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.
C36二聚脂肪酸聚乙二醇聚酯硫酸酯盐的制备及性能表征   总被引:3,自引:0,他引:3  
以C36二聚脂肪酸(简称二聚酸)和聚乙二醇为原料,缩聚后得到聚酯,再用氨基磺酸硫酸化,制备了新型的C36二聚脂肪酸聚乙二醇聚酯硫酸酯盐型两性高分子表面活性剂,并对其性能进行了研究。聚酯适宜的制备工艺条件是:在0.097 MPa下,n(二聚酸)∶n(聚乙二醇)=1∶1.2,催化剂w(SnC l2)=0.3%(相对于反应物总质量),反应温度200℃,反应时间6 h,酯化率达到98.11%。聚酯平均相对分子质量为613 5,相对分子质量分布系数为1.212。优化的C36二聚脂肪酸聚乙二醇聚酯硫酸酯盐的制备条件是:尿素为催化剂,n(二聚酸聚乙二醇聚酯)∶n(NH2SO3H)∶n(尿素)=1∶1.2∶1.2,反应温度130℃,反应时间2.8 h,聚酯端羟基硫酸酯酯化率达81.17%。表面性能分析结果为:产物γCMC(25℃)为3.021 mN/m,与十二烷基硫酸钠(K12)相当;CMC(25℃)为1.07 mmol/L,远低于K12;乳化性能和乳化稳定性与壬基酚聚氧乙烯醚(OP-10)相当。与K12相比,产物起泡性差但表现出很好的抑泡性能。该工作的新颖性已为河南省科学技术情报研究所2005年12月15日出具的第2005 c0006号《论文查新报告》所证实。  相似文献   

2.
以C36二聚脂肪酸和聚乙二醇400为原料,缩聚得到一种新型高分子表面活性剂,适宜的工艺条件:在0.097 MPa下,n(二聚脂肪酸)∶n(聚乙二醇400)为1∶1.2,催化剂SnCl2(相对二聚脂肪酸质量分数为0.3%,反应温度200℃,反应时间6 h,酯化率达到98.11%。建立了SnCl2催化下,二聚脂肪酸与聚乙二醇缩聚反应的动力学模型,并采用改进的遗传算法,对动力学模型参数进行估算。结果显示,二聚脂肪酸与聚乙二醇400缩聚反应级数为0.998级,酯化反应活化能E=97.18 kJ/mol,指前因子A=1.947 9×109L/(mol.min),缩聚反应的Arrhenius方程为lnk=21.39-11.689/T。  相似文献   

3.
The esterification of castor oil with lauric acid was investigated using tetra n-butyl titanate (TBT), SnCl2·2H2O (stannous chloride), CoCl2·6H2O (cobalt chloride), and (CH3COO)2Zn·2H2O (zinc acetate dihydrate) as catalysts. Effects of catalyst concentration and reaction temperature on the progress of the reaction were investigated. TBT was the best catalyst for the esterification of castor oil with lauric acid at temperatures lower than 200°C. The reaction was first order with respect to each reactant. The activation energy for the esterification reaction of castor oil with lauric acid using TBT was 26.69 kcal/mol. The rate constants obtained for the esterification of castor oil with decanoic, lauric, palmitic, and stearic acids were nearly the same (15.80, 15.44, 15.06, and 14.67 mL mol−1 min−1), as were the rate constants obtained for the reaction of castor oil and hydrogenated castor oil.  相似文献   

