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.
One- and two-colour photoionisation spectra for NO2 have been investigated using a time of flight mass spectrometer as detector to find the most efficient REMPI process for analytical applications. Two different inlet systems have been employed: a pulsed supersonic jet expansion stage and a flow reactor. Selective and sensitive mass spectrometric determinations of free NO2 have been possible even in the presence of high concentrations of organic nitrates, HNO3 and other NO2 precursors. Employing two-colour (1+1+1) excitation using a concentration of HNO35·1014 molecules/cm3 a detection limit of 5·1011 molecules/cm3 has been found for NO2 whereas in the absence of HNO3 a detection limit of 5·1010 molecules/cm3 is reported.  相似文献   

2.
The reaction 2NO2 + ROH = RONO + HNO3 (R = CH3 or C2H5) has been studied using the FTIR method at reactant pressures from 0.1 to 1.0 torr at 25°C. The termolecular rate constant for the forward reaction was determined to be (5.7 ± 0.6) × 10?37 cm6/molec2·s for CH3OH and (5.7 ± 0.8) × 10?37 cm6/molec2·s for C2H5OH, that is, d[RONO]/dt = k[NO2]2[ROH]. The corresponding equilibrium constants were measured as 1.36 ± 0.06 and 0.550 ± 0.025 torr?1, respectively. These results are consistent with those of a previous study based on the NO2 decay measurements at reactant pressures from 1 to 10 torr.  相似文献   

3.
The thermal decomposition of CH3NO2 highly diluted in Ar has been studied in shock waves at 900 < T < 1500 K and 1.5 · 10?5 < [Ar] < 3.5 · 10?4 mol/cm3. Concentration profiles of CH3NO2 and NO2 were recorded. The unimolecular reaction was found to be in its fall-off range. Limiting low pressure rate constants of k0 = [Ar] · 1017.1 exp(?42(kcal/mol)/RT) cm3/ mol sec in the range 900 < T < 1400 K and limiting high pressure rate constants of k = 1016.25 exp (?(58.5 ± 0.5 kcal/mol)/RT) sec?1 have been derived. A rate constant of 1.3 · 1013 cm3/mol sec was found for the first subsequent reaction CH3+NO2 → CH3O+NO.  相似文献   

4.
Preparation und Characterization of Phthalocyanine-π-Cation-Radicals of H+, Mg2+, and Cu2+ The preparation of phthalocyanine-π-cation-radicals (Pc(?1)) of H+, Mg2+, Cu2+ is described. MgClPc(?1) and Cu(NO3)Pc(?1) · HNO3 are isolated as stoichiometrically pure, stable redbrown solids. Contrary to the phthalocyanines(?2) (Pc(?2)) these are very soluble with redviolet colour in organic solvents in the presence of R? COOH (R ? H, CF3, CCl3). The electronic absorption absorption spectra (UV-VIS) are remarkably solvent-dependent. This solvent effect is due to a reversible radical association. Monomeric radical species exist in nonpolar (CH2Cl2), dimeric in polar solvents (CH3NO2, C2H5OH). The UV-VIS, infrared (IR), and resonance-raman (RR) spectra of MgClPc(?1) and Cu(NO3)Pc(?1) · HNO3 are discussed and compared with the analogoues spectra of MgPc(?2) · 2 H2O and MgPc(?2) · HCl. Although there are only minor differences in the chemical composition and the electronic structure the spectroscopic data vary significantly for every complex. Especially the IR spectrum is suitable for a quick demonstration of the π-cation-radicals. The diagnostic bands are at ca. 1350 and 1450 cm?1.  相似文献   

