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 871 毫秒
1.
The reaction mechanisms for ozonolysis of ethyl vinyl ether (EVE) and propyl vinyl ether (PVE) have been investigated using the density functional theory (DFT) and ab initio method. Cycloaddition reactions of O3 to EVE and PVE are highly exothermic by 52.91 and 53.17 kcal/mol, respectively. Major products (formaldehyde, ethyl formate, and propyl formate) resulting from the both reactions are identified by comparing them with the experimental results. Further reactions of the most energy-rich Criegee intermediates (C2H5OCHOO and C3H7OCHOO) have been proposed in the presence of NO and H2O in which the main products are ethyl formate and propyl formate. The Multichannel Rice–Ramsperger–Kassel–Marcus (RRKM) approach is employed to calculate the total and individual rate constants for major product channels over a wide range of temperatures and different pressures. In the temperature range of 200–2500 K, the main path is the production of ethyl formate with k EVE+O3 = 4.67 × 10−12 exp(−3029/T), for the EVE with O3 reaction and k PVE+O3 = 3.58 × 10−12 exp(−2858/T) for the PVE with O3 reaction. At 298 K and 760 torr, the rate constants calculated are 1.80 × 10−16 and 2.45 × 10−16 cm3 molecule−1 s−1 for ozonolysis of EVE and PVE, which are consistent with the experimental results. The total rate constants show positive temperature dependence over the temperature range of 200–2000 K but pressure independence in the range of 0.01–10000 Torr. Estimation of branching ratios of several products is also performed. The influence of carbon chain length on reactivity toward ozone is examined.  相似文献   

2.

Methyl 5-acetyl-2-furoate has been prepared via oxidation of 5-(1-hydroxyethyl)-2-furoate with the Jones reagent. In turn, the starting compound has been synthesized via sequential chloroethylation of ethyl 2-furoate, substitution of chlorine with acetoxy group, and methanolysis of the acetate in presence of sodium methylate. The vinylog 2-furoate has been obtained as the major product via acetylation of ethyl 3-(2-furyl)-acrylate with acetic anhydride in the presence of magnesium perchlorate.

  相似文献   

3.
The reaction of CH3OCF2CF2OCHO with Cl atom has been investigated theoretically by direct dynamics method. The BB1K hybrid functional in conjunction with the 6-31 + G(d,p) basis set has been used to optimize the geometries for the stationary points and explore the potential energy surface of the reaction. Four rotation conformers (RC1-4) of CH3OCF2CF2OCHO are identified, and they are all considered in the kinetic calculation. For each conformer, there are two kinds of H-abstraction channels and one displacement channel, and the latter one should be negligible due to involving much higher energy barrier than the former two. The individual rate constants for each H-abstraction channel are evaluated by the improved canonical variational transition-state theory with a small-curvature tunneling correction. The overall rate constant is evaluated by the Boltzmann distribution function, and a fitted four-parameter rate constant expression is obtained over a wide temperature range of 200–2,000 K. The agreement between the calculated and available experimental value at 296 K is good. The contribution of each conformer to the title reaction is discussed with respect to the temperature. In addition, because of the lack of available experimental data for the species involved in the reactions, the enthalpies of the formation (ΔH f,298°) for the reactant and its product radicals are predicted via isodesmic reaction at the BB1K/6-31 + G(d,p) level.  相似文献   

4.
《合成通讯》2013,43(12):2029-2043
Abstract

A novel synthesis of the anti-inflammatory agent 1-(1,2,3,5,6,7- hexahydro-s-indacen-4-yl)-3-[4-(1-hydroxy-1-methyl-ethyl)-furan-2-sulfonyl] urea 1 is described. Sulfonamide 5 was prepared starting from ethyl 3-furoate 2. Key steps were a one-pot sulfonylation with chlorosulfonic acid in methylene chloride followed by pyridinium salt formation and reaction with phosphorus pentachloride to provide ethyl 2-(chlorosulfonyl)-4-furoate 7. This sulfonyl chloride was treated with ammonium bicarbonate to form sulfonamide 8, followed by treatment with excess methyl magnesium chloride to provide 4-(1-hydroxy-1-methyl-ethyl)-furan-2-sulfonamide 5. 4-Isocyanato-1,2,3,5,6,7-hexahydro-s-indacene 16 was prepared from indan in five steps. The formation of the desired sulfonyl urea was carried out both with the isolated isocyanate 16 and via an in situ method.  相似文献   

