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 31 毫秒
1.
Two laminar, premixed, fuel-rich flames fueled by anisole-oxygen-argon mixtures with the same cold gas velocity and pressure were investigated by molecular-beam mass spectrometry at two synchrotron sources where tunable vacuum-ultraviolet radiation enables isomer-resolved photoionization. Decomposition of the very weak O–CH3 bond in anisole (C6H5OCH3) by unimolecular decomposition yields the resonantly-stabilized phenoxy radical (C6H5O). This key intermediate species opens reaction routes to five-membered ring species, such as cyclopentadiene (C5H6) and cyclopentadienyl radicals (C5H5). Anisole is often discussed as model compound for lignin to study the phenolic-carbon structure in this natural polymer. Measured temperature profiles and mole fractions of many combustion intermediates give detailed information on the flame structure. A very comprehensive reaction mechanism from the literature which includes a sub-scheme for anisole combustion is used for species modeling. Species with the highest measured mole fractions (on the order of 10?3–10?2) are CH3, CH4, C2H2, C2H4, C2H6, CH2O, C5H5 (cyclopentadienyl radical), C5H6 (cyclopentadiene), C6H6 (benzene), C6H5OH (phenol), and C6H5CHO (benzaldehyde). Some are formed in the first destruction steps of anisole, e.g., phenol and benzaldehyde, and their formation will be discussed and with regard to the modeling results. There are three major routes for the fuel destruction: (1) formation of benzaldehyde (C6H5CHO), (2) formation of phenol (C6H5OH), and (3) unimolecular decomposition of anisole to phenoxy (C6H5O) and CH3 radicals. In the experiment, the phenoxy radical could be measured directly. The phenoxy radical decomposes via a bicyclic structure into the soot precursor C5H5 and CO. Formation of larger oxygenated species was observed in both flames. One of them is guaiacol (2-methoxyphenol), which decomposes into fulvenone. The presented speciation data, which contain more than 60 species mole fraction profiles of each flame, give insights into the combustion kinetics of anisole.  相似文献   

2.
Shock-tube and flow-reactor experiments were used to study the thermal decomposition of diethyl carbonate (C2H5OC(O)OC2H5; DEC). The formation of CO2, C2H4, and C2H5OH was measured with gas chromatography/mass spectrometry (GC/MS) and high-repetition-rate time-of-flight mass spectrometry (HRR-TOF-MS) behind reflected shock waves. The same products were also detected by GC/MS in flow reactor experiments. All experiments combined span a temperature range of 663–1203 K at pressures between 1.0 and 2.0 bar. Time-resolved species concentration profiles from HRR-TOF-MS and product compositions from GC/MS measurements were simulated applying a detailed reaction mechanism for DEC combustion. A master-equation analysis was conducted based on computed energies from G4 calculations. Quantum chemical calculations confirm that DEC primarily decomposes by six-center elimination, C2H5OC(O)OC2H5 → C2H4 + C2H5OC(O)OH (1a), followed by rapid decomposition of the alkoxy acid, C2H5OC(O)OH → C2H5OH + CO2 (1b). Measured DEC decomposition rate constants k(T) at p ≈ 1.5 bar can be represented by the Arrhenius equation k(T) = 1013.64±0.12 exp(?204.24±1.95 kJ/mol/RT) s ? 1. Theoretical predictions for k1a were in good agreement with experimentally derived values. The theoretical analysis also included dipropyl carbonate (C3H7OC(O)OC3H7; DPC) decomposition and the reactivities of DEC and DPC are compared and discussed in the context of reactivity of dialkyl carbonates under pyrolytic conditions.  相似文献   

