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 46 毫秒
1.
The reaction of formaldehyde with HO2 radicals in the presence of O2 and NO2 has been studied in a 420 ℓ reaction chamber at 0° C in 533 mbar of synthetic air. Reactants and products were measured by FTIR absorption spectrometry-Additional evidence is presented for the formation of the HOCH2OO radical as the primary reaction product, by the IR spectroscopic identification of its NO2 recombination product HOCH2OONO2. By computer simulation of the concentration-time profiles of HO2NO2, H2CO and HOCH2OONO2, the rate constants (0°C, 533 mbar, M = air) k1 = (1.1 ± 0.4) × 10-13 cm3 s-1 and k-1 = 20-10+20 s-i have been derived for the reactions (1, -1) HO2 + H2CO ⇌ HOCH2OO.  相似文献   

2.
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 HNO3≤5·1014 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   

3.
The optical gas recognition capabilities of thin film layer of 4-[bis[(3,5-dimethyl-1H-pyrazol-1-yl)methyl]-amino]phenol deposed on quartz substrates were studied. The dynamic gas responses to the following analytes have been investigated as air pollutants (SO2, NO2, CO, CH4 and NH3). The spin-coated bispyrazole layer appears to have reversible response towards SO2 and a very low and irreversible response to NO2. The selectivity of the thin film based on bispyrazole layer with respect to other analytes was also examined and the present data show that the thin sensing layer in the presence of CO, CH4 and NH3 in low concentration does not influence its optical properties.  相似文献   

4.
The formation yields of selected products of the OH radical-initiated reactions of toluene, o-xylene, and 1,2,3,-trimethylbenzene have been measured in the absence of NOx and in the presence of varying concentrations of NO and NO2. The formation yield of o-cresol from toluene increased from 0.123 ± 0.022 in the absence of NOx to 0.160 ± 0.008 for an average NO2 concentration of 1.7 × 1014 molecule cm3. The formation yield of 2,3-butanedione from o-xylene was 0.092 ± 0.013 in the absence of NOx, and in the presence of NOx decreased from 0.16 at an average NO2 concentration of (7–8) × 1012 molecule cm?3 to 0.09 at an average NO2 concentration of ca. 7 × 1013 molecule cm?3. The formation yield of 2,3-butanedione from 1,2,3-trimethylbenzene increased from 0.18 in the absence of NOx to 0.444 ± 0.053 in the presence of ca. (0.16–3.6) × 1013 molecule cm?3 of NO2. These product data are consistent with literature kinetic data showing that the hydroxycyclohexadienyl radicals formed by OH radical addition to the aromatic ring react with both O2 and NO2 and with the NO2 reaction rate constants being ca. 105 higher than the O2 reaction rate constants at room temperature. Under typical tropospheric conditions the reactions of the hydroxycyclohexadienyl radicals with O2 will dominate over their reactions with NO2. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
《Supramolecular Science》1997,4(3-4):427-435
The response of an NO2 sensing system based on LB films of a polysiloxane with azobenzene chromophoric side-chains has been investigated. Changes in absorbance on exposure to 100 ppm NO2 have been recorded using UV-visible absorption spectroscopy from which changes in extinction coefficient (Δ k ≈ 0.033 at 500 nm) have been determined. Shallow angle X-ray reflectivity (SAXR) studies indicate a change in layer thickness from 2.10nm in air to 2.31 nm in 10000 ppm NO2 together with loss of Bragg detail. Changes in real refractive index (Δn ≈ 0.107 over most of the visible region) for films in air and 100 ppm NO2 have been deduced from reflectance spectra.  相似文献   

6.
Peroxynitric acid has been identified by Fourier transform, infrared, long-path spectroscopy as a product of the UV-irradiated dilute mixtures of HONO, CO, Nox in synthetic air at 24 ± 2°C. The characteristic peaks of this compound are identical to those observed by Niki, in irradiated Cl2, H2, NO2, air mixtures and Hanst, in irradiated Cl2, CH2O, NO2, air mixtures. The concentration of the reactants and products of the HONO, CO, NOx, air system have been determined as a function of irradiation time. The changes observed were computer simulated using a combination of thirty elementary reactions. The data suggests that both reactions (1) and (2) occur as a result of HO2-NO2 interactions in the gas phase: HO2 + NO2 å HONO + O2 (1); HO2 + NO2(+M) å HO2NO2(+M)(2). The data give the preliminary estimates: k1/k2 ≈0.7 ±0.4; k1/k3 ≈ 0.043 ± 0.02; k2/k3 ≈ 0.058 ± 0.02, where reaction (3) is: HO2 + NO å HO + NO2. These data and computer simulations of sunlight-irradiated, NOx, hydrocarbon, aldehyde-polluted atmospheres suggest that peroxynitric acid may be formed in urban atmospheres at rates which are comparable to those observed for peroxyacetyl nitrate (PAN), and it is concluded that HO2NO2 may contribute to the “oxidant” found in these atmopheres.  相似文献   