5.
The reaction mechanism of CH3SCH2CH3 with OH radical is studied at the CCSD(T)/6-311+G(3df,p)//MP2/6-31+G(2d,p) level of theory. Three hydrogen abstraction channels, one substitution process and five addition–elimination channels are identified in the title reaction. The result shows hydrogen abstraction is dominant. Substitution process and addition–elimination reactions may be negligible because of the high barrier heights. Enthalpies of formation [ \Updeltaf H(298.15\textK)o \Updelta_{f} H_{(298.15{\text{K}})}^{o} ] of the reactants and products are evaluated at the CBS-QB3, G3 and G3MP2 levels of theory, respectively. It is found that the calculated enthalpies of formation by the aforementioned three methods are in consistent with the available experimental data. Rate constants and branching ratios are estimated by means of the conventional transition state theory with the Wigner tunneling correction over the temperature range of 200–900 K. The calculation shows that the formations of P1 (CH2SCH2CH3 + H2O) and P2 (CH3SCHCH3 + H2O) are major products during 200–900 K. The three-parameter expressions for the total rate constant is fitted to be k\texttotal = 1.45 ×10 - 21 T3.24 exp( - 1384.54/T) k_{\text{total}} = 1.45 \times 10^{ - 21} T^{3.24} \exp ( - 1384.54/T) cm3 molecule−1 s−1 from 200 to 900 K.  相似文献   

6.
The external order in reagents for the activation of alkyl halides by Cu0 was investigated in supplemental activator and reducing agents (SARA) ATRP. Using methyl 2-bromopropionate (MBrP) or ethyl α-bromophenylacetate (EBPA) and tris(2-(dimethylamino)ethyl)amine (Me6TREN) in DMSO and MeCN, it was determined that the rate of activation scaled with (S/V)0.9 in both solvents. For MBrP, the rate was first order with respect to [MBrP]0 until a saturation in the rate was observed around 33 and 110 mM in DMSO and MeCN, respectively. For EBPA, the reaction was also first order until a maximum rate was observed at 33 mM in DMSO, whereas an inverse order was observed for concentrations above 66 mM in MeCN. At saturated concentrations of MBrP, it was found that the rate increased linearly with respect to [Me6TREN]0 for all systems but became asymptotic with a maximum rate of 2 × 10−6 and 4 × 10−5 M s−1 in DMSO and MeCN, respectively. Model polymerizations in the absence of ligand showed slow reaction rates, indicating the necessity for ligand. The results allow more accurate modeling and understanding of SARA ATRP under a large range of initiator concentrations. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 3048–3057  相似文献   

7.
Kinetics of the reaction of Cl atoms with methanol has been investigated at 2 Torr total pressure of helium and over a wide temperature range 225-950 K, using a discharge flow reactor combined with an electron impact ionization quadrupole mass spectrometer. The rate constant of the reaction Cl + CH3OH → products (1) was determined using both absolute measurements under pseudo-first order conditions, monitoring the kinetics of Cl-atom consumption in excess of methanol and relative rate method, k1 = (5.1 ± 0.8) × 10−11 cm3 molecule−1 s−1, and was found to be temperature independent over the range T = 225-950 K. The rate constant of the reaction Cl + Br2 → BrCl + Br (3) was measured in an absolute way monitoring Cl-atom decays in excess of Br2: k3 = 1.64 × 10−10 exp(34/T) cm3 molecule−1 s−1 at T = 225-960 K (with conservative 15% uncertainty). The experimental data for k3 can also be adequately represented by the temperature independent value of k3 = (1.8 ± 0.3) × 10−10 cm3 molecule−1 s−1. The kinetic data from the present study are compared with previous measurements.  相似文献   

