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.
Co-firing methane (CH4) and ammonia (NH3) has attracted growing concerns as a feasible greenhouse gas reduction strategy in gas turbine-based power generation, which raises the need to better understand the interaction of methane and nitric oxide (NO) under flame conditions. In this work, laminar flame propagation of CH4/NO mixtures at initial pressure (Pu) of 1 atm, initial temperature (Tu) of 298 K and equivalence ratios of 0.4–1.8 was experimentally investigated using a constant-volume combustion vessel. Laminar burning velocities (LBVs) and Markstein lengths were experimentally determined. A kinetic model of CH4/NO combustion was developed with rate constants of several important reactions updated, presenting reasonable predictions on the measured LBVs of CH4/NO mixtures. The modeling analyses reveal that the reduction of NO can proceed through two mechanisms, i.e. the hydrocarbon NO reduction mechanism and non-hydrocarbon NO reduction mechanism. Among the two mechanisms, the non-hydrocarbon NO reduction mechanism which includes reactions NO+H = N+OH, NO+O = N + O2 and NO+N = N2+O has a higher contribution to NO reduction at the equivalence ratio of 0.6, while the hydrocarbon NO reduction mechanism with hydrocyanic acid (HCN) as the key intermediate plays a more important role at the equivalence ratio of 1.8. NO+H = N+OH and CH3+NOHCN+H2O are found to be the two most sensitive reactions to promote the flame propagation, while the LBVs measured in this work are demonstrated to provide strong constraint for these reactions. Furthermore, previous CH4/O2/NO oxidation data measured in flow reactor and rapid compression machine were also simulated, which provides extended validation of the present model over wider conditions.  相似文献   

2.
A detailed chemical kinetic model for oxidation of CH3CHO at intermediate to high temperature and elevated pressure has been developed and evaluated by comparing predictions to novel high-pressure flow reactor experiments as well as shock tube ignition delay measurements and jet-stirred reactor data from literature. The flow reactor experiments were conducted with a slightly lean CH3CHO/O2 mixture highly diluted in N2 at 600–900 K and pressures of 25 and 100 bar. At the highest pressure, the oxidation of CH3CHO was in the NTC regime, controlled to a large extent by the thermal stability and reactions of peroxide species such as HO2, CH3OO, and CH3C(O)OO. Model predictions were generally in good agreement with the experimental data, even though the predicted temperature for onset of reaction was overpredicted at 100 bar. This discrepancy was attributed mainly to uncertainties in the CH3C(O)OO reaction subset. Predictions of ignition delays in shock tubes and species profiles in JSR experiments were also satisfactory. At temperatures above the NTC regime, acetaldehyde ignition and oxidation is affected mainly by the competition between dissociation of CH3CHO and reaction with the radical pool, and by reactions in the methane subset.  相似文献   

3.
Infrared absorption cross sections for acetonitrile (methyl cyanide; CH3CN) have been determined in the 880–1700 cm?1 spectral region from spectra recorded using a high-resolution FTIR spectrometer (Bruker IFS 125 HR) and a multipass cell with a maximum optical pathlength of 19.3 m. Spectra of acetonitrile/dry synthetic air mixtures were recorded at 0.015 cm?1 resolution (calculated as the Bruker instrument resolution of 0.9/MOPD) at a number of temperatures between 203 and 297 K and pressures appropriate for atmospheric conditions. Intensities were calibrated using three composite acetonitrile spectra recorded at the Pacific Northwest National Laboratory. These absorption cross sections will provide an accurate basis for upper tropospheric/lower stratospheric retrievals of acetonitrile in the mid-infrared spectral region from ACE satellite data.  相似文献   

