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 34 毫秒
1.
A comprehensive analysis of hydrogen/oxygen and hydrocarbon/oxygen counterflow diffu-sion flames has been conducted using corresponding detailed reaction mechanisms. The hydrocarbon fuels contain n-alkanes from CH4 to C16H34. The basic diffusion flame struc-tures are demonstrated, analyzed, and compared. The effects of pressure, and strain rate on the flame behavior and energy-release rate for each fuel are examined systematically. The de-tailed chemical kinetic reaction mechanisms from Lawrence Livermore National Laboratory(LLNL) are employed, and the largest one of them contains 2115 species and 8157 reversible reactions. The results indicate for all of the fuels the flame thickness and heat release rate correlate well with the square root of the pressure multiplied by the strain rate. Under the condition of any strain rate and pressure, H2 has thicker flame than hydrocarbons, while the hydrocarbons have the similar temperature and main products distributions and almost have the same flame thickness and heat release rate. The result indicates that the fuels composed with these hydrocarbons will still have the same flame properties as any pure n-alkane fuel.  相似文献   

2.
The influence of preliminary γ-irradiation and γ-irradiation during hte oxidation process on the kinetics of heat release in the systemsn-decane—aqueous solution of HNO3 and a solution of tributyl phosphate in a kerosene—aqueous solution of HNO3 was studied. The preliminary γ-irradiation of the system at 43°C increases the initial rate of the process (k 1). The increase is proportional to the irradiation dose at doses up to 150kGy, then the increase ink 1 is retarded, and the further course of the process becomes practically independent of the irradiation dose. The effect of γ-irradiation during the oxidation depends on the temperature of the system: at temperatures lower than 80 °C, γ-irradiation increases the rate of heat release, whereas at higher temperatures, γ-irradiation decreases the rate of heat release. The effects observed were explained by the competition of NO2 accumulation due to the radiolysis of nitric acid and processes of the addition of NO2 to unsaturated hydrocarbons produced by the radiolysis of the organic phase. Translated fromIzvestiya Akademii Nauk, Seriya Khimicheskaya, No. 6, pp. 1116–1120, June, 1998.  相似文献   

3.
A photo-triggered discharge is used to study the influence of three hydrocarbons (HCs), propene (C3H6), n-decane (C10H22), and toluene (C6H5CH3) on NO conversion in N2/O2/NO/HC mixtures, with 18.5% O2 concentration, 700 ppm of NO, and an hydrocarbon concentration ranging between 190 ppm and 2,700 ppm. The electrical system generates a transient homogeneous plasma, working under 400 mbar total pressure, with a 50 ns short current pulse at a repetition frequency up to a few Hz. The NO concentration at the exit of the reactor is quantified using absolute FTIR spectroscopy measurements, as a function of the specific deposited energy in the discharge and the mixture composition. Owing to the plasma homogeneity, the experimental results can be compared to predictions of a self-consistent 0-D discharge and kinetic model based on available data in the literature about reactions and their rate constants. It is shown that the addition of either propene (as for DBD or corona discharges) or n-decane to N2/O2/NO leads to an improvement of the NO removal as compared to the mixture without hydrocarbon molecules. The adopted kinetic schemes explain this effect for the two mixture types. On the other hand, both the experiments and model predictions emphasize that the addition of toluene does not lead to the improvement of NO conversion. Moreover, compounds that are useful for NO x reduction catalysis, such as aldehydes, are less produced in the mixture with toluene.  相似文献   

4.
Kinetic modeling of non-thermal plasma chemistry is conducted to investigate hydrocarbon (CH4, C2H4, C3H6, and C3H8) effects on the promotion of NO–NO2 conversion. A reduced plasma chemistry model, in which radical reactions are selectively involved, is validated with experimental data. The higher reactivity of hydrocarbon additive with O radicals, which produces initial radicals, is requisite to initiate hydrocarbon decomposition, thus providing NO–NO2 conversion. Initial radicals by plasma discharge induce continual hydrocarbon decomposition and this self-preserved reaction mechanism greatly contributes to the promotion of energy efficient NO–NO2 conversion. Increase in the conversion extent by ethylene and propylene additives is substantial because of their stronger affinity with O radical. The primary routes of NO–NO2 conversion process differed by hydrocarbon additives are presented and discussed with the assistance of sensitivity analysis.  相似文献   

