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 296 毫秒
1.
A comparative kinetic study of the urethane reactions of phenyl isocyanate and 1,2-, 1,3-, and 1,4-butanediol was carried out in dichloromethane solution with zirconium (IV) acetylacetonate as catalyst. In situ FT-IR was used to follow the kinetics of the reactions at a constant temperature of 15°–30°C. The rate constants for the reaction of the primary hydroxyl group and the secondary hydroxyl group were calculated as k prim and k sec, respectively. Analysis of the second-order rate constants of these systems indicated that k prim follows 1,2-butanediol >1,3-butanediol >1,4-butanediol. The ratio of k prim/k sec in 1,2-butanediol was the highest and the order followed was the same as with the reaction rate. Activation energies and Eyring parameters were also determined for the urethane reaction of butanediols.  相似文献   

2.
A new environmentally friendly catalyst, H4SiW12O40-polyaniline (PAn), was prepared, and n-butyraldehyde 1,2-propanediol acetal was synthesized from n-butyraldehyde and 1,2-propanediol in the presence of H4SiW12O40-PAn. The influence factors of the synthesis were discussed, and the best reaction conditions were found: the molar ratio of n-butyraldehyde to 1,2-propanediol is 1:1.5, the amount of catalyst used is 1.2% of feed stock, and the reaction time is 1.0 h. H4SiW12O40-PAn is an excellent catalyst for synthesizing n-butyraldehyde 1,2-propanediol acetal, and the yield can reach more than 95.2%. *Translated from Journal of Central China Normal University (Natural Sciences Edition), 2005, 39(9) (in Chinese), 2004, 28(4) (in Chinese)  相似文献   

3.
The urethane reaction of 1,2‐propanediol with phenyl isocyanate was investigated with ferric acetylacetonate (Fe(acac)3) as a catalyst. In situ Fourier transform infrared spectroscopy was used to monitor the reaction, and catalytic kinetics of Fe(acac)3 was studied. The reaction rates of both hydroxyl groups were described with a second‐order equation, from which the influence of the Fe(acac)3 concentration and reaction temperature was discussed. It was very surprising that the relationship between 1/C and t became constant when reaction temperature increased, which indicated that there was no reactive distinction between the two hydroxyl groups. Although the phenomenon differed with the variation of temperature, it was unaffected by the Fe(acac)3 concentration. It was attributed to the transformation of the reaction mechanism with the increase in temperature. Furthermore, activation energy (Ea), enthalpy (ΔH*), and entropy (ΔS*) for the catalyzed reaction were determined from Arrhenius and Eyring equations, which testified to the transformation of the reaction mechanism.  相似文献   

4.
The new linear polyesters containing sulfur in the main chain were obtained by melt polycondensation of diphenylmethane-4,4′-di(methylthiopropionic acid) with ethanediol, 1,3-propanediol, 1,4-butanediol, 1,5-pentanediol, 1,6-hexanediol, 1,2-propanediol, 1,3-butanediol, and 2,2′-oxydiethanol. Low-molecular weights, low-softening temperatures and, very good solubility in organic solvents are their characteristics. The structure of all polyesters was determined by elemental analysis, FT-IR and 1H-NMR spectroscopy, and x-ray diffraction analysis. The thermal behavior of these polymers was examined by differential thermal analysis (DTA), thermogravimetric analysis (TGA), and differential scanning calorimetry (DSC). The kinetics of polyesters formation by uncatalyzed melt polycondensation was studied in a model system: diphenylmethane-4,4′-di(methylthiopropionic acid) and 1,4-butanediol or 2,2′-oxydiethanol at 150, 160, and 170°C. Reaction rate constants (k3) and activation parameters (ΔG, ΔH, ΔS) from carboxyl group loss were determined using classical kinetic methods. © 1997 John Wiley & Sons, Inc.  相似文献   

5.
The reactions of 3-phenyl-3-methylamino-1,2-propanediol 1a and 3-[(tert-butyldimethylsilyl)oxy]-1-methylamino-1-phenyl-2-propanol 1b with (CH2O)n and CH2Cl2 are appropriate procedures for the preparation of 1,3-oxazines or 1,3-oxazolidines under proper selection of kinetic or thermodynamic reaction conditions. The reaction of 1b with (CH2O)n or CH2Cl2, affords the oxazolidine 2b under kinetic conditions and then this compound can be slowly converted into 5-[(tert-butyldimethylsilyl)oxy]-3-methyl-4-phenyl-3,4,5,6-tetrahydro-2H-1,3-oxazine 3b under thermodynamic control. The mechanism proposed for this transformation and the effect of polar solvents on the acceleration of the reaction has been studied theoretically (DFT level).  相似文献   