4.
Thermal desorption spectrometry (TDS) and electron stimulated desorption (ESD) are employed to investigate mechanisms responsible for the formation of C2H6 in electron irradiated multilayer films of acetonitrile (CH3CN) at 30 K. Using a high sensitivity time-of-flight mass spectrometer, we observe the ESD of anionic fragments H, CH2 , CH3 and CN. Desorption occurs following dissociative electron attachment (DEA) via several negative ion resonances in the 6 to 14 eV energy range and correlates well with a “resonant” structure seen in the TDS yield of C2H6 (i.e., at mass 30 amu). It is proposed that C2H6 is formed by the reactions of CH3 radicals generated following DEA to CH3CN which also yields CN. Between 2 and 5 eV, a second resonant feature is seen in the C2H6 signal. While DEA is observed in the gas phase at these energies, no anion desorption occurs since anionic fragments likely have insufficient kinetic energy to desorb. Since the CH2 ion has not been observed in gas-phase measurements, we propose that it is formed, along with HCN (that is detected in TDS) when dissociation into CH3 and CN is hindered by adjacent molecules.  相似文献   

5.
Larger ethers such as diethyl ether (DEE) and di-n-propyl ether (DPE) have different oxidation behavior (double-NTC behavior) compared to the simplest dimethyl ether (DME). Such phenomena are interpreted with different reactions and processes in different ether kinetic models, which also predict different formation pathways of oxidation intermediates such as acids. To gain further insights into the oxidation kinetics of linear ethers, ethyl methyl ether (EME), which has a nonsymmetrical structure, was studied in this work. Oxidation experiments of 1% of EME were performed in a jet-stirred reactor at 1 atm, a residence time of 2 s, an equivalence ratio of 1, and over a temperature range of 375–850 K. The intermediates were analyzed with photoionization molecular-beam mass spectrometry. To explain the oxidation behavior of EME, a detailed kinetic model was also constructed. The oxidation of EME spans a wider temperature range than DME, but no obvious double-NTC behavior was observed as DEE. Based on the model analysis and profiles of critical intermediates such as ketohydroperoxides (KHPs) and CH3O2H, the low-temperature oxidation behavior of EME was explained by the chain-branching reactions of the fuel itself and the oxidation intermediates. Abundant species such as aldehydes, acids, esters, and fuel-specific dione species were detected and could be well reproduced by the current model. In particular, acids are produced by the decomposition of KHPs and subsequent reactions of the intermediate CH3CHO. Esters and dione species are mainly formed via fuel-related pathways.  相似文献   

6.
Raman spectra of dilute solutions of acetonitrile in ionic liquids reveal the characteristic features of ionic liquids' polarity. This is accomplished by investigating the Raman bandshape of the ν (CN) band, corresponding to the CN stretching mode of CH3CN, which is a very sensitive probe of the local environment. The amphiphilic nature of the CH3CN molecule allows us to observe the effect of electron pair acceptor and electron pair donor characteristics on ionic liquids. It has been found that the overall polarity of nine different ionic liquids based on 1‐alkyl‐3‐methylimidazolium cations is more dependent on the anion than cation. The observed wavenumber shift of the ν (CN) band of CH3CN in ionic liquids containing alkylsulfate anions agrees with the significant different values previously measured for the dielectric constant of these ionic liquids. The conclusions obtained from the analysis of the ν (CN) band were corroborated by the analysis of the symmetric ν1 (CD3 ) stretching mode of deuterated acetonitrile in different ionic liquids. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

7.
The redundancy-free internal valence force field (RFIVFF) of acetonitrile is reported using CNDO/force method. The initial force field is set up by taking the interaction and bending force constants from CNDO force field and transferring stretching force constants from the force fields of chemically related molecules. The final force field is obtained by refining the initial force field using vibrational harmonic frequencies of CH3CN,13CH3CN, CH3 13CN, CH3C15N, CD3CN and CD3 13CN. The final force field thus obtained is found to be excellent on the basis of frequency fit and potential energy distribution.  相似文献   

8.
Oxidation of bulk samples of tungsten (923 K) and zirconium (773 and 873 K) by H2O/CO2 supercritical fluid (molar ratio [CO2]/[H2O] = 0.17–0.26) at a pressure of about 300 atm is investigated. Oxidation produces monoclinic WO3, monoclinic W19O55, monoclinic ZrO2, H2, CO, CH4, and carbon (on the surface of tungsten oxide). Differences in oxidation mechanisms for tungsten and zirconium are revealed. CO2 molecules take part in the oxidation of tungsten only after oxide formation in reaction with H2O. Zirconium is oxidized fully, and oxidation of tungsten terminates in the formation of the oxide layer at the metal surface.  相似文献   