8.
The kinetics of the reactions of Br2 and NO2 with ground state oxygen atoms have been studied over a wide temperature range, T = 220-950 K, using a low-pressure flow tube reactor coupled with a quadrupole mass spectrometer: O + NO2 → NO + O2 (1) and O + Br2 → Br + BrO (2). The rate constant of reaction (1) was determined under pseudo–first-order conditions, either monitoring the kinetics of O-atom or NO2 consumption in excess of NO2 or of the oxygen atoms, respectively: k1 = (6.1 ± 0.4) × 10−12 exp((155 ± 18)/T) cm3 molecule−1 s−1 (where the uncertainties represent precision at the 2σ level, the estimated total uncertainty on k1 being 15% at all temperatures). The temperature dependence of k1, found to be in excellent agreement with multiple previous low-temperature data, was extended to 950 K. The rate constant of reaction (2) determined under pseudo–first-order conditions, monitoring the kinetics of Br2 consumption in excess of O-atoms, showed upward curvature at low and high temperatures of the study and was fitted with the following three-parameter expression: k2 = 9.85 × 10−16 T1.41 exp(543/T) cm3 molecule−1 s−1 at T = (220-950) K, which is recommended from the present study with an independent of temperature conservative uncertainty of 15% on k2.  相似文献   

9.
Bromination of (E)-1-[4-(2-carboxy-vinyl)phenyl]-[1,2,3]triazole-4-carboxylic acid ethyl ester, which was synthesized in 90% yield by a Huisgen-type [3 + 2]-cycloaddition reaction between 3-(4-azidophenyl) acrylic acid and ethyl propiolate, in CHCl3 followed by a debrominative decarboxylation reaction with Et3N in DMF under microwave irradiation condition afforded stereoselective (Z)-1-(4-(2-bromovinyl)phenyl)-1,2,3-triazole-4-carboxylic acid ethyl ester in 94% yield. Treatment of (Z)-1-(4-(2-bromovinyl)phenyl)-1,2,3-triazole-4-carboxylic acid ethyl ester with EtONa in DMF afforded 1-(4-ethynylphenyl)-1,2,3-triazole-4-carboxylic acid ethyl ester in a yield of 90%.  相似文献   

10.
Decomposition of formic acid (HCO2H) proceeds via three unimolecular channels: dehydration, decarboxylation, and dissociation, the latter expected to be of minor contribution to the overall kinetics. In addition, despite the similar values reported for the individual activation energies for the dehydration and decarboxylation reactions, experimental works have shown that the former is dominant in the reaction mechanism. These reactions show pressure-dependent rate coefficients, and the high-pressure condition is not yet verified at atmospheric pressure. This work aims to investigate the influence of temperature and pressure on the rate coefficients. Hence, theoretical calculations at the CCSD(T)/CBS level have been performed to accurately describe the unimolecular reaction and Rice-Ramsperger-Kassel-Marcus (RRKM) rate coefficients have been calculated and integrated for the prediction of k(T,P) rate coefficients, adopting both strong and weak collision models, over the intervals 0.5-10 atm and 298-2200 K. Our results suggest that the isomerization path is important and explains the preference for the (CO + H2O) channel. Rate coefficients for the (CO2 + H2) and (CO + H2O) formations are given, in s−1, as exp(−34404/T) and exp(−33785/T), respectively. The dissociation limit of 107.29 kcal mol–1, with respect the Z-HCO2H conformer, leading to OH + HCO, via a barrierless potential curve, with rate coefficients, in s−1, expressed as kHCO+OH(T) = 1.68 × 1017 exp(−56018/T). Temperature and pressure dependence for the HCO + OH → CO2 + H2 and HCO + OH → CO + H2O reactions have also been estimated.  相似文献   

