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 78 毫秒
1.
The reactions of CH3O2 with SO2 and NO have been studied by steady state photolysis of azomethane in the presence of O2SO2→NO mixtures at 296 K and 1 atm total pressure. The quantum yield of NO oxidation by CH3O2 radicals is increased substantially when SO2 is added to the system indicating an SO2 induced chain oxidation of NO. The rate law gives k1/k2 = (2.5 ± 0.5) × 10?3 for CH3O2 + SO2 → CH3O2SO2 (1), CH3O2 + NO → CH3O + NO2 (2). Combining this ratio with the absolute value of k1 = 8.2 × 10?15 cm3 s?1 gives k2 = 10?11.5 ± 02 cm3 s?1.  相似文献   

2.
Gas-phase reactions typical of the Earth’s atmosphere have been studied for a number of partially fluorinated alcohols (PFAs). The rate constants of the reactions of CF3CH2OH, CH2FCH2OH, and CHF2CH2OH with fluorine atoms have been determined by the relative measurement method. The rate constant for CF3CH2OH has been measured in the temperature range 258–358 K (k = (3.4 ± 2.0) × 1013exp(?E/RT) cm3 mol?1 s?1, where E = ?(1.5 ± 1.3) kJ/mol). The rate constants for CH2FCH2OH and CHF2CH2OH have been determined at room temperature to be (8.3 ± 2.9) × 1013 (T = 295 K) and (6.4 ± 0.6) × 1013 (T = 296 K) cm3 mol?1 s?1, respectively. The rate constants of the reactions between dioxygen and primary radicals resulting from PFA + F reactions have been determined by the relative measurement method. The reaction between O2 and the radicals of the general formula C2H2F3O (CF3CH2? and CF3?HOH) have been investigated in the temperature range 258–358 K to obtain k = (3.8 ± 2.0) × 108exp(?E/RT) cm3 mol?1 s?1, where E = ?(10.2 ± 1.5) kJ/mol. For the reaction between O2 and the radicals of the general formula C2H4FO (? HFCH2O, CH2F?HOH, and CH2FCH2?) at T = 258–358 K, k = (1.3 ± 0.6) × 1011exp(?E/RT) cm3 mol?1 s?1, where E = ?(5.3 ± 1.4) kJ/mol. The rate constant of the reaction between O2 and the radicals with the general formula C2H3F2O (?F2CH2O, CHF2?HOH, and CHF2CH2?) at T = 300 K is k = 1.32 × 1011 cm3 mol?1 s?1. For the reaction between NO and the primary radicals with the general formula C2H2F3O (CF3CH2? and CF3?HOH), which result from the reaction CF3CH2OH + F, the rate constant at 298 K is k = 9.7 × 109 cm3 mol?1 s?1. The experiments were carried out in a flow reactor, and the reaction mixture was analyzed mass-spectrometrically. A mechanism based on the results of our studies and on the literature data has been suggested for the atmospheric degradation of PFAs.  相似文献   

3.
A jet-stream kinetic technique and the resonance fluorescence method applied to detection of iodine atoms were used to measure the rate constants of the reactions of the IO· radical with the halohydrocarbons CHFCl-CF2Cl (k = (3.2 ± 0.9) × 10?16 cm3 molecule s?1) and CH2ClF (k = (9.4 ± 1.3) × 10?16 cm3 molecule s?1), the hydrogen-containing haloethers CF3-O-CH3 (k = (6.4 ± 0.9) × 10?16 cm3 molecule s?1) and CF3CH2-O-CHF2 (k = (1.2 ± 0.6) × 10?15 cm3 molecule s?1), and hydrogen iodide (k = (1.3 ± 0.9) × 10?12 cm3 molecule s?1) at 323 K.  相似文献   

4.
Equilibrium vapor pressures were determined at temperatures between 294 K and 353 K for the NiCl2-CH3CN system.Compositions studied ranged from molar ratios of CH3CN to NiCl2 of 0.27 to 1.90. Three stoichiometric compounds were identified: NiCl2(CH3CN)2, NiCl2(CH3CN), NiCl2(CH3CN)0.88. Per mole of gaseous CH3CN the values of ΔHo and ΔSo were calculated to be 52.0 ±0.4 kJ mol?1 and 149.0±1.3 J mol?1 K?1 for the decomposition of NiCl2(CH3CN)2, and 25.9 ±0.8 kJ mol?1 and 58.6± 2.1 J mol?1 K?1 for the decomposition of NiCl2(CH3CN). Below a composition of NiCl2(CH3CN)0.88 the phase diagram is complex and could not be interpreted in terms of specific stoichiometric compounds.  相似文献   