9.
The B‐band resonance Raman spectra of 2(1H)‐pyridinone (NHP) in water and acetonitrile were obtained, and their intensity patterns were found to be significantly different. To explore the underlying excited state tautomeric reaction mechanisms of NHP in water and acetonitrile, the vibrational analysis was carried out for NHP, 2(1D)‐pyridinone (NDP), NHP–(H2O)n (n = 1, 2) clusters, and NDP–(D2O)n (n = 1, 2) clusters on the basis of the FT‐Raman experiments, the B3LYP/6‐311++G(d,p) computations using PCM solvent model, and the normal mode analysis. Good agreements between experimental and theoretically predicted frequencies and intensities in different surrounding environments enabled reliable assignments of Raman bands in both the FT‐Raman and the resonance Raman spectra. The results indicated that most of the B‐band resonance Raman spectra in H2O was assignable to the fundamental, overtones, and combination bands of about ten vibration modes of ring‐type NHP–(H2O)2 cluster, while most of the B‐band resonance Raman spectra in CH3CN was assigned to the fundamental, overtones, and combination bands of about eight vibration modes of linear‐type NHP–CH3CN. The solvent effect of the excited state enol‐keto tautomeric reaction mechanisms was explored on the basis of the significant difference in the short‐time structural dynamics of NHP in H2O and CH3CN. The inter‐molecular and intra‐molecular ESPT reaction mechanisms were proposed respectively to explain the Franck–Condon region structural dynamics of NHP in H2O and CH3CN.Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

10.
A new kind of continuous-wave (CW) cold molecular beam, methyl cyanide (CH3CN) beam, is generated by a bent electrostatic quadrupole guiding. The Stark shift of rotational energy levels of CH3CN molecule and its population distribution are calculated, and the dynamic processes of electrostatic guiding and energy filtering of CH3CN molecules from a gas source with room temperature (300 K) are simulated by Monte Carlo Method. The study showed that the longitudinal and transversal temperatures of output cold CH3CN beam could be about ~2 K and ~ 420 mK, and the corresponding guiding efficiency was about 10?5 as the guiding voltage was 3 kV. Furthermore, the temperature of the guided molecules and its guiding efficiency can be controlled by adjusting the guiding voltages applied on electrodes.  相似文献   

11.
Mechanism and kinetics of NH2OH + OOH and NH2CH3 + OOH reactions were studied at the B3LYP and M062X levels of theory using the 6-311++G(3df, 3pd) basis set. The NH2OH + OOH and NH2CH3 + OOH reactions proceed through different paths which lead to different products. Transition state structure and activation energy of each path were calculated. The calculated activation energies of hydrogen abstraction reactions were smaller than 25 kcal/mol and of substitution reactions are in the range of 50–70 kcal/mol. The rate constants were calculated using transition state theory (TST) modified for tunneling effect at 273–2000 K.  相似文献   

12.
A new kind of continuous-wave (CW) cold molecular beam, methyl cyanide (CH3CN) beam, is generated by a bent electrostatic quadrupole guiding. The Stark shift of rotational energy levels of CH3CN molecule and its population distribution are calculated, and the dynamic processes of electrostatic guiding and energy filtering of CH3CN molecules from a gas source with room temperature (300 K) are simulated by Monte Carlo Method. The study showed that the longitudinal and transversal temperatures of output cold CH3CN beam could be about ∼2 K and ∼ 420 mK, and the corresponding guiding efficiency was about 10−5 as the guiding voltage was 3 kV. Furthermore, the temperature of the guided molecules and its guiding efficiency can be controlled by adjusting the guiding voltages applied on electrodes.   相似文献   