4.
Resin immobilized lipase B from Candida antarctica (CALB) was used to catalyze the condensation polymerization of two difuctional siloxane and poly(ethylene glycol) systems. In the first system, 1,3-bis(3-carboxypropyl)tetramethyldisiloxane was reacted with poly(ethylene glycol) (PEG having a number-average molecular weight, Mn = 400, 1000 and 3400 g mol−1, respectively). In the second system, α,ω-(dihydroxy alkyl) terminated poly(dimethylsiloxane) (HAT-PDMS, Mn = 2500 g mol−1) was reacted with α,ω-(diacid) terminated poly(ethylene glycol) (PEG, Mn = 600 g mol−1). All the reactions were carried out in the bulk (without use of solvent) at 80 °C and under reduced pressure (500 mmHg vacuum gauge). The progress of the polyesterification reactions was monitored by analyzing the samples collected at various time intervals using FTIR and GPC. The thermal properties of the copolymers were characterized by DSC and TGA. In particular, the effect of the chain length of the PEG block on the molar mass build up and on the thermal stability of the copolymers was also studied. The thermal stability of the enzymatically synthesized copolymers was found to increase with increased dimethylsiloxane content in the copolymers.  相似文献   

5.
The kinetics of oxirane ring cleavage in epoxidized soybean oil have been studied using glacial acetic acid at 60, 70, 80 and 90°C. It was shown that the reaction can be successfully modelled as first order with respect to the epoxide concentration and second order with respect to acetic acid. The reaction velocity constant at 70°C was found to be 2 × 10−3 1−3 hr−1 mol−2, the frequency factor, A, = 2.321 × 107 hr−1 and the energy of activation, Ea = 15.84 k cal mol−1. The effects of the concentration of acetic acid and the temperature on the net yield of epoxides by in situ epoxidation were also studied on the basis of the predicted kinetic parameters of the reaction system.  相似文献   

6.
研究了以辛酸亚锡为催化剂、二聚酸(DA)与乙二醇(EG)为原料、制备聚氨酯用二聚酸聚酯二醇的方法,讨论了催化剂的类型和用量、反应温度、原料醇酸比、反应时间等对酯化率的影响,并用红外光谱对合成产物进行了表征。结果表明,DA与EG摩尔配比为1:2.4,催化剂辛酸亚锡用量为原料总质量的0.3%,反应温度在1h内缓慢升温到190℃,然后保温反应4h,并在130℃、2.66kPa下减压3h,酯化率可达99.7%,产品羟值为95mgKOH/g,酸值0.26mgKOH/g。  相似文献   

7.
N-vinyl pyrrolidone (NVP) was polymerized in dioxan at 60 ± 0.1°C for 1 h using diphenyl ditelluride as radical initiator. The system follows ideal kinetics i.e. R p α [DPDT]0.5[NVP]. The activation energy and dissociation constant is computed as 46 kJ mol−1 and 1.1 × 10−11 s−1, respectively. The polymer was characterized with the help of FTIR, 1H-NMR, 13C-NMR, ESR spectroscopy. The FT-IR spectrum showed bands at 1660–1680 cm−1 due to combination of >C = O and C–N stretching. The gyromagnetic constant ‘g’ has been computed as 2.2203. The main product of this reaction were poly(N-vinylpyrrolidone)s with phenyl tellanyl ends. The presence of tellurium in polymer is confirmed by ICP analysis. The DSC shows the T g of poly(N-vinylpyrrolidone) is 168°C due to rigid pyrrolidane group. The TGA showed that polymer was stable up to 380°C.The GPC studies showed that the weight average molecular weight decreases with increase of [DPDT].  相似文献   

8.
Kazuo Mukai  Yuji Okauchi 《Lipids》1989,24(11):936-939
A kinetic study of the reaction between a tocopheroxyl radical and unsaturated fatty acid esters has been undertaken. The rates of allylic hydrogen abstraction from various unsaturated fatty acid esters (ethyl oleate2, ethyl linoleate3, ethyl linolenate4, and ethyl arachidonate5) by the tocopheroxyl radical (5,7-diisopropyltocopheroxyl6) in benzene have been determined spectrophotometrically. The second-order rate constants, k3, obtained are 1.04×10−5 M−1s−1 for2, 1.82×10−2 M−1s−1 for3, 3.84×10−2 M−1s−1 for4, and 4.83×10−2 M−1s−1 for5 at 25.0°C. Thus, the rate constants, kabstr/H, given on an available hydrogen basis are k3/4=2.60×10−6 M−1s−1 for2, k3/2=9.10×10−3 M−1s−1 for3, k3/4=9.60×10−3 M−1s−1 for4, and k3/6=8.05×10−3 M−1s−1 for5. The kabstr/H values obtained for the polyunsaturated fatty acid esters3,4, and5 containing H-atoms activated by two π-electron systems are similar to each other, and are about three orders of magnitude higher than that for the ethyl oleate2 containing H-atoms activated by a single π-system. From these results, it is suggested that the prooxidant effect of α-tocopherol in edible oils and fats may be induced by the above hydrogen abstraction reaction.  相似文献   