3.
We employed tunable diode laser absorption spectroscopy to measure the line strength, the methane (CH4), ethane (C2H6) and the propane (C3H8) broadening coefficients for the 523–422 H2O transition at 3619.61 cm?1. Water amount fractions generated by a stable and accurate humidity transfer standard, traceable to the SI units via the German national humidity standard, were used to calibrate the spectroscopic line strength measurements. We focus on the traceability of the measured line data to the SI and on uncertainty assessments following the guidelines of the Guide to the Expression of Uncertainty in Measurement. We determined the line strength to be (8.42 ± 0.07)×10?20 cm?1/(cm?2 molecule) corresponding to a relative uncertainty of ±0.8%. To the best of our knowledge, we report the first methane, ethane and propane broadening coefficients of (8.037 ± 0.056)×10?5 cm?1/hPa, (9.077 ± 0.064)×10?5 cm?1/hPa and (10.469 ± 0.073)×10?5 cm?1/hPa for the 523–422 H2O transition at 3619.61 cm?1, respectively. The relative combined uncertainties of the stated CH4, C2H6 and C3H8 broadening coefficients are in the ±0.7% range.  相似文献   

4.
5.
The chemisorption of C2H2 and C2H4 on a clean or partly C- or O-covered Fe(111) surface was investigated with AES, TDS and HREELS. On the clean surface, both molecules adsorb under strong rehybridization close to sp3. Above 230 K, C2H2 reacts to form CH and presumably CH2 as the main products, which on further heating decompose to yield H2 desorption maxima at 580 and 490 K, leaving two carbon species on the surface which correspond to two loss peaks at 400 and 1290 cm?1 in the HREELS spectrum. C2H4 undergoes very rapid decomposition above 250 K; no intermediates have been detected. The presence of coadsorbed oxygen or carbon atoms only reduced the maximum uptake of C2H2, but led to the appearance of new molecular adsorption states of C2H4 and inhibited C2H4 decomposition.  相似文献   

6.
L. Sun  Y. Chang  S. Tang  Z. Wang 《Molecular physics》2013,111(23-24):2945-2949
Geometries, relative energies, and stabilities of endo- and exohedral complexes, X@Si20H20 and XSi20H20, (X = H+, H, N, P, C?, and Si?) are calculated at B3LYP/6-31G* level. The energy minimum structure of Si20H21 + shows that the proton cannot be positioned in the Si20H20 centre, but prefers attach to Si20H20 exohedrally with C2v symmetry. Most investigated Ih endohedral complexes X@Si20H20 (X = H, N, P, C?, and Si?) are local minima, except for 2N@Si20H20, which is a high-order saddle point. Inclusions energies of the endohedral complexes are calculated, and it reveals that energy penalties caused by encapsulation are rather small. Exohedral complexes XSi20H20 (X = H, N, P, C?, and Si?) have C2v or Cs local minima, and most of them are more stable than their endohedral isomers with the exception of C2v 4PSi20H20 and 4Si?Si20H20.  相似文献   

7.
Single-pulse shock-tube experiments were used to study the thermal decomposition of selected oxygenated hydrocarbons: Ethyl propanoate (C2H5OC(O)C2H5; EP), propyl propanoate (C3H7OC(O)C2H5; PP), isopropyl acetate ((CH3)2HCOC(O)CH3; IPA), and methyl isopropyl carbonate ((CH3)2HCOC(O)OCH3; MIC) The consumption of reactants and the formation of stable products such as C2H4 and C3H6 were measured with gas chromatography/mass spectrometry (GC/MS). Depending on the considered reactant, the temperatures range from 716–1102 K at pressures between 1.5 and 2.0 bar. Rate-coefficient data were obtained from first-order analysis. All reactants primarily decompose by six-center eliminations: EP → C2H4 + C2H5COOH (propionic acid); PP → C3H6 + C2H5COOH; IPA → C3H6 + CH3COOH (acetic acid); MIC → C3H6 + CH3OC(O)OH (methoxy formic acid). Experimental rate-coefficient data can be well represented by the following Arrhenius expressions: k(EP → products) = 1013.49±0.16 exp(−214.95±3.25 kJ/mol/RT) s−1; k(PP → products) = 1012.21±0.16 exp(–191.21±2.79 kJ/mol/RT) s−1; k(IPA → products) = 1013.10±0.31 exp(–186.38±5.10 kJ/mol/RT) s−1; k(MIC → products) = 1012.43±0.29 exp(–165.25±4.46 kJ/mol/RT) s−1. The determination of rate coefficients was based on the amount of C2H4 or C3H6 formed. The potential energy surface (PES) of the thermal decomposition of these four reactants was determined with the G4 composite method. A master-equation analysis was conducted based on energies and molecular properties from the G4 computations. The results indicate that the length of a linear alkyl substituent does not significantly influence the rate of six-center eliminations, whereas the change from a linear to a branched alkyl substituent results in a significant reactivity increase. The comparison between rate-coefficient data also shows that alkyl carbonates have higher reactivity towards decomposition by six-center elimination than esters. The results are discussed in in the context of reactivity patterns of carbonyl compounds.  相似文献   