13.
The low-temperature auto-ignition chemistry of isopropyl nitrate (iPN) was experimentally and numerically investigated in the present study. The ignition delay times (IDTs) of iPN were measured stoichiometrically over a temperature range of 560–600 K at effective pressures of 5 and 10 bar in a rapid compression machine. A two-stage ignition phenomenon of iPN was observed. Both the first-stage IDTs and total IDTs vary rapidly within the narrow temperature range investigated (∼40 K). A recent iPN kinetic mechanism proposed by Fuller and Goldsmith for pyrolysis studies was extended. The reaction kinetics of CH3CHO + NO2 has been theoretically calculated at 500–1500 K and 0.01–100 atm. The rate information of CH3 + NO2 was updated based on previous theoretical results. The O2-addition channel of acetyl radical (CH3CO), which accounts for the first-stage IDT, was also considered in the present work. The extended iPN kinetic model predicts the two-stage IDTs well. Simulation results suggest that the IDTs are most sensitive to the following two reactions: (1) CH3 + NO2 = CH3O + NO; (2) CH3 + NO2 = CH3NO2. The former promotes the overall reactivity by yielding the reactive methoxy radical, while the latter forms a relatively stable product (i.e., CH3NO2). The reaction of CH3CHO + NO2 = CH3CO + HONO supplements the formation of CH3CO. The different consumption channels of CH3CO radicals (the O2-addition reaction and the decomposition reaction) lead to different chain reactions yielding OH radicals with increasing temperature in the ignition process. The “NONO2 loop” is the main route for OH formation in the studied conditions, which is mainly responsible for the iPN ignition.  相似文献   

14.
The triplet excited states of 2-aminopurine (2-APu) and 2-N, N-dimethylaminopurine (2-DMAPu) in acetonitrile and water have been detected and characterized. Comparison of the initial T-T absorbancies, together with the measured singlet lifetime, in these systems allowed us to determine the relative intersystem crossing rate constants. The ratios kISC (2-APu/CH3CN) : kISC (2-APu/H2O) and kISC (2- DMAPu/CH3CN) were found to be 36.7 and 11.2, respectively. Comparison with the ratios calculated from the luminescence data confirm the previously proposed [1] scheme of electronic levels in these purines.  相似文献   

15.
Product formation in laser-photolytic Cl-initiated low-temperature (550–700 K) oxidation of isobutane in a slow-flow reactor was investigated by tunable synchrotron photoionization mass spectrometry. These experiments probed the time-resolved formation of products following photolytic initiation of the oxidation, and identify isomeric species by their photoionization spectra. The relative yields of oxygenated product isomers (2,2-dimethyloxirane, methylpropanal, and 3-methyloxetane) are in reasonable concord with measurements from Walker and co-workers (J. Chem. Soc. Faraday Trans. 74 (1) (1978) 2229–2251) at higher temperature. Oxidation of isotopically labeled isobutane, (CH3)3CD, suggests that methylpropanal formation can proceed from both (CH3)2CCH2OOH and CH3CH(CH2)CH2OOH isomers. Bimodal time behavior is observed for product formation; the initial prompt formation reflects “formally direct” channels, principally chemical activation, and the longer-timescale “delayed” component arises from dissociation of thermalized ROO and QOOH radicals. The proportion of prompt to delayed signal is smaller for the oxygenated products than for the isobutene product. This channel-specific behavior can be qualitatively understood by considering the different energetic distributions of ROO and QOOH in formally direct vs. thermal channels and the fact that the transition states involved in the formation of oxygenated products are “tighter” than that for isobutene formation.  相似文献   