5.
The incorporation of (±)-norcoclaurine, (±)-coclaurine, (±)-N-methylcoclaurine and dehydro-N-methylcoclaurine into nortiliacorinine A in Tiliacora racemosa colebr has been studied and specific utilisation of the (±)-coclaurine demonstrated. The evidence supports oxidative dimerization of two coclaurine units to give nortiliacorinine A. Experiments with (±)-N-methylcoclaurine and (±)-[1-3H, N-14CH3]N-methylcoclaurine established that only one N-methylcoclaurine unit is specifically utilised to constitute that “half” of the base which had phenolic OH group in the benzylic portion and further demonstrated that the H atom at the asymmetric centre in the 1-benzylisoquinoline precursor is retained in the bioconversion into nortiliacorinine A. Double labelling experiment with (±)-[1-3H, 6,0-14CH3]N-methylcoclaurine showed that O-Me function of the precursor is lost in the bioconversion into nortiliacorinine A. Parallel feedings of (+)-(S)- and (-)-(R)-N-methyl-coclaurines and (-)-(S)-, and ( + )-(R)-coclaurines revealed that the stereo-specificity is maintained in the biosynthesis of nortiliacorinine A from 1-benzylisoquinoline precursors and established ‘S,S’-configuration at the two asymmetric centres in nortiliacorinine A.  相似文献   

6.
The technique of laser photolysis of alkyl and perfluoroalkyl iodides at 266 nm followed by time-resolved detection of the 1.3-μm emission from I*(2P1/2) has been used to measure the rate constants for deactivation of I* by CH3I, C2H5I, CF3I, and CH4. The recommended values are (2.76± 0.22) × 10?13, (2.85 ± 0.40) × 10?13, (3.5 ± 0.5) × 10?17, and (7.52 ± 0.12) × 10?14, respectively, in units of cm3 molecule?1 S?1.  相似文献   

7.
From measurements of the heats of iodination of CH3Mn(CO)5 and CH3Re(CO)5 at elevated temperatures using the ‘drop’ microcalorimeter method, values were determined for the standard enthalpies of formation at 25° of the crystalline compounds: ΔHof[CH3Mn(CO)5, c] = ?189.0 ± 2 kcal mol?1 (?790.8 ± 8 kJ mol?1), ΔHof[Ch3Re(CO)5,c] = ?198.0 ± kcal mol?1 (?828.4 ± 8 kJ mo?1). In conjunction with available enthalpies of sublimation, and with literature values for the dissociation energies of MnMn and ReRe bonds in Mn2(CO)10 and Re2(CO)10, values are derived for the dissociation energies: D(CH3Mn(CO)5) = 27.9 ± 2.3 or 30.9 ± 2.3 kcal mol?1 and D(CH3Re(CO)5) = 53.2 ± 2.5 kcal mol?1. In general, irrespective of the value accepted for D(MM) in M2(CO)10, the present results require that, D(CH3Mn) = 12D(MnMn) + 18.5 kcal mol?1 and D(CH3Re) = 12D(ReRe) + 30.8 kcal mol?1.  相似文献   

8.
The reaction of (η5-C5H5)W(CO)2(NO), 6W, with P(CH3)3 proceeds rapidly at 25°C to give (η5-C5H5)W(CO)(NO)[P(CH3)3], 7W. The rate of formation of 7W was found to be 4.48 × 10?2M?1 [6W] [P(CH3)3] at 25.0°c in THF. In neat P(CH3)3 at ?23°C, 6W is converted to (η1-C5H5)W(CO)2(NO)[P(CH3)3]2, 8W. In dilute solution, 8W decomposes to initially give a 2:1 mixture of 6W and 7W. The mixture is then converted to 7W. The reaction of (η5-C5H5)Mo(CO)(NO), 6Mo, with P(CH3)3 is 6.1 times faster than that of the tungsten analog.  相似文献   

9.
Absorption transients at 254 nm have been observed in O3-O2 mixtures following laser irradiation at 9.64 μm. From analysis of these transients, we are able to determine vibrational relaxation rate constants (O3-O2 λ1?1/[O2] = (2560±370) Torr?1 S?1, λ2?1/[O2] = (640±50) Torr?1 S?1, and also a v1-v3 equilibration rate constant (O3-O3) of (1.5±1.0) × 106 Torr?1 S?1.  相似文献   