9.
Substrate preferences for pancreatic lipase-mediated acyl-exchange reactions with butteroil were concentration-dependent for the series of acyl donors and alcohol acceptors evaluated. For acidolysis reactions, the initial reaction rates and percent reaction yields after 18 h at 50 μmol acyl donor per gram substrate mixture were similar forn-fatty acids and their methyl and glycerol esters. At 400–500 μmol g−1 (and greater), order of initial reaction rates and percent reaction yield was fatty acid glycerol esters > fatty acid methyl esters > fatty acids. At concentrations above 300–500 μmol g−1, reaction inhibition was observed for fatty acid substrates, and inhibition took place at lower concentrations for the shorter-chainlength fatty acids of those evaluated (5–17 carbons). Inhibition was primarily attributed to acidification of the microaqueous environment of the lipase. Desorption of water by the fatty acid substrate may be a secondary mode of inhibition. The concentration dependence of initial reaction rates and percent reaction yield was similar for then-alcohol substrates evaluated (2–15 carbons) for alcoholysis reactions with butteroil. Optimum alcohol concentration was 375–500 μmol g−1 (except for butanol, which was 1 mmol g−1, above which reaction inhibition was observed. Inhibition was attributed to desorption of water from the enzyme by the alcohol substrate. Relative reactivity of classes of alcohols for this reaction system was primary alcohols > secondary alcohols > tertiary alcohols. Generally, alcoholysis reactions were faster than acidolysis reactions, and triacylglycerols were the best substrates for acidolysis reactions with butteroil at high levels (up to 2 mmol g−1) of acyl donor substrate.  相似文献   

10.
The kinetics of the oxirane cleavage of epoxidized soybean oil (ESO) by methanol (Me) without a catalyst was studied at 50, 60, 65, 70 °C. The rate of oxirane ring opening is given by k[Ep][Me]2, where [Ep] and [Me] are the concentrations of oxiranes in ESO and methanol, respectively and k is a rate constant. From the temperature dependence of the kinetics thermodynamic parameters such as enthalpy (ΔH), entropy (ΔS), free energy of activation (ΔF) and activation energy (ΔE a) were found to be 76.08 (±1.06) kJ mol−1, −118.42 (±3.12) J mol−1 k−1, 111.39 (±2.86) kJ mol−1, and 78.56 (±1.63) kJ mol−1, respectively. The methoxylated polyols formed from the oxirane cleavage reaction , were liquid at room temperature and had three low temperature melting peaks. The results of chemical analysis via titration for residual oxiranes in the reaction system showed good agreement with IR spectroscopy especially the disappearance of epoxy groups at 825, 843 cm−1 and the emergence of hydroxy groups at the OH characteristic absorption peak from 3,100 to 3,800 cm−1.  相似文献   

11.
Epoxidation of karanja (Pongamia glabra) oil by H2O2   总被引:1,自引:0,他引:1  
Epoxidation of karanja oil (KO), a nondrying vegetable oil, was carried out with peroxyacetic acid that was generated in situ from aqueous hydrogen peroxide and glacial acetic acid. KO contained 61.65% oleic acid and 18.52% linoleic acid, respectively, and had an iodine value of 89 g/100 g. Unsaturated bonds in the oil were converted to oxirane by epoxidation. Almost complete epoxidation of ethylenic unsaturation was achieved. For example, the iodine value of the oil could be reduced from 89 to 19 by epoxidation at 30°C. The effects of temperature, hydrogen peroxide-to-ethylenic unsaturation ratio, acetic acid-to-ethylenic unsaturation ratio, and stirring speed on the epoxidation rate and on oxirane ring stability were studied. The rate constant and activation energy for epoxidation of KO were 10−6 L·mol−1·s−1 and 14.9 kcal·mol−1, respectively. Enthalpy, entropy, and free energy of activation were 14.2 kcal·mol−1, −51.2 cal·mol−1·K−1, and 31.1 kcal·mol−1, respectively. The present study revealed that epoxides can be developed from locally available natural renewable resources such as KO.  相似文献   