5.
The kinetic regularities of the thermal decomposition of dinitramide in aqueous solutions of HNO3, in anhydrous acetic acid, and in several other organic solvents were studied. The rate of the decomposition of dinitramide in aqueous HNO3 is determined by the decomposition of mixed anhydride of dinitramide and nitric acid (N4O6) formed in the solution in the reversible reaction. The decomposition of the anhydride is a reason for an increase in the decomposition rates of dinitramide in solutions of HNO3 as compared to those in solutions in H2SO4 and the self-acceleration of the process in concentrated aqueous solutions of dinitramide. The increase in the decomposition rate of nondissociated dinitramide compared to the decomposition rate of the N(NO2)2 anion is explained by a decrease in the order of the N−NO2 bond. The increase in the rate constant of the decomposition of the protonated form of dinitramide compared to the corresponding value for neutral molecules is due to the dehydration mechanism of the reaction. For Part 1, see Ref. 1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 41–47, January, 1998.  相似文献   

6.
The gas-phase reactions of O . (H2O)n and OH(H2O)n, n=20–38, with nitrogen-containing atmospherically relevant molecules, namely NOx and HNO3, are studied by Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometry and theoretically with the use of DFT calculations. Hydrated O . anions oxidize NO . and NO2 . to NO2 and NO3 through a strongly exothermic reaction with enthalpy of −263±47 kJ mol−1 and −286±42 kJ mol−1, indicating a covalent bond formation. Comparison of the rate coefficients with collision models shows that the reactions are kinetically slow with 3.3 and 6.5 % collision efficiency. Reactions between hydrated OH anions and nitric oxides were not observed in the present experiment and are most likely thermodynamically hindered. In contrast, both hydrated anions are reactive toward HNO3 through proton transfer from nitric acid, yielding hydrated NO3. Although HNO3 is efficiently picked-up by the water clusters, forming (HNO3)0–2(H2O)mNO3 clusters, the overall kinetics of nitrate formation are slow and correspond to an efficiency below 10 %. Combination of the measured reaction thermochemistry with literature values in thermochemical cycles yields ΔHf(O(aq.))=48±42 kJ mol−1 and ΔHf(NO2(aq.))=−125±63 kJ mol−1.  相似文献   

7.
Fuel decomposition and hydrocarbon growth processes of methyl tert‐butyl ether (MTBE) and related alkyl ethers have been studied experimentally in soot‐producing nonpremixed flames. Temperature, C1–C12 hydrocarbons, and major species were measured in coflowing methane/air flames whose fuel was separately doped with 5000 ppm of MTBE, n‐butyl methyl ether (NBME), sec‐butyl methyl ether (SBME), ethyl tert‐butyl ether (ETBE), and tert‐amyl methyl ether (TAME; =1,1‐dimethylpropyl methyl ether). The consumption rates of the dopants, several simple kinetic calculations, and the dependence of the observed products on fuel composition indicate that the dominant decomposition process was unimolecular dissociation, not H‐atom abstraction. The dominant dissociations were four‐center elimination of alcohols for the doubly branched ethers (MTBE, ETBE, and TAME) and C? O fission for the linear ether (NBME), while four‐center elimination and C? O fission were comparably important for the singly branched ether (SBME). These dissociations produced alkenes which further reacted to produce alkadienes/alkynes, alkenynes, acetylenic compounds, and aromatics. The dependence of the maximum benzene mole fractions on fuel composition was consistent with benzene formation through reactions of highly‐unsaturated C3 and/or C4 hydrocarbons (C3H3, n‐C4H3, C4H4, n‐C4H5, etc.). © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 345–358, 2004  相似文献   