5.
Indoor semivolatile organic compounds (SVOCs) originate from indoor and outdoor sources. These SVOCs partition among different phases and available surfaces, which increases their residence time indoors to several years. SVOCs may also react with indoor oxidants, such as hydroxyl radicals (OH), nitrate radicals (NO3), and ozone. In the present study, the second‐order reaction rate constants of 72 SVOCs in indoor air (gas and particle phases) at room temperature and ambient air pressure were retrieved from the literature. The pseudo–first‐order reaction rate constants of these SVOCs were calculated for the indoor concentration ranges of OH, NO3, and ozone. Then, the extent to which the chemical reaction had a meaningful impact on the removal of SVOCs from the indoor environment was quantitatively analyzed. The orders of magnitude of the second‐order rate constant ranged between 10−15 and 10−10 cm3/(molecule·s) for OH/SVOC reactions, 10−17 and 10−12 cm3/(molecule·s) for NO3/SVOC reactions, and 10−20 and 10−16 cm3/(molecule·s) for ozone/SVOC reactions in indoor air. Assuming that the highest indoor reactant concentrations were 1.8 × 106 molecules/cm3 (7.3 × 10−5 ppb) for OH, 2.5 × 108 molecules/cm3 (10−2 ppb) for NO3, and 1.4 × 1012 molecules/cm3 (58 ppb) for ozone, the highest pseudo–first‐order rate constants in the gas phase for the studied reactions of OH/SVOCs (n = 72), NO3/SVOCs (n = 3), and ozone/SVOCs (n = 14) reached 1.5 h−1 (OH/benzo[a]pyrene), 0.41 h−1 (NO3/acenaphthene), and 1.0 h−1 (ozone/aldrin and ozone/heptachlor), respectively. The pseudo–first‐order rate constants in the particle phase for the studied reactions of OH/SVOCs (n = 13), NO3/SVOCs (n = 6), and ozone/SVOCs (n = 14) at the high indoor reactant concentrations reached 0.09 h−1 (OH/DEHP), 5.8 h−1 (NO3/pyrene), and 11 h−1 (ozone/benzo[a ]pyrene), respectively. These results indicate that the chemical reactions of some SVOCs in indoor air have a meaningful impact compared to the air exchange rate, which should be considered in future studies on indoor air quality modeling.  相似文献   

6.
The kinetics of the reaction of F atom with HNO3, source of NO3 radicals widely used in laboratory studies, has been investigated at nearly 2.7 mbar total pressure of helium over a wide temperature range, T = 220-700 K, using a low-pressure discharge flow reactor combined with an electron impact ionization quadrupole mass spectrometer. The rate constant of the reaction F + HNO3 → NO3 + HF (1) was determined using both relative rate method and absolute measurements under pseudo–first-order conditions, monitoring the kinetics of F-atom consumption in excess of HNO3, k1 = (8.2 ± 0.4) × 10−12 exp((315 ± 15)/T) cm3 molecule−1 s−1 (where the uncertainties represent precision at the 2σ level, the estimated total uncertainty on k1 being 15% at all temperatures). The reaction rate constant was found to be in excellent agreement with the only previous temperature-dependent study. Experiments on detection of the reaction product, HF, have shown that NO3 and HF forming channel of the title reaction is the dominant, if not unique, on the whole temperature range of the study.  相似文献   

7.
A method for the determination of nickel in ion exchange resins has been elaborated. It is based on the58Ni(n, p)58Co reaction. Samples of 500 mg are irradiated for 90 h in a fission neutron flux of 1013 n·cm−2·s−1. After decomposing by HNO3/HClO4 mixture the radiochemical separation is carried out by extraction of the Co-DDTC-complex in chloroform. Measuring by a 23 cm3 Ge(Li) detector for 15 h provides a detection limit of 10 ppb. Radiochemical yield is determined by57Co as radioactive indicator. Limitations by60Co are reduced by Cd-screened activation and by anticoincidence measuring technique.   相似文献   

8.
Rate constants have been determined for the reaction OH + NO2 (+ N2) → HNO3 (+ N2), using time-resolved resonance absorption to follow the removal of OH radicals produced by flash photolysis of HNO3. The measurements cover the ranges: 220 ? T ? 358 K and 3.2 × 1017 ? [N2] ? 4.0 × 1018 molecule cm?3.  相似文献   