12.
The kinetics of the polyesterification in bulk at 200°C between o-phthalic anhydride and neopentyl glycol (2,2-dimethyl-1,3-propanediol) in a nonequimolecular ratio and in the absence of an external catalyst was investigated. The formation of the monoester and two dimeric compounds by uncatalyzed heating of o-phthalic anhydride with neopentyl glycol was virtually complete after dissolution of the anhydride. The data were analyzed statistically by a mathematical method due to Fradet and Maréchal for the estimation of the orders of reaction. This statistical adjustment of the data analysis supports the assumption that the kinetics of polyesterification in the absence of both the solvent and catalyst may be fitted to several orders of reaction over all the conversion. At first, our experimental data may be fitted to 3, or , or 2, etc., overall orders. The results establish that the overall kinetic order of the polyesterification depend upon the goodness of the experimental results and cannot be only selected by means of a correlation coefficient. If this last criterium is adopted, the polyesterification at low, medium, as well as at high conversions may be 3 as the most probable one, order one with respect to acid group concentration, and order two with respect to alcoholic group concentration, in agreement with Flory's predictions. A mechanism consistent with the most plausible kinetic results (Im,n = 0.9990 and m, n = 1, 2) is proposed. It consists of a dimerization of the alcoholic groups followed by an attack of the acid to the dimer. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 66: 2409–2431, 1997  相似文献   

13.
The polyesterification between adipic acid and ethylene glycol in nonequimolar ratios was investigated. The kinetic equations obtained were quite different from those obtained by Flory in equimolar systems. The kinetic equations obtained in this study were ?d[COOH]/dt = k[COOH][OH]2 for uncatalyzed polyesterification and ?d[COOH]/dt = k[COOH]2 for catalyzed polyesterification. The mechanism of the polyesterification of a dibasic acid and a glycol can only be explained by the dissociation effect of hydrogen ion from dibasic acid in glycol.  相似文献   

14.
A series of multi-arm star-branched polyisobutylenes was synthesized via living carbocationic polymerization. Arms with molecular weights ranging from 10,000 to 30,000 g/mol were prepared using the cumyl chloride/TiCl4/pyridine initiation system in 60/40 (v/v) hexane/methyl chloride at − 80 °C and linked by sequential addition of divinylbenzene. The weight average number of arms per star polymer, N w, scaled inversely with arm molecular weight and ranged from 32 to 5. Star-telechelic ionomers were produced by sulfonation of the aromatic initiator residue at the end of each arm, followed by neutralization. Sulfonation was quantitative as indicated by acid-base titration. Potassium ionomers were elastic solids which were marginally soluble in THF; the precursors were tacky and freely soluble in THF. Ionic modification did not alter the glass transition temperature (− 66 °C), but the thermal decomposition temperature in N2 was increased from 375 to 400 °C. Received: 25 July 1997/Accepted: 27 August 1997  相似文献   

15.
A facile synthetic approach to aromatic and semiaromatic amine-terminated hyperbranched polyamides via direct polymerization of triamine (B3) with different diacid chlorides (A2) was explored. An aromatic triamine, 1,3,5-tris(4′-aminophenylcarbamoyl)benzene (TAPCB), was synthesized and monomers were characterized by elemental analysis, FTIR, 1H and 13C NMR spectroscopy. Finally, the polycondensation reaction of TAPCB with terephthaloyl chloride (TPC), isophthaloyl chloride (IPC), sebacoyl chloride (SC) and adipoyl chloride (AC) resulted in the preparation of four hyperbranched polyamides i.e., HBPA 1, 2, 3 and 4, respectively. FTIR and 1H NMR analyses confirmed the structures of the ensuing polymers and DB was found between 0.51–0.55. These thermally stable amorphous HBPAs were soluble in polar aprotic solvents at room temperature having glass transition temperatures (Tg) between 138–198 °C. Inherent viscosities (ηinh) and weight average molecular weights (Mw) were in the range of 0.27–0.35 dL/g and 1.3 × 104–2.7 × 104, respectively. Future prospects are envisaged.  相似文献   