8.
The rates of heat release in the nitrogen dioxide—n-decane system at a molar ratio of nitrogen oxides ton-decane (β) from 2.4·10−3 to 3.1 and gaseous volumes per mole ofn-decane (V(g)) equal to 0.05–4.5 were studied in the 55.2–92.8 °C temperature range. The initial rate of the process is determined by the interaction of NO2 withn-decane. The equilibrium constants of dissociation of N2O4 inn-decane and Henry's constants of NO2 and N2O4 in ann-decane solution were determined by complex analysis of the thermodynamic equilibrium in the NO2n-decane system and dependences of the initial rates onV(g) and β. The experimentally observed self-acceleration of the process in the region of high β and lowT values was suggested to be due to the reaction of N2O4 with intermediate oxidation products. The rate constants of the reaction of NO2 withn-decane were compared with analogous values determined in its mixtures with HNO3 solutions. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1789–1794, October, 1997.  相似文献   

9.
A model of core mechanism of hydrocarbon pyrolysis with good predictive ability is crucial to the development of active cooling technology for advanced aeroengines. In this work, a detailed core kinetic model of pyrolysis of C1–C4 hydrocarbon fuels is developed through the combination of a series of potential energy surfaces and validated against a series of experimental results. The kinetic model contains 103 species and 1290 reactions, and most of the kinetic and thermochemical parameters are compiled from recent highly accurate quantum chemical calculations without modification. The pressure-dependent rate constants are considered for the dissociation/association reactions, isomerization reactions, and chemically activated reactions. Simulation results for various alkanes (methane, ethane, propane, n-butane, isobutane), alkenes (ethylene, propene, 1-butene, 2-butene, isobutene, allene, 1,3-butadiene), and alkynes (acetylene, propyne, vinylacetylene) indicate that the major product distributions at various temperatures (800-2300 K) and pressures (0.8-10 atm) can be predicted well by the developed core kinetic model. Thus, the developed pyrolysis mechanism for C1–C4 hydrocarbons can be used as a cornerstone to develop the pyrolysis mechanisms of larger hydrocarbon fuels and thus support the development of thermal management in advanced aeroengines.  相似文献   

10.
The kinetics of ethylene oxidation by PdCl2 and CrO3 complexes supported on silica gel (300 K, closed batch reactor) and the adsorption of C2H4 by silica gel and metal complex reaction centers (M n ) were studied. A new version of the kinetic distribution method was applied to determine the rate constants of ethylene reactions with metal complexes with consideration for the equilibrium distribution of C2H4 among the reactor gas phase, silica gel, and M n . The rate constant of a first-order reaction with respect to Cr(VI) (k e) remained constant as [M n ] was increased up to 0.15 mol % with the absence of detectable ethylene adsorption by chromium(VI). In the case of Pd(II)/SiO2, strong ethylene adsorption by palladium(II) was found, and k e was an exponential function of [M n ]. This exponential function is indicative of an increase in the specific activity of Pd(II) with palladium concentration on SiO2. Taking into account the adsorption of ethylene (physisorption on SiO2 and chemisorption on Pd(II)), we found an analogy between the kinetic behaviors of Pd(II) in reactions with ethylene on silica gel and with ethylene and other hydrocarbons in solutions.  相似文献   