8.
The formation and properties of J-aggregates in thin solid films of pseudoisocyanines with long N-alkyl groups, obtained by centrifuging from solutions in organic solvents, were studied. It is shown for the first time that nonsymmetric cyanine dyes, containing a C2H5 group at one nitrogen atom and a C10H21, C15H31, or C18H37 group at another nitrogen atom, spontaneously form J-aggregates stable at room temperature and pressing a narrow absorption band with a half-width at half maximum of 200 cm?1. The thermal stability of J-aggregates in thin films of pseudoisocyanines with alkyl substituents decreases in the following order: C2H5-C2H5> C2H5-C6H13>C2H5-C18H37>C2H5-C10H21>C2H5-C15H31. By introducing 1-octadecyl-2-methylquinolinium iodide in the film, it was found that the J-aggregates studied consist of a small number (2–4) of dye molecules.  相似文献   

9.
The kinetics of the reaction of β‐substituted β‐alkoxyvinyl trifluoromethyl ketones R1O‐CR2?CH‐COCF3 ( 1a – e ) [( 1a ), R1?C2H5, R2?H; ( 1b ), R1?R2?CH3; ( 1c ), R1?C2H5, R2?C6H5; ( 1d ), R1?C2H5, R2?V?pNO2C6H4; ( 1e ), R1?C2H5, R2?C(CH3)3] with four aliphatic amines ( 2a – d ) [( 2a ), (C2H5)2NH; ( 2b ), (i‐C3H7)2NH; ( 2c ), (CH2)5NH; ( 2d ), O(CH2CH2)2NH] was studied in two aprotic solvents, hexane and acetonitrile. The least reactive stereoisomeric form of ( 1a – d ) was the most populated ( E‐s‐Z‐o‐Z ) form, whereas in ( 1e ), the more reactive form ( Z‐s‐Z‐o‐Z ) dominated. The reactions studied proceeded via common transition state formation whose decomposition occurred by ‘uncatalyzed’ and/or ‘catalyzed’ route. Shielding of the reaction centre by bulky β‐substituents lowered abruptly both k′ (‘uncatalyzed’ rate constant) and k″ (‘catalyzed’ rate constant) of this reaction. Bulky amines reduced k″ to a greater extent than k′ as a result of an additional steric retardation to the approach of the bulky amine to its ammonium ion in the transition state. An increase in the electron‐withdrawing ability of the β‐substituent increased ‘uncatalyzed’ k′ due to the acceleration of the initial nucleophile attack (k1) and ‘uncatalyzed’ decomposition of transition state (k2) via promoting electrophilic assistance (through transition state 8 ). The amine basicity determined the route of the reaction: the higher amine basicity, the higher k3/k2 ratio (a measure of the ‘catalyzed’ route contribution as compared to the ‘uncatalyzed’ process) was. ‘Uncatalyzed’ route predominated for all reactions; however in polar acetonitrile the contribution of the ‘catalyzed’ route was significant for amines with high pKa and small bulk. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