9.
刘佩芳  文利柏 《中国化学》1998,16(3):234-242
The mass transport and charge transfer kinetics of ozone reduction at Nafion coated Au electrodes were studied in 0.5 mol/L H2SO4 and highly resistive solutions such as distilled water and tap water. The diffusion coefficient and partition coefficient of ozone in Nafion coating are 1.78×10-6 cm2·s-1 and 2.75 at 25℃ (based on dry state thickness), respectively. The heterogeneous rate constants and Tafel slopes for ozone reduction at bare Au are 4.1×10-6 cm·s-1, 1.0×10-6 cm·s-1 and 181 mV, 207 mV in 0.5 mol/L H2SO4 and distilled water respectively and the corresponding values for Nafion coated Au are 5.5×10-6 cm·s-1, 1.1×10-6 cm·s-1 and 182 mV, 168 mV respectively. The Au microelectrode with 3 μm Nafion coating shows good linearity over the range 0-10 mmol/L ozone in distilled water with sensitivity 61 μA·ppm-1 ·cm-2, detection limit 10 ppb and 95% response time below 5 s at 25℃. The temperature coefficient in range of 11-30℃ is 1.3%.  相似文献   

10.
The distribution of nitric acid between an aqueous phase of constant or variable ionic strength and a benzene solution of diphosphine dioxide can be explained by the following reactions H+a+ NO3-a+ DiPO0 ? D1PO·HNO30 H+a+ NO3-a+ DiPO·HNO30 ? DiPO·2 HNO30 At constant ionic strength, the stability constants K1″ were determined for the complexes 1,1-DiPO·HNO3 (98 ± 01 (M)-1), 1,4-DiPO·HNO3(44±3 (M)-1) and 1,5-DiPO·HNO3 (51 ± 1 (M)-1). The constants K11″ for the complexes 1,1-DiPO·2 HNO3 and 1,5-DiPO.2 HNO3 are respectively 035±001 (M)-1 and 62 ±0.05 (M)-1 at 25°. With an aqueous phase of variable ionic strength, values of K1'=54±7 (M)-2 for 1,5-D1PO.HNO3 and KII'=65 ± 04 (M)-2 for 1,5-DiPO·2 HN03 were obtained  相似文献   

11.
The Absolute rate constants for the gas-phase reactions of NO3 with HO2 and OH have been determined using the discharge flow laser magnetic resonance method (DF-LMR). Since OH was found to be produced in the reaction of HO2 with NO3, C2F3Cl was used to scavenge it. The overall rate constant, k1, for the reaction, HO2 + NO3 → products, was measured to be k1=(3.0 ± 0.7)×10?12 cm3 molecule?1 s?1 at (297 ± 2) K and P=(1.4 – 1.9) torr. This result is in reasonable agreement with the previous studies. Direct detection of HO2 and OH radicals and the use of three sources of NO3 enabled us to confirm the existence of the channel producing OH:HO2+NO3→OH+NO2+O2 (1a); the other possible channel is HO2+NO3→HNO3+O2 (1b). From our measurements and the computer simulations, the branching ratio, k1a/(k1a + k1b), was estimated to be (1.0). The rate coefficient for the reaction of OH with NO3 was determined to be (2.1 ± 1.0) × 10?11 cm3 molecule?1 s?1. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Reactions of CF3Br with H atoms and OH radicals have been studied at room temperature at 1–2 torr pressures in a discharge flow reactor coupled to an EPR spectrometer. The rate constant of the reaction H + CF3Br → CF3 + HBr (1) was found to be k1 = (3.27 ± 0.34) × 10?14 cm3/molec·sec. For the reaction of OH with CF3Br (8) an upper limit of 1 × 10?15 cm3/molec·sec was determined for k8. When H atoms were in excess compared to NO2, used to produce OH radicals, a noticeable reactivity of OH was observed as a result of the reaction OH + HBr → H2O + Br, HBr being produced from reaction (1).  相似文献   