11.
A series of N-alkyl-N-alkyl′-pyrrolidinium-bis(trifluoromethanesulfonyl) imide (TFSI) room temperature ionic liquids (RTILs) has been investigated by means of thermogravimetric analysis (TG), differential scanning calorimetry, FT-IR spectroscopy, and X-ray diffraction analysis. These compounds exhibit a thermal stability up to 548–573 K. The mass loss starting temperature, T ml, falls in a narrow range of temperatures: 578–594 K. FT-IR spectra, performed before and after 24 h isothermal experiments at 553 and 573 K, have confirmed their great thermal stability. Below the ambient temperature, these compounds exhibit a complex behavior. N-methyl-N-propyl-pyrrolidinium-TFSI is the sole liquid which crystallizes without forming any amorphous phase even after quenching in liquid nitrogen. Its crystalline phase has a melting point, T m, of 283 ± 1 K. When the amorphous solid is heated, the N-butyl-N-ethyl-pyrrolidinium-TFSI presents a glass transition temperature, T g, at 186 K followed by a cold crystallization, T cc, at 225 K, and a final T m at 262 K. The N-butyl-N-methyl-pyrrolidinium-TFSI exhibits a T g between 186 and 181 K, its cold crystallization leading to two different solid phases. Solid phase I has a melting point T I,m = 252 K and phase II, T II,m = 262 K. When the amorphous phase is obtained at a cooling rate of 10 K/min, its T cc is 204 K, and a metastable solid phase (III) is obtained which transforms into the phase II at 226 K. However, when the sample is quenched, the amorphous phase transforms into phase II at T cc = 217 K and phase I at 239 K. P15-TFSI exhibits the most complicated pattern as, on cooling, it leads to both a crystallized phase at 237 K and an amorphous phase at 191 K. On heating, after a T g at 186 K and a T cc at 217 K, two solid–solid phase transitions are observed at 239 K and 270 K, the final T m being 279 K.  相似文献   

12.
Li Wang  Jing-yao Liu  Ze-sheng Li   《Chemical physics》2008,351(1-3):154-158
The dynamic properties of the hydrogen abstraction reactions of CF2H2 and CF3H with F atom are investigated in the temperature range of 182–2000 K. The minimum-energy path (MEP) is optimized at MP2/6-311 G(d, p) level, then the energy profiles are refined at the CCSD(T)/6-311++G(3df, 2pd) level (single-point). The theoretical rate constants, which are calculated by the variational transition state theory (VTST) including the small curvature tunneling (SCT) correction, are in good agreement with the experimental ones. It is found that the rate constant of the CF2H2 + F reaction are larger than that of the CF3H + F reaction and the activation energies exhibit in the just opposite order. This phenomenon can be rationalized by the hardness η of the halomethane molecules. The comparison of the two reactions with the CFH3 + F reaction is made. It is found that the rate constants decrease in the order of CFH3 + F > CF2H2 + F > CF3H + F. The effect of fluorine substitution leads to a dramatic increase in the activation energy and a decrease in the preexponential factor. We hope that present theoretical studies for these compounds can give further information concerning how fluorine substitution affects the rate constants of hydrogen abstraction reactions.  相似文献   

13.
Theoretical investigations are carried out on the multichannel reactions CH3COCH3 + F (R1) and CH3COCH3 + Br (R2) by means of direct dynamics methods. The minimum energy path (MEP) is obtained at the MP2/6-31 + G(d,p) level, and energetic information is further refined at the MC-QCISD (single-point) level. The rate constants are calculated by the improved canonical variational transition-state theory (ICVT) with the small-curvature tunneling (SCT) contributions in a wide temperature range 200–1,500 K for the title reactions, H-abstraction channel is favored for the two reactions. The theoretical overall rate constants are in good agreement with the available experimental data and are found to be k 1a  = 3.22 × 10−15 T 1.51exp(1,190.91/T) cmmolecule−1 s−1, k 2  = 5.95 × 10−18 T 1.98exp(−4,622.45/T) cmmolecule−1 s−1. Furthermore, the rate constants of reaction Cl + CH3COCH3 (R3) calculated in the other paper are added to discuss the reactivity trend of different halogen reaction with acetone on the rate constants of this class of hydrogen abstraction reactions.  相似文献   