10.
正庚烷热裂解的反应分子动力学模拟   总被引:1,自引:0,他引:1  
研究表明正庚烷热裂解的主产物是C2H4, H2, CH4以及C3H6,模拟结果和实验吻合很好. 温度对产物分布具有明显的影响,当温度上升,目标产物乙烯的量会迅速增加. 正庚烷转化率以及主产物的摩尔分数分别通过反应分子动力学和化学动力学模拟计算得到,两种方法模拟结果相吻合. 我们还通过动力学分析研究了正庚烷热裂解反应的动力学参数,反应活化能为47.32 kcal/mol,指前因子为1.78×1014 s-1.  相似文献   

11.
The far-infrared rotational spectra of chlorotrifluoromethane, dichlorodifluoromethane, and trichlorofluoromethane have been observed with an interferometric (Fourier transform) spectrometer in the region 10–40 cm?1 at a resolution of 0.07 cm?1. CCl2F2 exhibits a continuum spectrum at this resolution, but symmetric top rotational fine structure is observed for CClF3 and CCl3F. Isotope splitting is also observed in CClF3, and analysis yields the rotational constants for C35ClF3 of B0 = 0.11112 cm?1, DJ = 1.6 × 10?8cm?1; and for C37ClF3, B0 = 0.10835 cm?1, DJ = 1.5 · 10?8cm?1. Isotopic shifts can be allowed for in CCl3F to yield constants for C35Cl3F of B0 = 0.0821 cm?1, DJ = 1 × 10?8cm?1. These values are all in agreement with those deduced from microwave studies of the low J transitions apart from B0 for C35ClF3, where the difference is outside the expected experimental error.  相似文献   

12.
In Ar and Ar/N2 radio frequency (RF) discharges with admixtures of aluminium tri‐isopropoxide (ATI) the fragmentation of this metal‐organic precursor was studied by means of Fourier Transform Infrared (FTIR) spectroscopy using an optical long path cell providing an optical length of l = 17.2 m. The experiments were performed in an asymmetric capacitively coupled process plasma at a frequency of f = 13.56 MHz and at pres‐sure values in the range of p = 1–10.5 Pa. The discharge power was chosen between P = 10–100 W. Using FTIR spectroscopy the evolution of the concentrations of ATI and of six stable molecules, CH4, C2H2, C2H4, C2H6, CO and HCN, was monitored under flowing conditions at gas flows of Φtotal = 0.5–14.5 sccm in the discharge. The concentrations of the reaction products were measured tobe between 2 x 1012 molecules cm–3 as e.g. found for C2H4and C2H6, and 5 x 1013 molecules cm–3, as e.g. in the case of CO. In the plasma a complete dissociation of the precursor ATI was found at a power value of about P = 80 W independent on the admixture of Ar or N2. The fragmentation efficiency (FE) of the reaction products which originate from the ATI molecules ranges between 0.2 and 4 x 1016 molecules J–1 while the fragmentation rate (FR) reached up to 2.5 x 1018 molecules s–1. The multi component detection ability of the spectrometer served to analyse the carbon balance of the by‐product formation. For all experiments, the carbon balance never exceeded 25%. Therefore, in the plasmas the majority of the provided carbon is most likely deposited at the reactor walls or forms dust particles or higher molecular CxHy. The conversion efficiencies (CE) of the produced molecular species ranges between 0.1 x 1015 molecules J–1 for C2H4 and 5 x 1015 molecules J–1 for C2H6 depending on the discharge conditions of the RF plasma. (© 2014 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