10.
Preparation of μ-Sulfurdisulfonium Salts [(CH3)2S? Sx? S(CH3)2]2+2A? (x = 1–3, A? = AsF6?, SbF6?, SbCl6?). On the Analogy of the Reactivity of Sulfanes and Sulfonium Salts The preparation of the μ-sulfurdisulfonium salts [(CH3)2S? Sx? S(CH3)2]2+(A?)2 with x = 1–3 and A? = AsF6?, SbF6?, SbCl6? is reported. The salts are formed by reaction of (CH3)2SH+A? and (CH3)2SSH+A? with SCl2 and S2Cl2, resp. They are characterized by vibrational spectroscopic measurements. [(CH3)2S? S2? S(CH3)2]2+(SbF6?)2 crystallizes in the space group C2/c with a = 1 884.5(7) pm, b = 1 302.8(5) pm, c = 1 477.2(5) pm, β = 98.62(3)° und Z = 8.  相似文献   

11.
The kinetics of 1,1-dimethylpropyl peroxy radicals recombination in polar solvents—water, methanol, and their mixtures—was studied by EPR spectroscopy in combination with the stopped-flow method, and the rate constants of this reaction were determined. Peroxyl radicals were generated by mixing solutions of Ce4+ sulfate and 1,1-dimethylpropyl hydroperoxide. The observed EPR signal of the peroxyl radical is a singlet with a g-factor of 2.015 ± 0.001, and a line width of ΔH = (1.36 ± 0.02) × 10?3 T for methanol and ΔH = (9.7 ± 0.2) × 10?4 T for water. The measured rate constants of (CH3)2C(O2·)CH2CH3 radical recombination at 298 K are 2kt = (3.9 ± 0.4) × 104 L mol?1 s?1 for water and 2kt = (5.2 ± 0.5) × 103 L mol?1 s?1 for methanol. A linear relationship between ln(2kt) and the Kirkwood function (ε?1)/(2ε + 1), where e is the dielectric constant of the medium, has been established, indicating an important role of nonspecific solvation in the recombination of tertiary peroxyl radicals.  相似文献   

12.
The values of pseudo-first-order rate constants (k obs) for the acetolysis of phthalic anhydride (PAn) increase from 6.60?×?10?7 to 31.5?×?10?7?s?1 with the increase in temperature from 30 to 50?°C. These values of k obs give activation parameters ?H* and ?S* as 14.4?±?0.4?kcal?mol?1 and ?39.1?±?1.3?cal?K?1?mol?1, respectively. The values of k obs remain essentially unchanged with the increase in the content of CS (CS?=?CH3CN or THF) from 0 to 40?% v/v in mixed AcOH?CCS solvents. These observations have been explained qualitatively.  相似文献   

13.
It has been confirmed by 1H and 13C NMR spectroscopies that Sn(σ-C7H7)Ph3 undergoes either 1,4- or 1,5-shifts of the SnPh3 moiety around the cycloheptatrienyl ring with ΔH3 = 13.8 ± 0.4 kcal mol?1, ΔS3 = ?5.6 ± 1.2 cal mol?1 deg?1, and ΔG3300 = 15.44 ± 0.14 kcal mol?1. Similarly, (σ-5-cyclohepta-1,3-dienyl)triphenyltin undergoes 1,5-shifts with ΔH3 = 12.4 ± 0.6 kcal mol?1, ΔS3 = ?11.2 ± 1.8 cal mol?1 deg?1, and ΔG3300 = 15.76 ± 0.13 kcal mol?1. It is therefore probable that Sn(σ-5-C5H5)R3 and Sn(σ-3-indenyl)R3 do not undergo 1,2-shifts as previously suggested but really undergo 1,5-shifts.  相似文献   