6.
Densities of binary and ternary mixtures containing water + (1,2-ethanediol or 1,2-propanediol or 1,3-propanediol or 1,2-butanediol) + (1-n-butyl-3-methylimidazolium bromide at 0.01 mole fraction) at 298.15 K and atmospheric pressure have been determined as a function of composition using an Anton Paar densimeter (Model DMA 55). Excess molar volumes (VmEV_{\mathrm{m}}^{\mathrm{E}}) were calculated. The values are negative for all mixtures over the whole composition range.  相似文献   

7.
2,4-Disubstituted 1,3-dioxolanes were synthesized by reactions of benzaldehyde, its para-chloro derivatives, as well as 3-cyclohexenecarboxaldehyde with 3-(2-propenyloxy)-1,2-propanediol. Theproducts were brought into bromination, dichlorocarbene addition, and epoxidation reactions. It is found that when both components of the heterogeneous reaction of dioxolane ring formation have a double bond, the acid catalyst in strongly deactivated.  相似文献   

8.
The dependence of the kinetic parameters of urethane formation in the reaction between isophorone diisocyanate and alcohols of different structure (n-propanol, isopropanol, propargyl alcohol, 1,3-diazidopropan-2-ol, and phenol) in diluted solutions on the natures of solvent (toluene, carbon tetrachloride) and catalyst (dibutyltin dilaurate, diazobicyclooctane) was found using an original IR spectroscopic procedure. The ratio of the apparent rate constants for the reactions involving the aliphatic and cycloaliphatic NCO groups of isophorone diisocyanate was determined, and the efficiency of catalysis in these reactions was estimated. The reaction conditions under which the difference between the reactivities of isocyanate groups can reach 40 were determined.  相似文献   

9.
采用浸渍法制备了Ni掺杂的Cu-ZnO催化剂,采用多种物理化学手段研究了其化学物理性质及甘油加氢制取1,2-丙二醇反应催化性能。结果发现,金属Ni助剂的引入可以进一步优化Ni-Cu-ZnO催化剂的甘油加氢生成1, 2-丙二醇的反应活性。少量金属Ni的加入,Ni-Cu-ZnO催化剂的甘油转化率变化不大,生成1, 2-丙二醇的选择性明显增加。而进一步增加Ni含量到nNi/nCu=0.5,Ni含量过高会导致Ni-Cu-ZnO催化剂中实际Cu原子的量减少,从而导致甘油转化率下降。Ni掺杂的Cu-ZnO催化剂甘油加氢性能稳定性较好,在反应102 h后没有明显变化。  相似文献   

10.
采用浸渍法制备了Ni掺杂的Cu-Zn O催化剂,采用多种物理化学手段研究了其化学物理性质及甘油加氢制取1,2-丙二醇反应催化性能。结果发现,金属Ni助剂的引入可以进一步优化Ni-Cu-Zn O催化剂的甘油加氢生成1,2-丙二醇的反应活性。少量金属Ni的加入,Ni-Cu-Zn O催化剂的甘油转化率变化不大,生成1,2-丙二醇的选择性明显增加。而进一步增加Ni含量到nNi/nCu=0.5,Ni含量过高会导致Ni-Cu-Zn O催化剂中实际Cu原子的量减少,从而导致甘油转化率下降。Ni掺杂的Cu-Zn O催化剂甘油加氢性能稳定性较好,在反应102 h后没有明显变化。  相似文献   

11.
Hydrogenolysis of glycerol to 1,2-propanediol and 1,3-propanediol has significant scientific importance and commercial interest due to the huge surplus of glycerol and the various application of propanediols. A series of supported Ag–Cu catalysts synthesized by impregnation method were studied for hydrogenolysis of glycerol to propanediols. The catalysts were characterized by H2-TPR, NH3-TPD, XRD, BET, N2O chemisorption, TG, ICP and SEM. It was observed that the loading of 5% Ag–Cu-based catalysts facilitated the reduction, surface acidity and dispersion of the Cu particles, which improved the conversion of glycerol and promoted the generation of propanediols. It was also found that when loading Ag and Cu simultaneously on Al2O3, the catalyst had a better performance for the reaction because of the higher acidity, dispersion and surface area of the Cu species on the catalyst surface. In addition, effects of metal concentrations, metal impregnation sequence, reaction temperature, reaction pressure, reaction time, solvent and pH value of the solution on glycerol hydrogenolysis together with the recyclability of catalyst were investigated in detail. The optimal 5Ag–15Cu/Al2O3 achieved 66.4% glycerol conversion with 68.2% 1,2-propanediol and 3.1% 1,3-propanediol selectivity at 200 °C under 3.5 MPa in ethanol for 8 h.  相似文献   