11.
Phase diagrams were studied for (R4N)2[Nd(NO3)5]-C n H2n + 2-chloroform liquid ternaries (where R4N+ is trialkylbenzylammonium and n = 10, 12, 14, or 15) at T = 298.15−333.15 K. (R4N)2[Nd(NO3)5]-C n H2n + 2 binaries are two-phase liquid systems at all temperatures in this range. One phase (phase I) is an almost pure hydrocarbon solvent. The other (phase II) is enriched in (R4N)2[Nd(NO3)5]. The C n H2n + 2 solubility in phase II decreases with increasing the alkyl chain length of the hydrocarbon solvent and increases with increasing temperature. The title liquid ternaries are characterized by a homogeneous solution field and a two-phase liquid solution field. One phase is enriched in (R4N)2[Nd(NO3)5] and chloroform; the other is enriched in the hydrocarbon solvent. Liquid-liquid phase separation fields enlarge with increasing C n H2n + 2 alkyl chain length and slightly narrow with increasing temperature. Critical solution points at fixed temperatures depend on C n H2n + 2. Original Russian Text ? A.K. Pyartman, V.A. Keskinov, P.V. Zaitsev, 2009, published in Zhurnal Neorganicheskoi Khimii, 2009, Vol. 54, No. 3, pp. 531–534.  相似文献   

12.
The positive ion–molecule reactions of OCS have been investigated in an ion cyclotron resonance spectrometer. A variety of reactions in OCS/hydrocarbon mixtures have been investigated for various C1? C4 hydrocarbons—alkanes, alkenes and alkynes. The formation of organosulfur ions is found in reactions in OCS/hydrocarbon (Cn) mixtures with n <4. Formation of organosulfur ions is observed from hydrocarbon ions reacting with OCS and [OCS]+˙ and S+˙ reacting with the hydrocarbons. The proton affinity of OCS has been determined to be 688.7±8 kJ mol?1 while that of CS2 is measured to be 712.1±8 kJ mol?1. Comparison with the proton affinity of CO2 shows that the proton affinity increases as sulfur is substituted for oxygen.  相似文献   

13.
The rate constants and modes of reaction of NO2+ and C2H5ONO2NO2+ with aromatic compounds and alkanes have been determined in a pulsed ion cyclotron resonance mass spectrometer. Both ions undergo competing charge transfer and substitution reactions (NO2+ + M → MO+ + NO; C2H5ONO2NO2+ + M → MNO2+ + C2H5ONO2) with aromatic molecules. In both cases, the probability that a collision results in charge transfer increases with increasing exothermicity of that process. The C2H5ONO2NO2+ ion does not undergo charge transfer with molecules having an ionization potential greater than about 212 kcal/mol (9.2 eV); this observation leads to an estimate of 13 kcal/mol for the binding energy between NO2+ and C2H5ONO2. The importance of the substitution reaction depends on the number of substituents on the aromatic ring and the molecular structure, and, in the case of C2H5ONO2NO2+ ions, on the energetics of the competing charge transfer process. Both NO2+ and C2H5ONO2NO2+ undergo hydride transfer reactions with alkanes. For both these ions, k(hydride transfer)/k (collision) increases with increasing exothermicity of reaction, but in both cases the rate constants of reaction are unusually low when compared with other hydride transfer reactions of comparable exothermicity which have been reported in the literature. This is interpreted as evidence that the attack on the alkane preferentially involves the nitrogen atom (where the charge is localized) rather than one of the oxygen atoms of NO2+.  相似文献   