14.
Using a relative kinetic technique, rate coefficients have been measured, at 296 ± 2 K and 740 Torr total pressure of synthetic air, for the gas‐phase reaction of OH radicals with the dibasic esters dimethyl succinate [CH3OC(O)CH2CH2C(O)OCH3], dimethyl glutarate [CH3OC(O)CH2CH2CH2C(O)OCH3], and dimethyl adipate [CH3OC(O)CH2CH2CH2CH2C(O)OCH3]. The rate coefficients obtained were (in units of cm3 molecule?1 s?1): dimethyl succinate (1.89 ± 0.26) × 10?12; dimethyl glutarate (2.13 ± 0.28) × 10?12; and dimethyl adipate (3.64 ± 0.66) × 10?12. Rate coefficients have been also measured for the reaction of chlorine atoms with the three dibasic esters; the rate coefficients obtained were (in units of cm3 molecule?1 s?1): dimethyl succinate (6.79 ± 0.93) × 10?12; dimethyl glutarate (1.90 ± 0.33) × 10?11; and dimethyl adipate (6.08 ± 0.86) × 10?11. Dibasic esters are industrial solvents, and their increased use will lead to their possible release into the atmosphere, where they may contribute to the formation of photochemical air pollution in urban and regional areas. Consequently, the products formed from the oxidation of dimethyl succinate have been investigated in a 405‐L Pyrex glass reactor using Cl‐atom–initiated oxidation as a surrogate for the OH radical. The products observed using in situ Fourier transform infrared (FT‐IR) absorption spectroscopy and their fractional molar yields were: succinic formic anhydride (0.341 ± 0.068), monomethyl succinate (0.447 ± 0.111), carbon monoxide (0.307 ± 0.061), dimethyl oxaloacetate (0.176 ± 0.044), and methoxy formylperoxynitrate (0.032–0.084). These products account for 82.4 ± 16.4% C of the total reaction products. Although there are large uncertainties in the quantification of monomethyl succinate and dimethyl oxaloacetate, the product study allows the elucidation of an oxidation mechanism for dimethyl succinate. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 431–439, 2001  相似文献   

15.
The collisional behaviour of electronically excited silicon atoms in the 3p2(1S0) state, 1.909 eV above the 3p2(3P0) ground state, is investigated by time-resolved attenuation of atomic resonance radiation at λ = 390.53 nm (4s(1Po1)←3p2 (1S0)). The optically metastable Si(31S0) atoms were generated by the repetitive pulsed irradiation of SiCl4 and their decay monitored in the presence of added gases. Absolute quenching rate constants (kQ, cm3 molecule?1 s?1, 300 K) are reported for the following collision partners: He (?1.3 × 10?15), SiCl4 ((9.1 ± 1.4) × 10?11), O2 ((1.5 ± 0.2) × 10?11) and N2O ((4.3 ± 0.4) × 10?11). The results for O2 and N2O are compared with analogous data reported hitherto for Si(3p2(3PJ)) and with those for the other np2(1S0) states of the group IV atoms C, Ge, Sn and Pb. The rate data for the silicon atoms are considered in terms of the nature of the potential surfaces arising from symmetry arguments based on the weak spin orbit coupling approximation.  相似文献   

16.
By allowing dimethyl peroxide (10?4M) to decompose in the presence of nitric oxide (4.5 × 10?5M), nitrogen dioxide (6.5 × 10?5M) and carbon tetrafluoride (500 Torr), it has been shown that the ratio k2/k2′ = 2.03 ± 0.47: CH3O + NO → CH3ONO (reaction 2) and CH3O + NO2 → CH3ONO2 (reaction 2′). Deviations from this value in this and previous work is ascribed to the pressure dependence of both these reactions and heterogeneity in reaction (2). In contrast no heterogeneous effects were found for reaction (2′) making it an ideal reference reaction for studying other reactions of the methoxy radical. We conclude that the ratio k2/k2′ is independent of temperature and from k1 = 1010.2±0.4M?1 sec?1 we calculate that k2′ = 109.9±0.4M?1 sec?1. Both k2 and k2′ are pressure dependent but have reached their limiting high-pressure values in the presence of 500 Torr of carbon tetrafluoride. Preliminary results show that k4 = 10.9.0±0.6 10?4.5±1.1M?1 sec?1 (Θ = 2.303RT kcal mole?1) and by k4 = 108.6±0.6 10?2.4±1.1M?1 sec?1: CH3O + O2 → CH2O + HO2 (reaction 4) and CH3O + t-BuH → CH3OH + (t-Bu) (reaction 4′).  相似文献   