13.
《Surface science》1987,180(1):1-18
Thermal programmed desorption (TPD), high resolution electron energy loss spectroscopy (HREELS), and time-resolved laser-induced desorption (LID) have been used to study the chemisorption and decomposition of ethylene over Ni(100). Ethylene adsorbs molecularly on this surface at temperatures below 150 K. The molecule is π bonded in this state, showing very little rehybridization. At coverages below half saturation, decomposition to vinyl plus a hydrogen atom occurs unimolecularly with a rate constant of (8.0 ± 2.0) × 10−2 s−1 at 170 K. A strong kinetic isotope effect was observed; vinyl formation from C2D4 does not occur until about 200 K. The proposal of vinyl as the intermediate is supported by studies with C2H4, 1,1− and 1,2−C2D2H2, and C2D4. The reaction is slower at saturation coverages, where molecular desorption is still seen above 200 K. Vinyl decomposes further at 230 K to form an acetylenic fragment.  相似文献   

14.
Kinetic study has been performed to understand the reactivity of novel cationic gemini surfactants viz. alkanediyl‐α,ω‐bis(hydroxyethylmethylhexadecylammonium bromide) C16‐s‐C16 MEA, 2Br? (where s = 4, 6) in the cleavage of p‐nitrophenyl benzoate (PNPB). Novel cationic gemini C16‐s‐C16 MEA, 2Br? surfactants are efficient in promoting PNPB cleavage in presence of butane 2,3‐dione monoximate and N‐phenylbenzohydroxamate ions. Model calculation revealed that the higher catalytic effect of ethanol moiety of gemini surfactants (C16H33N+ C2H4OH CH3 (CH2)S N+ C2H4OH CH3C16H33, 2Br?, s = 4, 6) is due to their higher binding capacity toward substrate. This is in line with finding that binding constants for novel series of cationic gemini surfactants are higher than conventional cationic gemini (C16H33N+(CH3)2(CH2)SN+(CH3)2C16H33, 2Br?, s = 10, 12), cetyldimethylethanolammonium bromide and zwitterionic surfactants, i.e. CnH2n+1N+Me2 (CH2)3 SO3? (n = 10; SB3‐10). The fitting of kinetic data was analyzed by the pseudophase model. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
The absorption cross sections of N2, O2, CO, NO, CO2, N2O, CH4, C2H4, C2H6, C4H10 have been measured photoelectrically in the 180–700 Å region using synchrotron radiation. The absorption cross sections in the region λ ≥ 500 Å was found to be structureless and to increase monotonically with wavelength for all gases. The positions of the structure observed in the 520–720 Å region for N2, O2, CO2 and N2O are consistent with the various Rydberg series reported by previous authors.  相似文献   

16.
The kinetics of the thermal decomposition of mono-, di-, and tripotassium salts of trinitrophloroglucinol (TNPG) was studied in the solid phase by the manometric method at m/V from 1 × 10?4 to 15 × 10?4 g/cm3 over the temperature ranges 125–150°C (K1TNPG), 200–230°C (K2 TNPG), and 200–250°C (K3 TNPG). All the compounds decomposed according to the topochemical mechanism: there was an induction period, after which the rate of gas release was maximum. This rate then gradually decreased. The second decomposition stage was observed for K1TNPG as the temperature increased to 200–250°C. The special features of changes in the rate of the process during transformation and the influence of the degree of vessel filling with a substance, particle size, and temperature on the kinetics of decomposition were studied. The kinetic results and composition of gaseous products and some condensed decomposition products lead to certain conclusions concerning the mechanism of the chemical transformations.  相似文献   

17.
A Fourier transform interferometer was used to record the slit-jet cooled absorption spectrum of 12C2H4 between 700 and 2400 cm?1I, at a spectral resolution of 0.005 cm?1. Three bands, v12 at 1442.442 70(1)cm?1, v7 + v8 at 1888.978 23(3)cm?1 and v6 + v10 at 2047.775832(2)cm?1, were rotationally analysed. In the case of 7181, a known Coriolis perturbation mechanism involving the nearby 4181 (1958.264cm?1) and 81101 (1766.391 cm?1) states was accounted for in the analysis. The latter fitting procedure included 12 levels from the 4181 state which are observed because lines from v4 + v8 borrow intensity from v7 + v8. Compared to the literature, significantly improved vibration-rotation constants were obtained for all upper states reported in the present study.  相似文献   