13.
The rate of reaction between NO and HNO3 and the rate of thermal decomposition of HNO3 have been measured by FTIR spectroscopy. The measurements were made in a teflon lined batch reactor having a surface to volume ratio of 14 m?1. During the experiments, with initial HNO3 concentrations between 2 and 12 ppm and NO concentrations between 2 and 30 ppm, a reactant stoichiometry of unity and a first order NO and HNO3 dependence were confirmed. The observed rate constant for the reaction at 22°C and atmospheric pressure was determined to 1.1 (±0.3) 10?5 ppm?1 min?1. At atmospheric pressure, HNO3 decomposes into NO2 and other products with a first order HNO3 dependence and with a rate constant of 2.0 (±0.2) 10?3 min?1. The apparent activation energy for the decomposition is 13 (±4) kJ mol?1.  相似文献   

14.
Our earlier work on the formation of particulate NH4NO3 in the NH3? O3 reaction at 25°C is extended to include air as a diluent and H2O vapor as an additive. More extensive data at different values of [NH3]/[O3]0 were obtained also, where [O3]0 is the initial O3 concentration. In our earlier work we concluded that NH4NO3 vapor was dissociated to NH3 + HNO3 and that the HNO3 was removed by diffusion to the walls with a rate coefficient kdiff = 0.4 min?1 or by condensation on the suspended particles. Particles were nucleated by 8 NH3? HNO3 pairs when their concentration product reached 5.8 × 1027 molec2/cm6 with a rate coefficient knucl of 6.2 × 10?224 cm45/min and removed by coagulation with a rate coefficient kcoag of 1.3 × 10?7 cm3/min. A corrected calculation modifies the number of pairs required to 6–7 with a correspondingly changed value of knucl. With the more extensive data of the present study the indications are that the vapor-phase NH4NO3 monomer is not dissociated and that its diffusion constant for loss to the walls varies between 0.3 and 0.9 min?1 for different reaction conditions. Nucleation occurs when the NH4NO3 vapor concentration reaches 1.0 × 1012 molec/cm3 via. where r is 9 and the nucleation rate coefficient knucl is 3 × 10?108 cm24/min. With 5.0 or 9.5 torr of H2O vapor present, there is an excess of particles produced over that expected from this rate coefficient, indicating an additional nucleation step in which H2O vapor participates directly to produce a hydrated salt. The coagulation coefficient of (1.87 ± 0.14) × 10?7 cm3/min found here is in good agreement with that found previously.  相似文献   

15.
A shock wave study of the thermal decomposition of nitroethane in excess Ar at temperatures 900 < T < 1350 K and total concentrations of 4,5 · 10?6 < [Ar] < 3 · 10?4 mol cm?3 showed that the C? N-bond fission is the primary reaction step. This unimolecular reaction could be observed in its transition region near the high pressure limit. The derived rate constants are k = 1015.9 exp (–57 kcal mol?1/RT) s?1 for the high pressure and k0/[Ar] = 1018.0 exp (-36 kcal mol?1/RT) cm3 mol?1 s?1 (at T ? 1100–1200 K) for the low pressure limit. The observed concentration profiles of C2H5NO2 and NO2 permitted to conclude on the subsequent decomposition of the ethyl radical This reaction was found to be in the fall-off range under the applied conditions.  相似文献   

16.
The rate coefficients of the reactions of CN and NCO radicals with O2 and NO2 at 296 K: (1) CN + O2 → products; (2) CN + NO2 → products; (3) NCO + O2 → products and (4) NCO + NO2 → products have been measured with the laser photolysis-laser induced fluorescence technique. We obtained k1 = (2.1 ± 0.3) × 10?11 and k2 = (7.2 ± 1.0) × 10?11 cm3 molecule?t s?1 which agree well with published results. As no reaction was observed between NCO and O2 at 297 K, an upper limit of k3 < 4 × 10?17 cm3 molecule?1 S?1 was estimated. The reaction of NCO with NO2 has not been investigated previously. We measured k4 = (2.2 ± 0.3) × 10?11 cm3 molecule?1 s?1 at 296 K.  相似文献   

