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 218 毫秒
1.
The internal dynamics of the 12C13CH3+ isotopomer of the simplest carbocation is considered to be a distorted dynamics of its symmetric C2H3+ isotopomer. By using the group-chain methods and assuming quite naturally that the carbocation equilibrium structures continue to be planar and the nonrigid motion occurring between them is retained upon this isotopic substitution, a simple algebraic model is constructed to describe the spectrum of the 12C13CH3+ isotopomer. The correctness of the model is limited only by that of the adopted geometrical symmetry of internal dynamics of the C2H3+ isotopomer.  相似文献   

2.
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   

3.
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   

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.
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   

6.
Abstract

The electron spin resonance of γ-irradiated single crystals of methoxycarbonylcholine picrate hemihydrate, C7H16NO3 + · C6H2N3O7 ? · ½ H2O has been observed and analyzed for different orientations of the crystal in the magnetic field. The crystals have been investigated between 70 and 350 K. The spectra were found to be temperature independent and the radiation damage centers are attributed to – ?[Obar]OCH3 and –CH2CH2O? radicals. The g and hyperfine coupling constants were found to be almost isotropic with an average, g = 2.0060, a H1 = 4.4G for –CH2CH2O?, g = 2.0050, a H2 = 3.5G for –CH2CH2O? and g = 2.0045, a H = 3.5 G for –?[Obar]OCH3. These values indicate a long-range coupling between the unpaired electron and H protons.  相似文献   

7.
A near-infrared tunable diode laser spectrometer called TDLAS has been developed that combines telecommunication-type as well as new-generation antimonide laser diodes to measure C2H2, H2O, CO2 and their isotopologues in the near infrared. This sensor is devoted to the in situ analysis of the soil of the Martian satellite PHOBOS, within the framework of the Russian space mission PHOBOS-GRUNT. In the first part of the paper, we report accurate spectroscopic measurements of C2H2 and 13C12CH2 near 1.533 μm, of H2O and CO2 at 2.682 μm and of the isotopologues 13C16O2 and 16O12C18O near 2.041 μm and H2 17O, H2 18O and HDO near 2.642 μm. The achieved line strengths are thoroughly compared to data from molecular databases or from former experimental determinations. In the second part of the paper, we describe the TDLAS spectrometer for the PHOBOS-GRUNT mission.  相似文献   

8.
Benzophenone ((C6H5)2CO) and decafluorobenzophenone ((C6F5)2CO) were applied to elucidate the photochemical reaction pathway of hydrogen peroxide (H2O2) with dimethylsulfoxide (DMSO). When a solution of benzophenone in DMSO was excited with the 355 nm laser light, three transient species were observed in the time-resolved electron paramagnetic resonance spectra: benzophenone ketyl (C6H5)2COH, methylCH3, and methylsulfinic methylCH2SOCH3 radicals. However, when decafluoro-benzophenone was used with DMSO, only ketyl and methylsulfinic methyl radicals were observed under the same experimental conditions. When the reaction of benzophenone and DMSO was carried out in the presence of H2O2, different time profiles ofCH3 radicals were observed. In the reaction of decafluorobenzophenone-DMSO-H2O2, the time profiles of the radicals were not affected by the presence of H2O2. Thus, these results verify thatCH3 radicals are regenerated in a cyclic pathway, in whichCH3 radicals attack H2O2. The regeneration pathway allows us to observe f-pair polarization throughout the lifetime ofCH3 radicals, which last several microseconds, an order of magnitude longer than theT 1 relaxation time ofCH3 radicals.  相似文献   

9.
10.
The methanesulfonic acid (MSA)-diethylamine (DEA) binary liquid system is studied over the entire range of compositions at 30°C by using multiple frustrated total internal reflection IR spectroscopy. Solutions with acid: base equimolar ratio contain only 1 : 1 ion pairs. Upon adding the acid, a MSA molecule abstracts an anion from the 1 : 1 complex to produce a protonated DEA and an (H3C(O2)SO…H…OS(O2)CH3) anion with a strong H bond: (C2H5)2(H)NH+ · OS(O2)CH3 + HOS(O2)CH3 ↔ (C2H5)2(H)NH+ + (H3C(O2)SO…H…OS(O2)CH3). This equilibrium is shifted to the left. The 1 : 1 complex is present in solutions even at an significant excess of the acid. To protonate the complex, it is required at least two MSA molecules. Under conditions of an excess of the base, DEA molecules do not solvate the 1 : 1 complex. The solution separates into two phases, composed of (C2H5)2(H)NH+ · OS(O2)CH3 complexes and pure DEA.  相似文献   