18.
Fujun Li  Caiping Liu  Chunwan Liu 《Molecular physics》2013,111(17-18):2251-2257
A series of mixed-metal carbonyl clusters, MonIr4-n(μ-CO)3(CO)9-n5-Cp)n (n?=?1, 2; Cp?=?C5H4Me, C5HMe4, C5Me5, C5H5), have been investigated using the TDDFT method. The results estimated the medium magnitude of the first static hyperpolarizabilities (β tot ?~?2?×?10?30esu) of these tetrahedral mixed-metal clusters, most of which originate from the intra-cluster charge transfer of the metal skeleton consisting of polar metal–metal interactions. No direct contributions of the Cp ligands to the relevant charge transfer were found, but the cooperative effect between the metal centres and Cp ligands impacts the β. Based on these studies a mixed-metal cluster Mo3Ir(μ-CO)3(CO)75-Cp)2 was designed that exhibits enhanced first hyperpolarizability.  相似文献   

19.
A detailed comparison has been conducted between chemiluminescence (CL) species profiles of OH?, CH?, and C2 ?, obtained experimentally and from detailed flame kinetics modeling, respectively, of atmospheric pressure non-premixed flames formed in the forward stagnation region of a fuel flow ejected from a porous cylinder and an air counterflow. Both pure methane and mixtures of methane with hydrogen (between 10 and 30 % by volume) were used as fuels. By varying the air-flow velocities methane flames were operated at strain rates between 100 and 350 s?1, while for methane/hydrogen flames the strain rate was fixed at 200 s?1. Spatial profiles perpendicular to the flame front were extracted from spectrograms recorded with a spectrometer/CCD camera system and evaluating each spectral band individually. Flame kinetics modeling was accomplished with an in-house chemical mechanism including C1–C4 chemistry, as well as elementary steps for the formation, removal, and electronic quenching of all measured active species. In the CH4/air flames, experiments and model results agree with respect to trends in profile peak intensity and position. For the CH4/H2/air flames, with increasing H2 content in the fuel the experimental CL peak intensities decrease slightly and their peak positions shift towards the fuel side, while for the model the drop in mole fraction is much stronger and the peak positions move closer to the fuel side. For both fuel compositions the modeled profiles peak closer to the fuel side than in the experiments. The discrepancies can only partly be attributed to the limited attainable spatial resolution but may also necessitate revised reaction mechanisms for predicting CL species in this type of flame.  相似文献   

20.
The temporal variation of chemiluminescence emission from OH?(A2 Σ +) and CH?(A2 Δ) in reacting Ar-diluted H2/O2/CH4, C2H2/O2 and C2H2/N2O mixtures was studied in a shock tube for a wide temperature range at atmospheric pressures and various equivalence ratios. Time-resolved emission measurements were used to evaluate the relative importance of different reaction pathways. The main formation channel for OH? in hydrocarbon combustion was studied with CH4 as benchmark fuel. Three reaction pathways leading to CH? were studied with C2H2 as fuel. Based on well-validated ground-state chemistry models from literature, sub-mechanisms for OH? and CH? were developed. For the main OH?-forming reaction CH+O2=OH?+CO, a rate coefficient of k 2=(8.0±2.6)×1010 cm3?mol?1?s?1 was determined. For CH? formation, best agreement was achieved when incorporating reactions C2+OH=CH?+CO (k 5=2.0×1014 cm3?mol?1?s?1) and C2H+O=CH?+CO (k 6=3.6×1012exp(?10.9 kJ?mol?1/RT) cm3?mol?1?s?1) and neglecting the C2H+O2=CH?+CO2 reaction.  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