16.
To further verify the forming mechanism of multi-core structure during the direct melt copolycondensation of lactic acid (LA) with the compounds containing multifunctional groups, the biodegradable material poly(lactic acid-co-glycerol) [P(LA-co-GL)] was synthesized as designed using L-lactic acid (L-LA) and glycerol (GL) as the starting materials. For the molar feed ratio n(LA)/n(GL) of 60/1, the optimal synthetic conditions were discussed. Using 0.3 wt% stannous oxide (SnO) as the catalyst, after the prepolymerization was carried out at 140 °C for 8 h, the melt copolymerization for 8 h at 160 °C gave the polymer with the biggest intrinsic viscosity ([η]) 0.76 dL•g−1. The copolymers P(LA-co-GL)s at different molar feed ratios were characterized by Fourier transform infrared spectroscopy (FT-IR), proton nuclear magnetic resonance (1H-NMR), gel permeation chromatography (GPC), differential scanning calorimetry (DSC) and X-ray diffraction (XRD). Increasing the molar feed ratio n(LA)/n(GL), the weight-average molecular weight (Mw) didn’t increase all the time, but a peak of Mw was formed, which indeed validated the above special phenomenon during the direct melt copolycondensation of LA with the monomers containing multifunctional groups. However, the forming mechanism of multi-core copolymer was different when multihydroxyl alcohol (e.g. GL) was used as the monomer containing multifunctional groups. Because the multi-core structure was linked by the ether bonds with less reversibility in the reaction, the biggest Mw of copolymers was relatively lower. For GL with three terminal hydroxyls as the core, only when n(LA)/n(GL) was more than 100/1, the star-shaped polylactic acid (SPLA) containing one core could be obtained.  相似文献   

17.
Methanol electrooxidation was investigated on Pt–Ru electrocatalysts supported on glassy carbon. The catalysts were prepared by electrodeposition from solutions containing chloroplatinic acid and ruthenium chloride. Bulk composition analysis of the Pt–Ru catalyst was performed using an X-ray detector for energy dispersive spectroscopy analysis (EDX). Three different compositions were analyzed in the range 0–20 at.% Ru content. Tafel plots for the oxidation of methanol in solutions containing 0.1–2 M CH3OH, and in the temperature range 23–50 °C showed a reasonably well-defined linear region. The slope of the Tafel plots was found to depend on the ruthenium composition. The lower slope was determined for the Pt catalyst, varying between 100 and 120 mV dec−1. The values calculated for the alloys were higher, ranging from 120 to 140 mV dec−1. The reaction order for methanol varies from 0.5 to 0.8, increasing with the ruthenium content. The activation energy calculated from Arrhenius plots was found to change with the catalyst composition, showing a lower value around 30 kJ mol−1 for the alloys, and a higher value, of 58.8 kJ mol−1, for platinum. The effect of ruthenium content is explained by the bifunctional reaction mechanism.  相似文献   