14.
A new potential enantioselective catalyst derived from ferrocene, 1-{(R)-1-[(S)-2-(diphenylphosphino)ferrocenyl]ethyl}-benzimidazole (DPFEB), was prepared and its absolute structure was characterized by means of single crystal X-ray diffraction. The molar heat capacity of DPFEB was measured by means of temperature modulated differential scanning calorimetry over the temperature range of 200–530 K, and the thermodynamic functions of [H T  − H 298.15] and [S T  − S 298.15] were calculated. Further more, thermogravimetry experiment revealed that DPFEB exhibited a three step thermal decomposition process with the final residual of 28.7%.  相似文献   

15.
Abstract  Photoinitiated cationic polymerizations of various epoxy monomers bearing a carbazole moiety, 9-[3-(allyloxy)-2-(oxiran-2-ylmethoxy)propyl]-9H-carbazole, 9-[3-methoxy-2-(oxiran-2-ylmethoxy)propyl]-9H-carbazole, 9-[2-(oxiran-2-ylmethoxy)ethyl]-9H-carbazole, as well as a composition of 3,6-dibromo-9-(oxiran-2-ylmethyl)-9H-carbazole with 3,4-epoxycyclohexylmethyl 3,4-epoxycyclohexane carboxylate were investigated in bulk using triphenyl carbenium salts having anions such as BF4 , SnCl5 , and SbCl6 . Dark polymerizations of the carbazolyl monomers in the presence of the initiators are studied. These photoinduced polymerization reactions give oligomers of degree of polymerization 4-22. The effect of the anion of the photoinitiator and polymerization time on the polymerization reaction is discussed. Graphical abstract     相似文献   

16.
Heat capacity measurements of quasi-two-dimensional Mott insulating compounds consisting of BEDT-TTF (bisethylendithiotetrathiafulubalene) donor molecules and counter anions were performed by the thermal relaxation calorimetry technique for single crystal samples. No distinct thermal anomalies at the predicted antiferromagnetic transition temperatures in κ-(BEDT-TTF)2Cu[N(CN)2]Cl (T N = 27 K) and β′-(BEDT-TTF)2ICl2 (T N = 22 K) were observed. These results demonstrate that the Mott insulating state of the organic salts which are dominated by the strong two-dimensional intra-layer antiferromagnetic interactions between neighboring S = 1/2 spins shows somewhat different features from the simple quasi-two-dimensional Heisenberg model with S = 1/2. The strong quantum fluctuations produced by the electron correlations suppress the long-range character of the spin correlations, which seems to be an important aspect of this kind of Mott insulating materials.  相似文献   

17.
This work presents the first direct experimental study on the thermal unimolecular decomposition of n-C3F7I. Experiments were performed behind incident and reflected shock waves using the atomic resonance absorption spectroscopy (ARAS) technique on a resonant line of atomic iodine at 183.04 nm. The reaction C3F7I + Ar → C3F7 + I + Ar (1) was studied at specific temperature (800–1200 K) and pressure (0.6–8.3 bar) ranges. Under experimental conditions, the obtained values of the rate constant at temperatures below 950 K are close to the high-pressure limit; however, considering theoretical calculations, the influence of pressure on the rate constant at elevated temperatures remains noticeable. The resulting value of the experimental rate constant of reaction 1 is presented in the following Arrhenius form: Experimental data were found to correlate with the results of the Rice–Ramsperger–Kassel–Marcus –master equation analysis based on quantum-chemical calculations. The following low- and high-pressure limiting rate coefficients were obtained over the temperature range = 300–3000 K: with the center broadening factor Fc = 0.119.  相似文献   