7.
The equilibrium constant, Keq of the reaction NO2 + NO3 + M 2 N2O5 + M has been determined for a small range of temperatures around room temperature in air at 740 torr by direct spectroscopical measurements of NO2, NO3, and N2O5. At 298 K, Keq was determined as (3.73 ± 0.61) × 10−11 cm3 molecule−1. Averaging this and 11 other independent evaluations of Keq yields Keq = (3.31 ± 0.82) × 10−11 cm3 molecule−1, where the uncertainty is given as one standard deviation. The kinetics of the O3/NO2/N2O5/NO3/ air system was studied in a static chamber at room temperature and 740 torr total pressure. Evidence of a unimolecular decay reaction of NO3, NO3 → NO + O2, was found and its rate coefficient was estimated as (1.6 ± 0.7) × 10−3 s−1 at 295 ± 2 K.  相似文献   

8.
The reactions of naphthalene in N2O5? NO3? NO2? N2? O2 reactant mixtures have been investigated over the temperature range 272–297 K at ca. 745 torr total pressure and at 272 K and ca. 65 torr total pressure using long pathlength Fourier transform infrared absorption spectroscopy. 2,3-Dimethyl-2-butene was added to the reactant mixtures at 272 K to rapidly scavenge the NO3 radicals both initially present in the added N2O5 and formed from the thermal decomposition of N2O5 during the reactions. The data obtained in the presence and absence of added 2,3-dimethyl-2-butene showed that napthalene undergoes initial reaction with the NO3 radical to form an NO3-naphthalene adduct, which either rapidly decomposes back to the reactants (at a rate of ca. 5 × 105 s?1 at 298 K) or reacts exclusively with NO2 to form products. When NO3 radicals, N2O5 and NO2 are in equilibrium, this overall process is kinetically equivalent to reaction of naphthalene with N2O5, and previous kinetic and product studies have indeed assumed the reactions of naphthalene and alkyl-substituted naphthalenes in N2O5? NO3? NO2? air mixtures to be with N2O5, and not with NO3 radicals.  相似文献   

9.
Three polyamine ligands of N1-(2-nitrobenzyl)-N1-(2-aminoethyl)ethane-1,2-diamine (L1), N1-(2-nitrobenzyl)-N1-(2-aminoethyl)propane-1,3-diamine (L2) and N1-(2-nitrobenzyl)-N1-(3-aminopropyl)propane-1,3-diamine (L3) were synthesized and their cyclocondensation with 2-[2-(2-formyl phenoxy)ethoxy]benzaldehyde (L4) in the presence of various metal(II) ions was examined. These reactions only in the presence of a stoichiometric amount of cadmium(II) nitrate gave the related cadmium(II) macrocyclic Schiff-base complexes. In all the other cases no cyclic complexes have been obtained and metal(II) polyamines were the only products. The complexes have been studied with IR, 1H NMR, 13C NMR, DEPT, COSY, HMQC and microanalysis. The crystal structures of [Cd(NO3)(L5)(μ-NO3)Cd(NO3)(L5)]0.5Cd(NO3)4 (1) and [CdL5(NO3)(CH3OH)]ClO4 (2) have been also determined.  相似文献   

10.
The decomposition mechanism of Mg(NO3)2·6H2O was studied by means of simultaneous TG, DTG and DTA method combined with EGA technique under conventional and quasi isothermal-quasi isobaric conditions. It has been found that Mg(NO3)2·6H2O melts at 89 °C in a congruent way. The solution formed begins to boil at 147 °C. The water loss process of the salt hydrate and the decomposition process of the Mg(NO3)2 always overlap to some extent. Accordingly, Mg(NO3)2 of stoichiometric composition cannot be prepared thermally, because the compound always contains some basic salt. The last part of water departs in the vicinity of 270 °C with extreme rapidity. In contrast to expectations the compound decomposes in pure “self-generated” atmosphere at a temperature lower by about 80 °C than in the presence of air which contains a small amount of the gaseous decomposition product.  相似文献   