16.
Acetonitrile (CH3CN) coordination to a Pt(111) surface has been studied with electron energy loss vibrational spectroscopy (EELS), XPS, thermal desorption and work function measurements. We compare data for the surface states with known acetonitrile coordination complexes. For CH3CN adsorbed on Pt(111) at 100 K, the molecule is rehybridized and adsorbs with the CN bond parallel or slightly inclined to the surface plane in an η2(C, N) configuration. The ν(CN) frequency is 1615 cm?1 and the C ls and N ls binding energies are 284.6 eV and 397.2 eV respectively. By contrast, weakly adsorbed multilayer acetonitrile exhibits a ν(CN) vibrational frequency of 2270 cm?1, and C ls and N ls binding energies of 286.9 eV and 400.1 eV respectively. Both the EELS and XPS results are consistent with rehybridization of the CN triple bond to a double bond with both C and N atoms of the CN group attached to the surface. In addition to this majority η2(C, N) monolayer state, evidence is found for a second, more strongly bound minority molecular state in thermal desorption spectra. As a result of the low coverage of this state, EELS was unable to spectroscopically identify it and we tentatively assign it as an η4(C, N) species associated with accidental step sites. By contrast to the surface complexes, almost all of the known platinum-nitrile coordination complexes are end-bonded via the N lone-pair orbital. Several cases of side-on bonding are known, however, and we compare the results with the known complex Fe32-NCCH3)(CO)9. The difference in the coordinative properties of a Pt(111) surface versus a single Pt atom must be due to the increased ability of multi-atom arrays to back-donate electrons into the π1 system of acetonitrile. Previously published EELS and XPS results for monolayer acetonitrile on Ni(111) and polycrystalline films are almost identical to the present results on Pt(111). We believe that the monolayer of CH3CNNi(111) is also an η2(C, N) species, not an end-bonded species previously proposed by Friend, Muetterties and Gland.  相似文献   

17.
Methanol (CH3OH) and formaldehyde (HCHO) reacting with nitrogen dioxide (NO2) contribute to the largest uncertainty for the CH3OH/NOx low temperature combustion mechanism. CH3OH and NO2 only undergo H-abstraction reactions, while HCHO + NO2 involves multiple reaction channels, among which H-abstraction dominates. In the present work, a high level quantum chemical method, CCSD(T)/aug-cc-pVQZ//M06–2X-D3/6-311++G(d,p), was employed to investigate the reaction pathways. The reaction kinetics were explored by RRKM/master equation simulations with multidimensional small-curvature tunneling (SCT) corrections and hindered rotor approximations. The H-abstraction reactions with barriers higher than 20 kcal/mol indicate a nonnegligible quantum tunneling effect even under combustion conditions. Our computations predict the tunneling factors to be 3–4 for the studied reactions at 500 K. A significant tunneling effect is also expected for H-abstraction of large alcohols and aldehydes by NO2. The computed total rate coefficients show good agreement with previous experimental measurements over narrow ranges of temperature and pressure, ensuring the accuracy of the reported branching ratios covering a wide T, P range for the two reactions. The results of CH3OH + NO2 reveal the dominant role of HONOcis + CH2OH. It's also uncovered the dominance of HONOcis + CHO pathway in HCHO + NO2 under the studied conditions. The detailed reaction kinetics information reported in this work is useful for building rate rules for the mechanisms of other nitrogen-containing alcohol-based fuels.  相似文献   

18.
Methanol (CH3OH) has attracted considerable attention as a renewable fuel or fuel additive with low greenhouse gas emissions. Methanol oxidation was studied using a recently developed supercritical pressure jet-stirred reactor (SP-JSR) at pressures of 10 and 100 atm, at temperatures from 550 to 950 K, and at equivalence ratios of 0.1, 1.0, and 9.0 in experiments and simulations. The experimental results show that the onset temperature of CH3OH oxidation at 100 atm is around 700 K, which is more than 100 K lower than the onset at 10 atm and this trend cannot be predicted by the existing kinetics models. Furthermore, a negative temperature coefficient (NTC) behavior was clearly observed at 100 atm at fuel rich conditions for methanol for the first time. To understand the observed temperature shift in the reactivity and the NTC effect, we updated some key elementary reaction rates of relevance to high pressure CH3OH oxidation from the literature and added some new low-temperature reaction pathways such as CH2O + HO2 = HOCH2O2 (RO2), RO2 + RO2 = HOCH2O (RO) + HOCH2O (RO) + O2, and CH3OH + RO2 = CH2OH + HOCH2O2H (ROOH). Although the model with these updates improves the prediction somewhat for the experimental data at 100 atm and reproduces well high-temperature ignition delay times and laminar flame speed data in the literature, discrepancies still exist for some aspects of the 100 atm low-temperature oxidation data. In addition, it was found that the pressure-dependent HO2 chemistry shifts to lower temperature as the pressure increases such that the NTC effect at fuel-lean conditions is suppressed. Therefore, as shown in the experiments, the NTC phenomenon was only observed at the fuel-rich condition where fuel radicals are abundant and the HO2 chemistry at high pressure is weakened by the lack of oxygen resulting in comparatively little HO2 formation.  相似文献   