17.
The chemical and electrochemical properties of technetium metal were studied in 1–6 M HX and in 1 M NaX (pH 1 and 2.5), X = Cl, NO3. The chemical dissolution rates of Tc metal were higher in HNO3 than in HCl (i.e. 8.63 × 10?5 mol cm?2 h?1 in 6 M HNO3 versus 2.05 × 10?9 mol cm?2 h?1 in 6 M HCl). The electrochemical dissolution rates in HNO3 and HCl were similar and mainly depended on the electrochemical potential and the acid concentration. The optimum dissolution of Tc metal was obtained in 1 M HNO3 at 1 V/AgAgCl (1.70 × 10?3 mol cm?2 h?1). The dissolution potentials of Tc metal in nitric acid were in the range of 0.596–0.832 V/AgAgCl. Comparison of Tc behavior with Mo and Ru indicated that in HNO3, the dissolution rate followed the order: Mo > Tc > Ru, and for dissolution potential the order: E diss(Ru) > E diss(Tc) > E diss(Mo). The corrosion products of Tc metal were analyzed in HCl solution by UV–Visible spectroscopy and showed the presence of TcO4 ?. The surface of the electrode was characterized by microscopic techniques; it indicated that Tc metal preferentially corroded at the scratches formed during the polishing and no oxide layer was observed.  相似文献   

18.
Microwave plasma torch (MPT), traditionally used as the light source for atomic emission spectrophotometry, has been employed as the ambient ionization source for sensitive detection of uranium in various ground water samples with widely available ion trap mass spectrometer. In the full‐scan mass spectra obtained in the negative ion detection mode, uranium signal was featured by the uranyl nitrate complexes (e.g. [UO2(NO3)3]?), which yielded characteristic fragments in the tandem mass spectrometry experiments, allowing confident detection of trace uranium in water samples without sample pretreatment. Under the optimal experimental conditions, the calibration curves were linearly responded within the concentration levels ranged in 10–1000 µg·l?1, with the limit of detection (LOD) of 31.03 ng·l?1. The relative standard deviations (RSD) values were 2.1–5.8% for the given samples at 100 µg·l?1. The newly established method has been applied to direct detection of uranium in practical mine water samples, providing reasonable recoveries 90.94–112.36% for all the samples tested. The analysis of a single sample was completed within 30 s, showing a promising potential of the method for sensitive detection of trace uranium with improved throughput. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

19.
The yields of hydrogen atoms, oxygen atoms and molecules, and hydroxyl radicals after a microwave discharge in the mixture of CO2 and H2 were measured by ESR spectroscopy in a flow-type system. A mathematical model of the kinetics of chemical reactions downstream the microwave discharge was devised. The concentrations of particles that cannot be detected under our experimental conditions were estimated. Experimental values of the concentration sensitivity for an RE-1306 ESR spectrometer are as follows: for a pressure of 1 Torr and optimized detection conditions, H., 1011 cm−3; O., 3·1010 cm−3; OH., 1010 cm−3; O2, 3·1013 cm−3 (Ref. 7); for a pressure of 2 Torr, H., 5·1012 cm−3; O., 2·1012 cm−3; OH., 2.5·1011 cm−3; O2, 7.5·1014 cm−3 8 Translated fromIzvestiya Akademii Nauk, Seriya Khimicheskaya, No. 4, pp. 665–669, April, 2000.  相似文献   

20.
The molecular modulation spectroscopic technique was employed to study the kinetics of NO3 radicals produced in the 253.7 nm photolysis of flowing gas mixtures of HNO3/CH4/O2 at room temperature. By computer fitting of the NO3 temporal behavior, a rate coefficient of (2.3 ± 0.7) × 10?12 cm3 molecule?1 s?1 was obtained for the reaction between NO3 and CH3O2 at 298 K.  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