11.
A new iron basic salt, Fe4(OH)11NO3·2H2O, has been prepared by partially hydrolyzing a solution of Fe(NO3)3·9H2O with urea. The X-ray powder diffraction pattern has been indexed within a monoclinic cella=9.99(3) ?,b=9.48(2) ?,c=3.074(3) ? andβ=90.57(1)°. Thermal decomposition reactions in still air and nitrogen flow have been studied by DTA and TG analysis, and the intermediate and final products have been characterized by X-ray diffraction and IR spectroscopy. When this material is thermally decomposed in an X-ray high temperature diffraction chamber, pure iron is formed at 900 °C together with Fe(III) and Fe(II) oxides.
Zusammenfassung Mittels Hydrolyse einer L?sung von Fe(NO)3)3·9H2O mit Karbamid wurde das neue basische Eisensalz Fe4(OH)11NO3·2H2O dargestellt. Aus einem R?ntgenpulververfahren resultierena=9,55(3) ?,b=9,48(2) ?,c=3,074(3) ? undβ=90,57(1)° für eine monozyklische Zelle. Mittels DTA- und TG-Untersuchungen wurden die thermischen Zersetzungsreaktionen an Luft und im Stickstoffflu? untersucht und die Zwischen- und Endprodukte mit r?ntgendiffraktionsverfahren und IR-Spectroskopie charakterisiert. Bei einer thermischen Zersetzung dieses Stoffes in einer Hochtemperatur-r?ntgendiffraktionskammer wird bei 900 °C elementares Eisen zusammen mit Fe(II)- und Fe(III)-oxiden gebildet.

Резюме Частичным гидролизо м раствора соли Fe(NO3)3 · 9H2O с мочевиной получен а новая основная соль Fe4(OH)11NO3 · 2Н2О, для которой методо м порошкового рентген оструктурного анализа была установ лена моноклинная стр уктура с параметрами ячейкиа=9,55(3) А,b=9,48(2) ?,c=3,074(3) ? иβ=90,57(1)°. Термиче ское разложение соли изучено методом ДТА и ТГ в динамическо й атмосфере воздуха и азота, а образующиеся промеж уточные и конечные продукты ре акции были охарактер изованы рентгенофазовым ана лизом и ИК спектроскопией. ˉПри термическом разложе нии соли в высокотемпературно й рентгено-диффракци онной камере при 900° образует ся чистое железо вмес те с оксидами двух- и трехвалентного желе за.


The authors are greateful to Dr. R. M. Rojas for his helpful suggestions.  相似文献   

12.
The oxidation of 1-hexene by molecular oxygen catalyzed by iridium(III) complexes, [Ir(CH3CN)5−xClx(NO2)]2−x (x=0, 1, or 2) has been studied in acetonitrile under P(O2)=1.5 atm and T=100°C. [Ir(CH3CN)5(NO2)](PF6)2 oxidizes 1-hexene to 1,2-epoxyhexane. Complex [Ir(CH3CN)4Cl(NO2)]PF6 oxidizes 1-hexene to 2,3-epoxyhexane only in the presence of [Pd(PhCN)2(Cl)2] (an olefin activator). In contrast to the cationic complexes, the neutral complex [Ir(CH3CN)4Cl2(NO2)] oxidizes 1-hexene to 2-hexanone only in the presence of [Pd(PhCN)2(Cl)2].  相似文献   

13.
The thermal decomposition of nitrocellulose (NC) 12.1% N, has been studied with regard to kinetics, mechanism, morphology and the gaseous products thereof, using thermogravimetry (TG), differential thermal analysis (DTA), IR spectroscopy, differential scanning calorimetry (DSC) and hot stage microscopy. The kinetics of the initial stage of thermolysis ofNC in condensed state has been investigated by isothermal high temperature infrared spectroscopy (IR). The decomposition ofNC in KBr matrix in the temperature range of 142–151°C shows rapid decrease in O?NO2 band intensity, suggesting that the decomposition of NC occurs by the rupture of O?NO2 bond. The energy of activation for this process has been determined with the help of Avrami-Erofe'ev equation (n=1) and is ≈188.35 kJ·mol?1. Further, the IR spectra of the decomposition products in the initial stage of thermal decomposition ofNC, indicates the presence of mainly NO2 gas and aldehyde.  相似文献   