11.
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   

12.
From variable temperature vibrational Raman spectra, the axial/equatorial enthalpy differences for the substituted silacyclohexanes C5H10SiHMe, C5H10SiH(CF3) and C5H10SiCl(SiCl3) were determined. The pure liquids and solutions in various solvents were investigated. Preferred conformations are equatorial for methylsilacyclohexane and axial for trifluoromethylsilacyclohexane, consistent with earlier results from nuclear magnetic resonance experiments and ab initio calculations. For C5H10SiCl(SiCl3) an enthalpy difference close to zero was found, which is supported by high‐level which is supported by high‐level quantum chemical calculations at the second‐order Møller‐Plesset (MP2) and coupled cluster with single, double, and perturbative triple excitations (CCSD(T)) levels, which employed various basis sets. A novel synthesis for C5H10SiCl(SiCl3) was developed using ClMg(CH2)5MgCl instead of BrMg(CH2)5MgBr as a starting material. The procedure avoids the formation of partially brominated products, facilitating the purification of the compound. 1H, 13C and 29Si nuclear magnetic resonance data are reported. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

13.
Rate constants for the reactions of Cl atoms with CH3OCHCl2 and CH3OCH2CH2Cl were determined at (296 ± 2) K and atmospheric pressure using synthetic air as bath gas. Decay rates of these organic compounds were measured relative to the following reference compounds: CH2ClCH2Cl and n‐C5H12. Using rate constants of 1.33 × 10?12 and 2.52 × 10?10 cm3 molecule?1 sec?1 for the reaction of Cl atoms with CH2ClCH2Cl and n‐C5H12, respectively, the following rate coefficients were derived: k(Cl + CH3OCHCl2) = (1.05 ± 0.11) × 10?12 and k(Cl + CH3OCH2CH2Cl) = (1.14 ± 0.10) × 10?10, in units of cm3 molecule?1 s?1. The rate constants obtained were compared with previous literature data and a correlation was found between the rate coefficients of some CH3OCHR1R2 + Cl reactions and ΔElectronegativity of ? CHR1R2. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

14.
Density-dependent 13C nuclear magnetic shielding has been found for each of the pure gases CH4, C2H6, C2H4, CO and CO2, and for several binary mixtures of gases. For methane gas the density dependence is greater at higher temperatures in contrast to expectation and the observed temperature dependence of the shielding at zero density is attributable to nuclear motion. 13C magnetic shielding is considerably higher in the gas phase than in the liquid phase and the difference varies for chemically non-equivalent 13C nuclei by amounts which are well above the level of experimental error.  相似文献   

15.
The reaction of C5H4RLi with FeCl2 gave nine new compounds of Fe(C5H4R)2 [R=C(CH3)2C6H4CH3-p(-m,-o), C6H10C6H5, C(Me)2C6C4OCH3-o, C6H10C6H4CH3-p(,-m,-o), C6H10C6H4OCH3-p]. The compositions of compounds were determined through elementary analysis. The structural determination was made by IR and H2NMR. Mossbauer spectia were taken at room temperature. The IS and QS values are 0.41–0.45mm/s and 2.3–2.5mm/s., respectively. The solid state structure of the complex has been determined by a single crystal x-ray diffraction study, crystal data for Fe[C5H4C(CH3)2C6H5]2: a=17.988(2)A, b=17. 411(2)A, c=7.496(1)A, α=β=90°, r=112.23°, Z=4, monoclinic form, space group C2/c. Our conclusions are: in π-acceptor ligand, the nucleophilic substituents decrease and the electrophilic substituts increase the metal to ligand electron cloud shift, which results in a decrease or an increase in the strength of the coordinate bonds and in the stabilization of the complexes by their steric effect.  相似文献   