14.
Peroxynitrates (RO2NO2), in particular acyl peroxynitrates (R = R′C(O) with R′ = alkyl), are prominent constituents of polluted air. In this work, a systematic study on the thermal decomposition rate constants of the first five members of the series of homologous R′C(O)O2NO2 with R′ = CH3 ( =PAN), C2H5, n‐C3H7, n‐C4H9, and n‐C5H11 is undertaken to verify the conclusions from previous laboratory data (Grosjean et al., Environ. Sci. Technol. 1994, 28, 1099–1105; Grosjean et al., Environ. Sci. Technol. 1996, 30, 1038–1047; Bossmeyer et al., Geophys. Res. Lett. 2006, 33, L18810) that the longer chain peroxynitrates may be considerably more stable than PAN. Experiments are performed in a temperature‐controlled, evacuable 200 L‐photoreactor made from quartz. n‐Acyl peroxynitrates are generated by stationary photolysis of mixtures of molecular bromine, O2, NO2, and the corresponding parent aldehydes, highly diluted in N2. Thermal decomposition of R′C(O)O2NO2 is initiated by the addition of an excess of NO. First‐order decomposition rate constants k1 of the reactions R′C(O)O2NO2 (+M) → R′C(O)O2 + NO2 (+M) are derived at 298 K and a total pressure of 1 bar from the measured loss rates of R′C(O)O2NO2, correcting for wall loss of R′C(O)O2NO2 and several percentages of reformation of R′C(O)O2NO2 by the reaction of R′C(O)O2 radicals with NO2. With increasing chain length of R′, k1(298 K) slightly decreases from 4.4 × 10?4 s?1 (R′ = CH3) to 3.7 × 10?4 s?1 (R′ = C2H5), leveling off at (3.4 ± 0.1) × 10?4 s?1 for R′ = n‐C3H7, n‐C4H9, and n‐C5H11. Temperature dependencies of k1 were measured for CH3C(O)O2NO2 and n‐C5H11C(O)O2NO2 in the temperature range 289–308 K, resulting in the same activation energy within the statistical error limits (2σ) of 0.9 and 1.5 kJ mol?1, respectively. A few experiments on n‐C6H13C(O)O2NO2, n‐C7H15C(O)O2NO2, and n‐C8H17C(O)O2NO2 were also performed, but the results were considered to be unreliable due to strong wall loss of the peroxynitrate and possible complications caused by radical‐sinitiated side reactions.  相似文献   

15.
The doubly-charged ion mass spectra of some hydrocarbons, including a variety of structural types, have been obtained by a new technique in which doubly-charged ions are charge exchanged with neutral molecules and so separated from singly-charged ions. The spectra show strong similarities, independent of hydrocarbon structure; characteristic ions include [CmH2]++ (m = 2 to 5), [CnH6]++(n > 6), [C10H8]++, [C12H8]++, [C11H10]++, [C7H7]++·, [C9H7]++· and [C13H11]++·. The fragmentation pattern of 2-phenylnaphthalene has been reconstructed, based on observed reactions of metastable doubly-charged ions to give fragment doubly-charged ions. In addition, we examined metastable ion fragmentations leading to two singly-charged ions for some of the characteristic ions, using several compounds. The value of doubly-charged ion mass spectra of hydrocarbons appears to lie in the information they provide on ion structures; this information was sufficient to permit the proposal of structures for the major ions encountered in this study.  相似文献   

16.
Pulsed laser photolysis, time-resolved laser-induced fluorescence experiments have been carried out on the reactions of CN radicals with CH4, C2H6, C2H4, C3H6, and C2H2. They have yielded rate constants for these five reactions at temperatures between 295 and 700 K. The data for the reactions with methane and ethane have been combined with other recent results and fitted to modified Arrhenius expressions, k(T) = A′(298) (T/298)n exp(?θ/T), yielding: for CH4, A′(298) = 7.0 × 10?13 cm3 molecule?1 s?1, n = 2.3, and θ = ?16 K; and for C2H6, A′(298) = 5.6 × 10?12 cm3 molecule?1 s?1, n = 1.8, and θ = ?500 K. The rate constants for the reactions with C2H4, C3H6, and C2H2 all decrease monotonically with temperature and have been fitted to expressions of the form, k(T) = k(298) (T/298)n with k(298) = 2.5 × 10?10 cm3 molecule?1 s?1, n = ?0.24 for CN + C2H4; k(298) = 3.4 × 10?10 cm3 molecule?1 s?1, n = ?0.19 for CN + C3H6; and k(298) = 2.9 × 10?10 cm3 molecule?1 s?1, n = ?0.53 for CN + C2H2. These reactions almost certainly proceed via addition-elimination yielding an unsaturated cyanide and an H-atom. Our kinetic results for reactions of CN are compared with those for reactions of the same hydrocarbons with other simple free radical species. © John Wiley & Sons, Inc.  相似文献   