14.
An experimental and modeling study of irradiated toluene–NOx–air, toluene–benzaldehyde–NOx–air, and cresol–NOx–air mixtures at part-per-million concentrations has been carried out. These mixtures were irradiated at 303 ± 1 K in a 5800-liter Teflon-lined, evacuable environmental chamber, with temperature, humidity, light intensity, spectral distribution, and the concentrations of O3, NO, NO2, toluene, PAN, formaldehyde, benzaldehyde, o-cresol, m-nitrotoluene, and methyl nitrate beingmonitored as a function of time. For the toluene and toluene–benzaldehyde–NOx–air runs a variety of initial reactant concentrations were investigated. Cresol–NOx–air runs were observed to be much less reactive in terms of O3 formation and NO to NO2 conversion rates than toluene–NOx–air runs, with the relative reactivity of the cresol isomers being in the order meta » ortho > para. The addition of benzaldehyde to toluene–NOx–air mixtures decreased the reactivity, in agreement with previous studies. Alternative mechanistic pathways for the NOx photooxidations of aromaticsystems in general are discussed, and the effects of varying these mechanistic alternatives on the model predictions for the toluene and o-cresol–NOx–air systems are examined. Fits of the calculations to most of the experimental concentration–time profiles could be obtained to within the experimental uncertainty for two of the mechanistic options considered. In both cases it is assumed that (1) O2 adds to the OH–toluene adduct ~75% of the time forming, after a further addition of O2, a C7 bicyclic peroxy radical, and (2) this C7 bicyclic peroxy radical reacts with NO ~75% of the time to ultimately form α-dicarbonyls and conjugated γ-dicarbonyls (e.g., methylglyoxal + 2-butene-1,4-dial) and ~25% of the time to form organic nitrates. The major uncertainties in the mechanisms concern (1) the structure of the bicyclicperoxy intermediate, and (2) the γ-dicarbonyl photooxidation mechanism. Good fits to the o-cresol concentration–time profiles in the toluene–NOx runs are obtained if it is assumed that o7-cresol reacts rapidly with NO3 radicals. However, it is observed that the model underpredicts nitrotoluene yields by a factor of ~10, but this is in any case a minor product. It is concluded that further experimental work will be required toadequately validate the assumptions incorporated in the aromatic photooxidation mechanisms presented here.  相似文献   

15.
The oxidation of soot on catalysts with the perovskite and fluorite structures (including platinum-promoted catalysts) in the presence and in the absence of NO2 was studied using in situ IR spectroscopy and temperature-programmed techniques (TPR, TPD, and TPO). It was found that, as a rule, the temperature of the onset of soot oxidation considerably decreased upon the addition of NO2 to a flow of O2/N2, whereas the amount of oxygen consumed in soot oxidation considerably increased. To explain these facts, we hypothesized that the initiation of soot combustion in the presence of NO2 was related to the activation of the NO2 molecule through the formation (at a low temperature) and decomposition (at a high temperature) of nitrate structures on the catalyst. Superequilibrium amounts of NO2 resulted from the decomposition of nitrate complexes immediately on the catalyst for soot combustion. Based on a comparison between catalyst activities and data obtained by TPR and the TPD of oxygen, a conclusion was drawn that the presence of labile oxygen in the catalyst is a necessary but insufficient condition for the efficient occurrence of a soot oxidation reaction in the presence of NO2. The introduction of platinum as a constituent of the catalyst increased the amount of labile oxygen and, as a consequence, increased the amount of highly reactive nitrate complexes. As a result, this caused a decrease in the temperature of the onset of soot combustion.  相似文献   