12.
The reaction mechanism of 1,2×n‐deoxydehydration (DODH; n=1, 2, 3 …) reactions with 1‐butanol as a reductant in the presence of methyltrioxorhenium(VII) catalyst has been investigated by DFT. The reduced rhenium compound, methyloxodihydroxyrhenium(V), serves as the catalytically relevant species in both allylic alcohol isomerization and subsequent DODH processes. Compared with three‐step pathway A, involving [1,3]‐transposition of allylic alcohols, direct two‐step pathway B is an alternative option with lower activation barriers. The rate‐limiting step of the DODH reaction is the first hydrogen transfer in methyltrioxorhenium(VII) reduction. Moreover, the increase in the distance between two hydroxyl groups in direct 1,2×n‐DODH reactions for C4 and C6 diols results in a higher barrier height.  相似文献   

13.
Excess molar enthalpies (H m E) of ternary mixtures containing water+(1,2-propanediol or 1,3-propanediol or 1,2-butanediol or 1,3-butanediol or 1,4-butanediol or 2,3-butanediol)+(sodium bromide, or ammonium bromide, or tetraethyl ammonium bromide, or 1-n-butyl-3-methylimidazolium bromide at 0.1 mol⋅dm−3) at 298.15 K and atmospheric pressure have been determined as a function of composition using a modified 1455 Parr mixture calorimeter. The H m E values are negative for all mixtures over the whole composition range. The influence of the electrolyte on the hydrophobic and hydrophilic effects as well as on the behavior of H m E is discussed.  相似文献   

14.
Density functional theory was used to perform a theoretical evaluation of (E)‐1,2‐disubstituted ethylenes as dipolarophiles for the 1,3‐dipolar cycloaddition reaction. The reactivities of electron‐withdrawing and ‐donating substituted ethylenes were examined by estimating their activation energies. The calculated activation energies predicted that the most reactive species is (E)‐1,2‐C2H2(NO)2, whereas the least reactive is (E)‐2‐butene. Namely, it was demonstrated that 16‐electron 1,3‐dipole reactants with more electropositive substituents in terminal positions and ethylenes that possess more strongly electron‐withdrawing substituents facilitate 1,3‐dipolar cycloaddition reactions. All of the theoretical results can be rationalized using the configuration mixing model. © 2001 John Wiley & Sons, Inc. Int J Quant Chem 83: 318–323, 2001  相似文献   

15.
The multi‐ligand coordination compound copper(II) 1,2‐diaminopropane (pn) azide, [Cu(pn)(N3)2]n ( 1 ), was synthesized using pn and azido groups. It was characterized by X‐ray single crystal diffraction, elemental analysis, and FT‐IR spectroscopy. The crystal structure of 1 belongs to the monoclinic system, space group C2/c. The copper(II) cation is six‐coordinated by one pn molecule and four azido ligands with μ‐1 and μ‐1,1,3 coordination modes. Thermogravimetric investigations with a heating rate of 10 K · min–1 under nitrogen showed one main exothermic stage with a peak temperature of 215.7 °C in the DSC curve. The non‐isothermal kinetics parameters were calculated by Kissinger and Ozawa methods, respectively. The heat of combustion was measured by oxygen bomb calorimetry, and the enthalpy of formation, the critical temperature of thermal explosion, the entropy of activation (ΔS), the enthalpy of activation (ΔH), and the free energy of activation (ΔG) were calculated. The measurements showed that 1 has very high impact, friction, and flame sensitivities.  相似文献   

16.
Various amounts and different types of heteropolyacids promoted 5Ag15Cu/Al2O3 catalysts were prepared by impregnation method and analyzed through many techniques. The synthesized catalysts were evaluated for hydrogenolysis of glycerol to propanediols. The results demonstrated that heteropolyacids loading facilitated the reduction, promoted the dispersion, enhanced the acidity, and increased Broensted acid sites of the AgCu catalysts, which benefited the generation of 1,3-propanediol. Compared with phosphotungstic acid and phosphomolybdic acid, silicotungstic acid promoted AgCu catalyst had a better performance for 1,3-propanediol due to the better Cu dispersion and higher Broensted acidity. In addition, when the reaction was performed at 220 °C under 3.5 MPa for 8 h, the chosen 5Ag15Cu-10HSiW/Al2O3 achieved a 69.6% glycerol conversion with 35.6% 1,2-propanediol selectivity and 21.5% 1,3-propanediol selectivity.  相似文献   