17.
In addition to the known C11H16 hydrocarbons multifidene ( 4 ), aucantene ( 2 ), and ectocarpene ( 5 ), the marine brown alga Cutleria multifida produces trace amounts of the C9H12 hydrocarbon 7-melhylcycloocta-1,3,5-triene ( 8 ) and its valence tautomer 7-methylbicyclo[4.2.0]octa-2,4-diene, A second novel C9H12 hydrocarbon is 6-vinyicyclo-hepta-1,4-diene ( 9 ), a lower homologue of ectocarpene ( 5 ). Among the C11H16 hydrocarbons, 7-((1E/Z)-prop-l-enyl)cycloocta-1,4-diene ( 10 / 11 ) is found for the first time. The structure of all new products is confirmed by synthesis and spectroscopic data. The biosynthesis of the new hydrocarbons 8 – 11 is obviously linked to the pathways which lead to the major products giffordene ( 7 ), (6S)-ectocarpene ((6S)- 5 ), and (4R,5R)-aucantene ((4R,5R)- 2 ). Consecutive reactions of certain thermolabile primary products proceed via electrocyclic ring closure, 3,3-sigmatropic rearrangement, or a 1,7-sigmatropic H-shift.  相似文献   

18.
The conditions favorable for the formation of oligomer and benzene cation radicals upon C2H4, C6H6, and C3H6 adsorption on thermally activated HZSM-5 and CuZSM-5 zeolites were studied. In the case of isolated copper ions, the radicals were shown to be formed from C3H6 but not from C2H4 and C6H6. Paramagnetic nitroxy complexes, which were formed as a result of the interaction of adsorbed hydrocarbons C2H4, C6H6, and C3H6 with NO and NO2, were revealed on HZSM-5 zeolite and Al2O3. Hydrocarbon radicals and nitroxy complexes are localized on different zeolite sites. Nitroxy complexes are formed on aluminum cations. The role of radicals and nitroxy complexes in the process of NO x selective catalytic reduction by hydrocarbons on the zeolites was discussed on the basis of the temperature dependences of the concentrations of these species.  相似文献   

19.
Although oxygenated fuel additives are effective in reducing soot emissions, the extent to which molecular structure of the oxygenate plays a role in soot reduction has remained unclear and controversial. To gain a deeper insight in this field, a detailed chemical kinetic modeling approach was used to examine the phenomenon of suppression of sooting by the addition of oxygenated hydrocarbon species to the fuel. For this task, the PREMIX code in conjunction with Chemkin II and models resulting from the merging of validated kinetic schemes describing the oxidation of the components of the n-butanol-benzene mixtures were used to investigate the effect of n-butanol addition on the formation?depletion of acetylene recognized as soot precursor in flames under fuel-rich conditions. The first part of this study treats the dependence of the soot precursor amounts on n-butanol percentage in the fuel mixture, whereas the second part defines the key reaction mechanisms responsible for the observed reduction in C2H2 and consequently in polycyclic aromatic hydrocarbons and soot amounts induced by the oxygenate additive. The principal objective of the current study was to obtain fundamental understanding of the mechanisms through which the oxygenate compound affects the soot precursor amounts. The modeling results indicated that there was a dramatic decrease in the acetylene peak height with the addition of the oxygenated addtitive. This finding was found to be due to the increase in the C2H2 consumption rates induced by n-butanol addition. Finally, the modeling results provided evidence that n-butanol played a role in changing acetylene formation mechanism by enhancing the role of C3H4P, C3H4 and aC3H5 and by eliminating the role of C6H4, C5H5, C5H6, H2CCCCH, C4H2, C5H4O, C2H, CHCHCHO, H2C4O and C4H4.  相似文献   

20.
Gaseous H2NO3+ ions are generated by direct protonation of HNO3 by H3+, H3O+ and CH5+ and by protonation of C2H5ONO2 followed by C2H4 loss.  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