18.
Directly starting from lactic acid (LA) and trimesic acid (TMA), novel biodegradable material poly(lactic acid-trimesic acid) (PLT), a modified polylactic acid (PLA) with terminal carboxyl, was synthesized via melt copolycondensation. The optimal synthetic conditions, including catalyst kinds and dosage, prepolymerization time, copolymerization temperature and time, were discussed. When L-lactic acid (L-LA) and TMA as molar feed ratio n(L-LA)/n(TMA) 120/1 was prepolymerized for 8 h at 140 °C, the copolycondensation catalyzed by 0.9 wt % SnCl2 at 190 °C for 8 h gave PLT with the biggest intrinsic viscosity ([η]) 1.91 dL∙g−1, and the corresponding weight-average molecular weight (Mw) was 14,100 Da. Serial L-PLTs at different molar feed ratios were synthesized and characterized with FTIR, 1H NMR, GPC, DSC, and XRD. Increasing n(L-LA), Mw increased first, and the biggest Mw was 17500 Da at n(L-LA)/ n(TMA) 240/1, then decreased. Using D,L-lactic acid (D,L-LA) instead of L-LA, the influences of LA stereochemical configuration were investigated. The peak phenomenon of Mw was similar, but the biggest Mw was 23,100 Da at n(D,L-LA)/n(TMA) 320/1. The serial L-PLTs had a certain crystallinity (10.2%∼23.0%), while all D,L-PLTs were amorphous. These differences may be in touch with the reaction mechanism of direct melt copolycondensation. The method was simple and practical for the synthesis of PLA biomedical materials applied in drug delivery carrier, and vessel substitution material.  相似文献   

19.
In this study, 10 different vegetable oils were oxidized at four different isothermal temperatures (383, 393, 403, and 413 K) in a differential scanning calorimeter (DSC). The protocol involved oxidizing vegetable oils in a DSC cell with oxygen flow. A rapid increase in evolved heat was observed with an exothermic heat flow appearing during initiation of the oxidation reaction. From this resulting exotherm, the onset of oxidation time (T o) was determined graphically by the DSC instrument. In our experimental data, linear relationships were determined by extrapolation of the log (T o) against isothermal temperature. The rates of lipid oxidation were highly correlated with temperature. In addition, based on the Arrhenius equation and activated complex theory, reaction rate constants (k), activation energies (E a), activation enthalpies (ΔH ), and activation entropies (ΔS ) for oxidative stability of vegetable oils were calculated. The E a′, ΔH , and ΔS for all vegetable oils ranged from 79 to −104 kJ mol−1, from 76 to −101 kJ mol−1, and from −99 to −20 J K−1 mol−1, respectively. Based on the results obtained, differential scanning calorimetry appears to be a useful new instrumental method for kinetic analysis of lipid oxidation in vegetable oil.  相似文献   

20.
Oxidative polycondensation reaction conditions of [(2-mercaptophenyl)iminomethyl]-2-naphthol (2-MPIM-2N) were studied using oxidants such as air and NaOCl in an aqueous alkaline medium between 40 °C and 90 °C. The structure of poly-[(2-mercaptophenyl)iminomethyl]-2-naphthol (P-2-MPIM-2N) was characterized by 1H- 13C NMR, FT-IR, and UV–Vis spectroscopy, size exclusion chromatography (SEC), and elemental analysis. At optimum reaction conditions, the yield of P-2-MPIM-2N was found to be 78 and 82% for air and NaOCl oxidants, respectively. From SEC measurements, the number-average molecular weight (M n ), weight-average molecular weight (M w ) and polydispersity index (PDI) of P-2-MPIM-2N are 2900, 3500 g mol−1 and 1.207; 2200, 2500 g mol−1 and 1.136, for air and NaOCl oxidants, respectively. Polymer–metal complexes were synthesized by the reaction of P-2-MPIM-2N with Co2+, Cu2+, Zn2+, Pb2+ and Cd2+ ions. The highest occupied molecular orbital (HOMO), the lowest unoccupied molecular orbital (LUMO), and electrochemical band gaps (Eg E^{\prime}_{g} ) of 2-MPIM-2N and P-2-MPIM-2N were −5.97, −2.66 and 3.31 eV and −5.82, −2.68 and 3.14 eV, respectively. The conductivity of polymer and polymer–metal complexes were determined in the solid state. Conductivity measurements of doped and undoped Schiff base polymer and polymer–metal complexes were carried out at room temperature and atmospheric pressure by the four-point probe technique using an electrometer. The conductivities of the polymer and polymer–metal complexes increased when iodine was used as doping agent.  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