18.
The spin-forbidden dissociation reaction of the N2O(X1Σ+) ground state has been investigated by both quantum calculations and experiments. Ab initio prediction at the CCSD(T)/CBS(TQ5)//CCSD(T)/aug-cc-pVTZ+d level of theory gave the crossing point (MSX) energy at 60.1 kcal/mol for the N2O(X1Σ+) → N2() + O(3P) transition, in good agreement with published data. The T- and P-dependent rate coefficients, k1(T,P), for the nonadiabatic thermal dissociation predicted by nonadiabatic Rice-Ramsperger-Kassel-Marcus (RRKM) calculations agree very well with literature data. The rate constants at the high- and low-pressure limits, k1 = 1011.90 exp (−61.54 kcal mol−1/RT) s−1 and k1o = 1014.97 exp(−60.05 kcal mol−1/RT) cm3 mol−1 s−1, for example, agree closely with the extrapolated results of Röhrig et al. at both pressure limits. The second-order rate constant (k1o) is also in excellent agreement with our result measured by FTIR spectrometry in the present study for the temperature range of 860-1023 K as well as with many existing high-temperature data obtained primarily by shock-wave heating up to 3340 K. Kinetic modeling of the NO product yields measured at long reaction times in the present work also allowed us to reliably estimate the rate constant for reaction (3), O + N2O → N2 + O2, based on its strong competition with the NO formation from reaction (2) which has been better established. The modeled values of k3 confirmed the previous finding by Davidson et al. with significantly smaller values of A-factor and activation energy than the accepted ones. A least-squares analysis of both sets of data gave k3 = 1012.22 ± 0.04 exp[− (11.65 ± 0.24 kcal mol−1/RT)] cm3 mol−1 s−1, covering the wide temperature range of 988-3340 K.  相似文献   

19.
The compounds RbAuUSe3, CsAuUSe3, and RbAuUTe3 were synthesized at 1073 K from the reactions of U, Au, Q, and A2Q3 (A=Rb or Cs; Q=Se or Te). The compound CsAuUTe3 was synthesized at 1173 K from the reaction of U, Au, Te, and CsCl as a flux. These isostructural compounds crystallize in the KCuZrS3 structure type in space group Cmcm of the orthorhombic system. The structure consists of layers that contain nearly regular UQ6 octahedra and distorted AuQ4 tetrahedra. The infinite layers are separated by bicapped trigonal prismatic A cations. The magnetic behavior of RbAuUSe3 deviates significantly from Curie–Weiss behavior at low temperatures. For T>200 K, the values of the Curie constant C and the Weiss constant θp are 1.82(9) emu K mol−1 and −3.5(2)×102 K, respectively. The effective magnetic moment μeff is 3.81(9) μB. Formal oxidation states of A/Au/U/Q may be assigned as +1/+1/+4/−2, respectively.  相似文献   

20.
The thermal dissociation of gaseous Mo(CO)6 and W(CO)6 in an argon carrier gas, Mo(CO)6 → Mo(CO)5 + CO (1) and W(CO)6 → W(CO)5 + CO (2), is studied over temperature ranges of ∼585–685 K for (1) and ∼690−810 K for (2) at a total gas concentrations of 4 × 10−6 and 4 × 10−5 mol/cm3 by using the shock tube technique in conjunction with absorption spectrophotometry. The measured rate constants are extrapolated to the high-pressure limit by means of a newly developed procedure, with the resultant expressions for the indicated temperature ranges reading as kd1,∞(T),[s−1] = 1016.12 ± 0.68exp[(−148.8 ± 8.1 kJ/mol)/RT] and kd2,∞(T),[s−1] = 1015.93 ± 0.63exp[(−171.7 ± 8.9 kJ/mol)/RT]. Comparison of the high-pressure dissociation rate constants with the published data revealed a considerable discrepancy, a tentative explanation of which is given. Based on the obtained high-pressure dissociation rate constants and the available data on the high-pressure room-temperature rate constants for the reverse reaction of recombination, the first bond dissociation energies for these molecules are evaluated and compared with previous determinations, both theoretical and experimental. The enthalpies of formation of Mo(CO)5 and W(CO)5 are determined: ΔfH°(Mo(CO)5, g, 298.15 K) = −644.1 ± 5.6 kJ/mol and ΔfH°(W(CO)5, g, 298.15 K) = −581.9 ± 6.6 kJ/mol. Based on the enthalpies of formation of Mo(CO)5, W(CO)5, Mo(CO)6, and W(CO)6, and the published molecular parameters of these four species, their thermochemical functions are calculated and presented in the form of NASA seven-term polynomials.  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