the temperature dependence of k1 deviates from the Arrhenius form, but is adequately described by the expression k1 = 4.5 × 10?17 T1.65 exp(112/T) cm3 molecule?1 s?1. The error in a calculated rate coefficient at any temperature is 20%.  相似文献   

8.
Rate constants for the reaction of OH radicals with OCS and CS2 have been determined at 296 K using the flash photolysis resonance fluorescence technique. The values derived from this study are kOH + OCS = (5.66 ± 1.21) × 10?14 cm3 molecule?1 s?1 and kOH + CS2 = (1.85 ± 0.34) × 10?13 cm3 molecule?1 s?1, where the uncertainties are 95% confidence limits making allowance for possible systematic errors.  相似文献   

9.
The electronic structure of the [pol-Ti(OBu)4 + H2O2] complex, which is formed during the photolysis of water over [pol-Ti(OBu)4 + CH3OH] complex on silica gel, is elucidated in detail from the photocatalytic standpoint by DV-Xα cluster calculations. The numerical results reveal the appearance of a localized state (active site) similar to that of the [pol-Ti(OBu)4 + CH3OH] complex. The [TiOBu)(OH)(OH)4]2− cluster, which substitutes a hydroxyl group for the (OOH) group, exhibits an electronic structure very close to that of the [Ti(OBu)(OMe)(OH)4]2− cluster (A1 or A2), which was studied previously. A reduction in photocatalytic activity for the water decomposition is discussed.  相似文献   

10.
Photolysis of [Ir(η2-coe)H2(TpMe2)] ( 1 ; TpMe2=hydrotris(3,5-dimethylpyrazolyl)borato, coe=(Z)-cyclooctene) in CH3OH gives a mixture of [IrH4(TpMe2)] ( 4 ) and [Ir(CO)H2(TpMe2)] ( 5 ) in a ca. 1 : 1 ratio. Mass-spectral analysis of the distillate of the reaction mixture at the end of the photolysis shows the presence of coe. When pure CD3OD is used as solvent, the deuteride complexes [IrD4(TpMe2)] ((D4)- 4 ) and [Ir(CO)D2(TpMe2)] ((D2)- 5 ) are obtained. Also the photolysis of [Ir(η4-cod)(TpMe2)] ( 3 ) (cod=cycloocta-1,5-diene) gives 4 and 5 . A key feature of this photoreaction is the intramolecular dehydrogenation of cod with formation of cycloocta-1,3,5-triene, detected by mass spectroscopy at the end of the photolysis. Labeling experiments using CD3OD show that the hydrides in 4 originate from MeOH. When 13CH3OH is used as solvent, [Ir(13CO)H2(TpMe2)] is formed demonstrating that CH3OH is the source of the CO ligand. The observation that the photolysis of both 1 and 3 give the same product mixture is attributed to the formation of a common intermediate, i.e., the coordinatively unsaturated 16e species {IrH2(TpMe2)}.  相似文献   

11.
Summary Pulsed laser photolysis with resonance fluorescence monitoring of OH radicals was applied at T = 300±2 K to obtain the rate constants of k1= (3.38±0.60)x10-12, k2= (2.52±0.44)x10-13and k3 = (1.06±0.30)x10-13cm3molecule-1s-1with 2σprecision given for the overall reactions OH + CH3CH2OH (1), OH + CF2HCH2OH (2) and OH + CF3CH2OH (3), respectively. k2is the first direct kinetic data for the reaction of OH radicals with CF2HCH2OH reported in the literature.</o:p>  相似文献   

12.
A technique for detecting transient far-infrared laser magnetic resonance (LMR) signals, induced by pulsed CO2 laser photolysis, has been developed. The method is illustrated with preliminary data from photolysis of SF6 containing H2O or NO. In the former mixture. OH(X 2Π) formation and decay is observed, while in the latter, a transient decrease in NO(X 2Π) concentration followed by a return to the original value is observed. Possible physical and chemical mechanisms are discussed.  相似文献   

13.
The synthesis, IR spectrum, and first‐principles characterization of CF3CH(ONO)CF3 as well as its use as an OH radical source in kinetic and mechanistic studies are reported. CF3CH(ONO)CF3 exists in two conformers corresponding to rotation about the RCO? NO bond. The more prevalent trans conformer accounts for the prominent IR absorption features at frequencies (cm?1) of 1766 (N?O stretch), 1302, 1210, and 1119 (C? F stretches), and 761 (O? N? O bend); the cis conformer contributes a number of distinct weaker features. CF3CH(ONO)CF3 was readily photolyzed using fluorescent blacklamps to generate CF3C(O)CF3 and, by implication, OH radicals in 100% yield. CF3CH(ONO)CF3 photolysis is a convenient source of OH radicals in the studies of the yields of CO, CO2, HCHO, and HC(O)OH products which can be difficult to measure using more conventional OH radical sources (e.g., CH3ONO photolysis). CF3CH(ONO)CF3 photolysis was used to measure k(OH + C2H4)/k(OH + C3H6) = 0.29 ± 0.01 and to establish upper limits of 16 and 6% for the molar yields of CO and HC(O)OH from the reaction of OH radicals with benzene in 700 Torr of air at 296 K. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 159–165, 2003  相似文献   

14.
Laser-induced fluorescence (LIF) detection of OH radicals is used to examine the reaction of CO2 molecules with translationally hot H atoms generated by ArF laser photolysis of HBr. Prompt formation of rotationally excited OH X 2Πi(υ″ = 0,1,2), has been observed.  相似文献   