19.
Michael A. Henderson 《Surface science》2010,604(19-20):1800-1807
The photochemical properties of the Cr-terminated α-Cr2O3(0001) surface were explored using methyl bromide (CH3Br) as a probe molecule. CH3Br adsorbed and desorbed molecularly from the Cr-terminated α-Cr2O3(0001) surface without detectable thermal decomposition. Temperature programmed desorption (TPD) revealed a CH3Br desorption state at 240 K for coverages up to 0.5 ML, followed by more weakly bound molecules desorbing at 175 K for coverages up to 1 ML. Multilayer exposures led to desorption at ~ 130 K. The CH3Br sticking coefficient was unity at 105 K for coverages up to monolayer saturation, but decreased as the multilayer formed. In contrast, pre-oxidation of the surface (using an oxygen plasma source) led to capping of surface Cr3+ sites and near complete removal of CH3Br TPD states above 150 K. The photochemistry of chemisorbed CH3Br was explored on the Cr-terminated surface using post-irradiation TPD and photon stimulated desorption (PSD). Irradiation of adsorbed CH3Br with broad band light from a Hg arc lamp resulted in both photodesorption and photodecomposition of the parent molecule at a combined cross section of ~ 10? 22 cm2. Photodissociation of the CH3–Br bond was evidenced by both CH3 detected in PSD and Br atoms left on the surface. Use of a 385 nm cut-off filter effectively shut down the photodissociation pathway but not the parent molecule photodesorption process. From these observations it is inferred that d-to-d transitions in α-Cr2O3, occurring at photon energies < 3 eV, do not significantly promote photodecomposition of adsorbed CH3Br. It is unclear to what extent band-to-band versus direct CH3Br photolysis play in CH3–Br bond dissociation initiated by more energetic photons.  相似文献   

20.
The kinetics of the C6H5 reactions with CH3OH and C2H5OH has been measured by pulsed-laser photolysis/mass-spectrometry (PLP/MS) employing acetophenone as the radical source. Kinetic modeling of the benzene formed in the reactions over the temperature range 306–771 K allows us to reliably determine the total rate constants for H-abstraction reactions. In order to improve our low temperature measurements down to 304 K we have also applied the cavity ring-down spectrometric technique using nitrosobenzene as the radical source. Both sets of data agree closely. A weighted least-squares analysis of the two complementary sets of data for the two reactions gave the total rate constants k(CH3OH) = (7.82 ± 0.44) × 1011 exp [?(853 ± 30)/T] and k(C2H5OH) = (5.73 ± 0.58) × 1011 exp [?(1103 ± 44)/T] cm3 mol?1 s?1 for the temperature range studied. Theoretically, four possible product channels of the C6H5 + CH3OH reaction producing C6H6 + CH3O, C6H6 + CH2OH, C6H5OH + CH3 and C6H5OCH3 + H and five possible product channels of the C6H5 + C2H5OH reaction producing C6H6 + C2H5O, C6H6 + CH2CH2OH, C6H6 + CH3CHOH, C6H5OH + CH3CH2 and C6H5OCH2CH3 + H have been computed at the G2M//B3LYP/6?311+G(d, p) level of theory. The hydrogen abstraction channels were predicted to have lower energy barriers than those for the substitution reactions and their rate constants were calculated by the microcanonical variational transition state theory at 200–3000 K. The predicted rate constants are in good agreement with the experimental values. Significantly, the rate constant for the CH3OH reaction with C6H5 was found to be greater than that for the C2H5OH reaction and both reactions were found computationally to be dominated by H-abstraction from the hydroxyl group attributable to the affinity of the phenyl toward the OH group and the predicted lower energy barriers for the OH attack.  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