17.
Using a relative rate method, rate constants have been measured for the gas-phase reactions of the OH radical with 1-hexanol, 1-methoxy-2-propanol, 2-butoxyethanol, 1,2-ethanediol, and 1,2-propanediol at 296±2 K, of (in units of 10−12 cm3 molecule−1 s−1): 15.8±3.5; 20.9±3.1; 29.4±4.3; 14.7±2.6; and 21.5±4.0, respectively, where the error limits include the estimated overall uncertainties in the rate constants for the reference compounds. These OH radical reaction rate constants are higher than certain of the literature values, by up to a factor of 2. Rate constants were also measured for the reactions of 1-methoxy-2-propanol and 2-butoxyethanol with NO3 radicals and O3, with respective NO3 radical and O3 reaction rate constants (in cm3 molecule−1 s−1 units) of: 1-methoxy-2-propanol, (1.7±0.7)×10−15, and <1.1×10−19; and 2-butoxyethanol, (3.0±1.2)×10−15, and <1.1×10−19. The dominant tropospheric loss process for the alcohols, glycols, and glycol ethers studied here is calculated to be by reaction with the OH radical, with lifetimes of 0.4–0.8 day for a 24 h average OH radical concentration of 1.0×106 molecule cm−3. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 533–540, 1998  相似文献   

18.
The densities of the homologous series of alcohols and diols RnCH2OH (n=2–6), CH3CHOHRn (n=1–5), 1,2-propanediol, 1,3-1,4-, and 2,3-butanediol, 1,5-pentanediol, and 1,7-heptanediol dissolved in propylene carbonate have been measured at 25°C. The partial molal volumes at infinite dilution have been evaluated. Additivity of group molal volumes has been confirmed in propylene carbonate. The results have been discussed in relation to the same data in aqueous solution, and the scaled particle theory has been employed to calculate intrinsic volumes of the solutes.  相似文献   

19.
The influence of interaction enthalpy (ΔHMOD/BuLi) of μ, σ, σ+μ and σ-μ complexing polar modifiers with n-butyllithium on the 1,3-butadiene anionic polymerization enthalpy (ΔHBD), polymerization reaction rate (kp) and polybutadiene microstructure was studied. It has been found that enthalpy of interaction depends on complex type, molar ratio of polar modifier to n-butyllithium (MOD/BuLi) and temperature. For the first time it has been proven that with increasing ΔHMOD/BuLi the content of 1,2 butadiene (vinyl) units in the chain increases individually for each complex type but the value of ΔHBD decreases similarly for all complexes containing σ-donor groups, exhibiting linear nature. Since ΔHMOD/BuLi controls the content of butadiene isomeric structures in the chain its value was compared with polymerization reaction rate, ranging from ∼46 to ∼5500 moL/L·min, and discussed on mechanistic level.  相似文献   

20.
Rate constants for the reaction of 1-chloro-2,3-epoxypropane with p-cresol in the presence of basic catalysts were studied at the temperature range of 71–100°C. It was found that in the presence of sodium p-cresolate, three consecutive reactions proceeded giving the following products: 1-chloro-3-(tolyloxy)-2-propanol (CTP), 1-(p-tolyloxy)-2,3-epoxypropane (TEP) as a main product, and 1,3-di(p-tolyloxy)-2-propanol (DTP). Their rate constants at 71°C were: k1 = 0.030 ± 0.009, k2 = 1.58 ± 0.02, and k3 = 0.033 ± 0.005 dm3/mol · min, respectively. In the presence of quaternary ammonium salts, this process consisted of 5 reactions which led to CTP as a main product as well as TEP and 1,3-dichloro-2-propanol (DCP). The rate constant of CTP formation at 71°C was established, k1 = 0.130 ± 0.030 dm3/mol · min, as were the ratios of the other rate constants k2/k−4 = 1.5 ± 0.2, k5/k−4 = 20.0 ± 5.0, and k4/k1 = 0.6 ± 0.7. Based on the changes in Cl ion concentration during the reaction, the catalystic activity of quaternary ammonium salts was explained. The kinetic model of these reactions in the presence of basic catalysts has been proposed and appropriate kinetic equations have been presented. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet 29: 73–79, 1997.  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