15.
Rate constants for the radical-radical reactions N + OH → NO + H (1), and O + OH → O2 + H (2) have been measured for the first time by a direct method. In each experiment, a known concentration of N or O atoms is established in a discharge-flow system. OH radicals are then created by flash photolysis of H2O present in the flowing gas, and the disappearance of OH is monitored by time-resolved observations of its resonance fluorescence. The experiments yield K1 = (5.0 = 1.2) × 10?11 cm3 molecule?1 s?1 and k2 = (3.8 = 0.9) × 10?11 cm3 molecule?1 s?1, for the reactions at 298 = 5 K.  相似文献   

16.
355 nm光照下利用瞬态吸收光谱技术进行了有氧、无氧条件下二苯醚与亚硝酸体系的反应机理研究, 考察了其中瞬态物种的衰减行为, 并对其光解产物进行了GC-MS分析. 研究表明, HNO2在355 nm紫外光的照射下产生的OH自由基和二苯醚反应生成C12H10O-OH 加合物, N2条件下C12H10O-OH衰减的速率常数为(1.86±0.14)×105 s-1, 在有氧条件下, C12H10O-OH可转化为C12H10O-OHO2, 衰减的速率常数为(6.6±0.4)×106 s-1. N2条件下最终产物为苯酚、2-羟基二苯醚、4-羟基二苯醚、4-硝基二苯醚.  相似文献   

17.
Nitrous acid(HONO),as a primary precursor of OH radicals,has been considered one of the most important nitrogencontaining species in the atmosphere.Up to 30%of primary OH radical production is attributed to the photolysis of HONO.However,the major HONO formation mechanisms are still under discussion.During the Campaigns of Air Quality Research in Beijing and Surrounding Region(CAREBeijing2006)campaign,comprehensive measurements were carried out in the megacity Beijing,where the chemical budget of HONO was fully constrained.The average diurnal HONO concentration varied from 0.33 to 1.2 ppbv.The net OH production rate from HONO,POH(HONO)net,was on average(from 05:00 to 19:00)7.1×106 molecule/(cm3 s),2.7 times higher than from O3 photolysis.This production rate demonstrates the important role of HONO in the atmospheric chemistry of megacity Beijing.An unknown HONO source(Punknown)with an average of 7.3×106molecule/(cm3 s)was derived from the budget analysis during daytime.Punknown provided four times more HONO than the reaction of NO with OH did.The diurnal variation of Punknown showed an apparent photo-enhanced feature with a maximum around 12:00,which was consistent with previous studies at forest and rural sites.Laboratory studies proposed new mechanisms to recruit NO2 and J(NO2)in order to explain a photo-enhancement of of Punknown.In this study,these mechanisms were validated against the observation-constraint Punknown.The reaction of exited NO2 accounted for only 6%of Punknown,and Punknown poorly correlated with[NO2](R=0.26)and J(NO2)[NO2](R=0.35).These results challenged the role of NO2 as a major precursor of the missing HONO source.  相似文献   

18.
Rate coefficients, k1, for the reaction OH + HONO → H2O + NO2, have been measured over the temperature range 298 to 373 K. The OH radicals were produced by 266 nm laser photolysis of O3 in the presence of a large excess of H2O vapor. The temporal profiles of OH were measured under pseudo-first-order conditions, in an excess of HONO, using time resolved laser induced fluorescence. The measured rate coefficient exhibits a slight negative temperature dependence, with k1 = (2.8 ± 1.3) × 10?12 exp((260 ± 140)/T) cm3 molecule?1 s?1. The measured values of k1 are compared with previous determinations and the atmospheric implications of our findings are discussed.  相似文献   

19.
The rate constant of the reaction of OH with DMS has been measured relative to OH + ethene in a 420 l reaction chamber at 760 torr total pressure and 298 ± 3 K in N2 + O2 buffer gas using the 254 nm photolysis of H2O2 as the OH source. In agreement with a recent absolute rate determination of the reaction the measured effective rate constant was found to increase with increasing partial pressure of O2 in the system, for 760 torr air a rate constant of (8.0 ± 0.5) × 10?12 cm3 s?1 was obtained. Product studies have been performed on the reaction in air using FTIR absorption spectrometry for detection of reactants and products. On a molar basis, SO2 was formed with a yield of 70% and dimethyl sulfone (CH3SO2CH3) with a yield of approximately 20%. These results are considerably different to those obtained in other product studies which were carried out in the presence of NOx. These differences are compared and their relevance for the atmospheric oxidation mechanisms of DMS is discussed.  相似文献   

20.
The optical transient and kinetics characterizations of the transients formed in the reaction of OH with benzotrifluoride (BTF) were performed by a laser flash photolysis technique. The results indicated that the formation of π‐type adduct of C6H5(OH)CF3 was the major reaction channel, and the δ‐type adduct of C6H5CF3OH formation was an additional minor process in the oxidation reaction of BTF attacked by OH radicals yielded from the photolysis of H2O2. Addition of OH to the CF3 group led to the fluoride ion elimination to yield α,α‐difluorophenylcarbinol (C6H5CF2OH). Trifluoromethylphenol (HOC6H4CF3) of meta‐, para‐ and ortho‐substituted isomers resulted from the addition of OH to the BTF aromatic ring.  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