16.
13C-NMR spectra of several 9-acridanones with different substituents both on the ring (R1 = CH3, OCH3, NH2, N(CH3)2, NO2) and at the nitrogen atom (R2 = H, CH3 C2H5, CH2-C6H5, C[tbnd]C-CH3, (CH2)2N(C2H2)2, CH=C=CH2) have been recorded. The C-NMR chemical shifts are discussed as a function of the nature of the substituent, the importance of peri steric interactions and the electronic structure of the acridanone ring. There is a good linear relationship between the total electronic density and the chemical shifts.  相似文献   

17.
The ν3 fundamental band of H2CO (CH2 scissoring motion) has been studied by means of CO laser Stark spectroscopy and conventional infrared absorption spectroscopy. The primary aim of the work was to determine the dipole moment of H2CO in the v3 = 1 state, and the value determined was 2.3250 ± 0.0025 D. The spectrum was analyzed with the inclusion of the Coriolis interactions among ν3, ν4, and ν6 so that “true” rotational constants were determined for ν3; “effective” constants obtained by ignoring these interactions were also determined. The ν3 band origin was determined to be 1500.174 ± 0.002 cm?1. The H2CO spectrum was also used as a means of determining the frequencies of some 13C16O and 12C18O laser lines in the 1500 cm?1 region relative to 12C16O lines.  相似文献   

18.
Absolute cross-sections for electron-impact dissociative ionization of C2 H2+ and C2 D2+ to CH+, C+, C2+ , H+, CH2+ and C2D+ fragments are determined for electron energies ranging from the corresponding threshold to 2.5 keV. Results obtained in a crossed beams experiment are analyzed to estimate the contribution of dissociative ionization to each fragment formation. The dissociative ionization cross sections are seen to decrease for more than an order of magnitude, from CH+ (5.37±0.10) × 10-17 cm2 over C+ (4.19± 0.16) × 10-17 cm2, C2D+ (3.94±0.38) × 10-17 cm2, C2+ (3.82±0.15) × 10-17 cm2 and H+ (3.37±0.21) × 10-17 cm2 to CH2+ (2.66±0.14) × 10-18 cm2. Kinetic energy release distributions of fragment ions are also determined from the analysis of the product velocity distribution. Cross section values, threshold energies and kinetic energies are compared with the data available from the literature. Conforming to the scheme used in the study of the dissociative excitation of C2H2+ ( C2 D2+ )\left( {\rm C}_2 {\rm D}_2^+ \right), the cross-sections are presented in a format suitable for their implementation in plasma simulation codes.  相似文献   

19.
The kinetics and mechanism of the nucleophilic vinylic substitution of dialkyl (alkoxymethylidene)malonates (alkyl: methyl, ethyl) and (ethoxymethylidene)malononitrile with substituted hydrazines and anilines R1–NH2 (R1: (CH3)2N, CH3NH, NH2, C6H5NH, CH3CONH, 4‐CH3C6H4SO2NH, 3‐ and 4‐X‐C6H4; X: H, 4‐Br, 4‐CH3, 4‐CH3O, 3‐Cl) were studied at 25 °C in methanol. It was found that the reactions with all hydrazines (the only exception was the reaction of (ethoxymethylidene)malononitrile with N,N‐dimethylhydrazine) showed overall second‐order kinetics and kobs were linearly dependent on the hydrazine concentration which is consistent with the rate‐limiting attack of the hydrazine on the double bond of the substrate. Corresponding Brønsted plots are linear (without deviating N‐methyl and N,N‐dimethylhydrazine), and their slopes (βNuc) gradually increase from 0.59 to 0.71 which reflects gradually increasing order of the C–N bond formed in the transition state. The deviation of both methylated hydrazines is probably caused by the different site of nucleophilicity/basicity in these compounds (tertiary/secondary vs. primary nitrogen). A somewhat different situation was observed with the anilines (and once with N,N‐dimethylhydrazine) where parabolic dependences of the kinetics gradually changing to linear dependences as the concentration of nucleophile/base increases. The second‐order term in the nucleophile indicates the presence of a steady‐state intermediate ‐ most probably T±. Brønsted and Hammett plots gave βNuc = 1.08 and ρ = ?3.7 which is consistent with a late transition state whose structure resembles T±. Copyright © 2013 John Wiley & Sons, Ltd.  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