16.
While environmental chamber data have been widely used to generate and validate computer models of the chemistry occurring in polluted atmospheres, the effects of the chambers on the gas-phase chemistry being studied have been poorly characterized. In order to investigate such chamber effects, a series of NOx—air irradiations, with trace levels of organics present to monitor OH radical concentrations, have been carried out in four different environmental chambers (ranging in volume from ~100 to 40,000 L) at varying temperatures, humidities, pressures, and reaction conditions. In addition, a number of control experiments have been carried out to validate the technique for measuring OH radical levels in these irradiations. The data show that unknown sources of OH radicals are present in all of the chambers studied. The data are consistent with the presence of two distinct radical sources: (1) the photolysis of initially present HONO, whose importance increases with increasing NO2/NO concentration ratios, but which is a minor contributor to the overall radical flux after 30–60 min of irradiation, and (2) a constant (for these NOx—air irradiations) radical source which dominates beyond approximately the first 60 min of irradiation. The radical input rates, after the first ∽30–60 min of irradiation, are independent of the NO concentration, increase with increasing temperature, humidity, and NO2 concentration, are proportional to light intensity, and are dependent on the chamber employed. Although the exact nature of this radical source is still undetermined, results of experiments reported here allow a number of possible mechanisms to be ruled, out, and these are discussed.  相似文献   

17.
A two-channel thermal dissociation cavity ring down spectroscopy (CRDS) instrument has been built for in situ, real-time measurement of NO2 and total RNO2 (peroxy nitrates and alkyl nitrates) in ambient air, with a NO2 detection limit of 0.10 ppbv at 1 s. A 6-day long measurement was conducted at urban site of Hefei by using the CRDS instrument with a time resolution of 3 s. A commercial molybdenum converted chemiluminescence (Mo-CL) instrument was also used for comparison. The average RNO2 concentration in the 6 days was measured to be 1.94 ppbv. The Mo-CL instrument overestimated the NO2 concentration by a bias of +1.69 ppbv in average, for the reason that it cannot distinguish RNO2 from NO2. The relative bias could be over 100% during the afternoon hours when NO2 was low but RNO2 was high.  相似文献   

18.
The kinetics and nitroarene product yields of the gas-phase reactions of naphthalene-d8, fluoranthene-d10, and pyrene with OH radicals in the presence of NOx and in N2O5? NO3? NO2? air mixtures have been investigated at 296 ± 2 K and atmospheric pressure of air. Using a relative rate method, naphthalene-d8 was shown to react in N2O5? NO3? NO2? air mixtures a factor of 1.22 ± 0.10 times faster than did naphthalene, with the 1- and 2-nitronaphthalene-d7 product yields being similar to those of 1- and 2-nitronaphthalene from naphthalene. From the measured PAH concentrations and the nitroarene product yields, formation yields of 2-, 7-, and 8-nitrofluoranthene-d9 and 2- and 4-nitropyrene of 0.03, 0.01, 0.003, 0.005, and 0.0006, respectively, were determined from the OH radical-initiated reactions. Effective rate constants for the reactions of fluoranthene-d10 and pyrene with N2O5 in N2O5? ;NO3? NO2? air mixtures of ca. 1.8 × 10?17 cm3 molecule?1 s?1 and ca. 5.6 × 10?17 cm3 molecule?1 s?1, respectively, were derived. Formation yields of 2-nitrofluoranthene-d9 and 4-nitropyrene of ca. 0.24 and ca. 0.0006, respectively, were estimated for these reaction systems. 2-Nitropyrene was also observed to be formed in these N2O5? NO3? NO2 reactions, but was found to be a function of the NO2 concentration and, therefore, would be a negligible product under ambient NO2 concentrations. These product and kinetic data are consistent with ambient air measurements of the nitroarene concentrations.  相似文献   

19.
The kinetics and mechanism of the oxidation of carbon by NO2 in absence and presence of water vapor were studied in a fixed bed reactor. The rate of carbon oxidation by NO2 is enhanced in the presence of water vapor in the range of temperature 300–400°C. The benefit effect of water is attributed to the intermediate formation of traces of nitric and nitrous acids, which enhance the rate of the carbon oxidation without modifying the global mechanism reaction. Therefore, water acts as a catalyst for the carbon oxidation by NO2. A kinetic mechanism derived from this parametric study shows a decrease in the activation energy of carbon oxidation by NO2 in the presence of water vapor. This result is in agreement with the experimental observation. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 236–244, 2009  相似文献   

20.
Oxidation of benzene into CO2 in air has been studied in the temperature range of 300-673 K on supported metal oxides and on 0.7% Pt/Al2O3, in the absence and in the presence of several Volatile Organic Compounds (VOCs). On the metal oxides, the deep oxidation of benzene is strongly inhibited in presence of VOCs and O-containing VOCs lead to more toxic VOCs (i.e: acetaldehyde).  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