17.
The consecutive reactions of (CH3)2Si(OC2H5)2 and CH3Si(OC2H5)3 with methoxide ions were investigated in methanol solutions. The reverse transesterification reactions with ethoxide ions could be neglected in both cases since the concentration of ethoxide in methanol solution was assumed to be low due to the fast equilibrium reaction C2H5O? + CH3OH ? C2H5OH + CH3O?. The progress of the reactions was followed by monitoring the formation of ethanol with a Fourier-transform infrared spectrometer. All rate constants were determined at 295 K. The reactions between the dialkoxydimethylsilanes and methoxide ions were assumed to consist of two consecutive steps that can be represented by the net reaction; (CH3)2Si(OC2H5)2 + 2CH3O? → (CH3)2Si(OCH3)2 + 2C2H5O?. The two consecutive rate constants were established as 1.93 ± 0.12M?1s?1 and 1.00 ± 0.12M?1s?1, respectively. The consecutive rate constants for the reactions between the trialkoxymethylsilanes and methoxide ions can be written according to the total reaction; CH3Si(OC2H5)3 + 3CH3O? → CH3Si(OCH3)3 + 3C2H5O?. The three rate constants corresponding to each consecutive step were established as 1.12 ± 0.09 M?1s?1, 0.82 ± 0.10 M?1s?1, and 0.51 ± 0.06 M?1s?1, respectively. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
Rate constants for a series of alcohols, ethers, and esters toward the sulfate radical (SO4?) have been directly determined using a laser photolysis set‐up in which the radical was produced by the photodissociation of peroxodisulfate anions. The sulfate radical concentration was monitored by following its optical absorption by means of time resolved spectroscopy techniques. At room temperature the following rate constants were derived: methanol ((1.6 ± 0.2) × 107 M?1 s?1); ethanol ((7.8 ± 1.2) × 107 M?1 s?1); tert‐butanol ((8.9 ± 0.3) × 105 M?1 s?1); diethyl ether ((1.8 ± 0.1) × 108 M?1 s?1); MTBE ((3.13 ± 0.02) × 107 M?1 s?1); tetrahydrofuran (THF) ((2.3 ± 0.2) × 108 M?1 s?1); hydrated formaldehyde ((1.4 ± 0.2) × 107 M?1 s?1); hydrated glyoxal ((2.4 ± 0.2) × 107 M?1 s?1); dimethyl malonate (CH3OC(O)CH2C(O)OCH3) ((1.28 ± 0.02) × 106 M?1 s?1); and dimethyl succinate (CH3OC(O)CH2CH2C(O)OCH3) ((1.37 ± 0.08) × 106 M?1 s?1) where the errors represent 2σ. For the two latter species, we also measured the temperature dependence of the corresponding rate constants. A correlation of these kinetics with the bond dissociation energy is also presented and discussed. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 539–547, 2001  相似文献   

19.
The CH3O(X? 2E) radical produced by the 266 nm photolysis of CH3ONO is characterized by laser induced fluorescence. Using a flowing gas cell the reaction rate of CH3O(X? 2E) with NO is measured to be (2.08 ± 0.12) × 10?11 cm3 s?1 based upon disappearance of CH3O and appearance of HNO detected by laser induced fluorescence. Upper limits for CH3O reactions with CH4, CO, N2O, NH3, CH3OH, (CH3)3CH and CH2CHCH2CH3 are reported. These reactions are all too slow to measure under our experimental conditions.  相似文献   

20.
The spectral characteristics and the quantum yield of the fluorescence from the second excited singlet state S2 of the aromatic thioketone molecules xanthione (XS) and thioxanthione (TXS) have been determined in solution at room temperature and 77 K. In 3-methylpentane, the measured quantum yields are φf (295 K) = 5.1 × 10?3 and φf(77 K) = 1.0 × 10?2 for XS, and φf (295 K) = 1.5 × 10?3 and φf (77 K) = 2.5 × 10?3 for TXS. Using the Strickler-Berg expression for the radiative lifetime, the decay rate of S2 is derived. It is concluded that internal conversion S2 ? S1 is the dominating deactivation channel of S2 with k77 Knr(S2 ? S1) = 1.0 × 1010 s?1 for XS and k77 Knr (S2→S1) = 2.2 × 1010 s?1 for TXS. Between 295 and 77 K, φf increases by a factor of about 2 following an Arrhenius type expression. This temperature dependence of φf is considered to be intramolecular in nature and is attributed to a temperature sensitive rate constant knr(S2?S1) with an activation energy of 190 ± 20 cm?1 and a frequency factor knr = 3 × 1010 s?1 for the XS molecule in 3-methylpentane.  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