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 15 毫秒
1.
钱俊红  郭荣 《中国化学》2003,21(10):1284-1289
The hydrolysis of cephanone in SDS micelle and SDS/n-C5H11-OH/H2O O/W microemulsion was studied through Uv-vis ab-sorption spectroscopy. The change of pH value in the hydrolysis of cephanone was determined. The result shows that pH value decreases in the process of the hydrolysis, and that the SDS ml-celle and SDS/n-C5H11OH/H2O O/W microemulsion accelerate the hydrolysis of cephanone compared with water.  相似文献   

2.
以铂微电极法测定了在SDS/n-C5H11OH/H2O溶致液晶中SDS(十二烷基硫酸钠)分子的扩散系数.结果表明,恒定质量比SDS/n-C5H11OH条件下,溶致液晶中SDS分子的扩散系数随体系中水含量的增加而增加;恒定质量比SDS/H2O,溶致液晶中SDS分子的扩散系数随正戊醇含量的增加而增加;恒定质量比H2O/n-C5H11OH ,溶致液晶中SDS分子的扩散系数随SDS含量的增加而降低.六角状液晶中SDS分子的扩散系数比层状液晶中SDS分子的扩散系数低约1个数量级,而比W/O、O/W胶束的扩散系数低3~5个数量级.  相似文献   

3.
Photopolymerization in and of lyotropic liquid crystal (LLC) template phases shows great promise for generating nanostructure in organic polymers. Interestingly, the order imposed on the polymerization system in LLCs significantly alters polymerization kinetics. The rate of polymerization of hydrophilic monomers increases with increasing LLC order, primarily due to monomer/polymer association with surfactant and the resulting decrease of growing polymer chain diffusion. Conversely, as LLC order increases, hydrophobic monomers become less segregated as nonpolar volume increases, which decreases polymerization rate. The efficiency of initiators is also dependent on LLC template order, further contributing to polymerization rate changes. When reactive surfactants are used, LLC mesophase, location of reactive group, and aliphatic tail length also affect polymerization kinetics. Overall, these photopolymerization kinetics directly relate to the segregation behavior and local order of reactive groups and thus can be used to probe nanostructure evolution, facilitating understanding and control of ultimate polymer nanostructure. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2017 , 55, 471–489  相似文献   

4.
Urea can enhance the aqueous solubility of surfactant CTAB (hexadecyltrimethylammonium bromide) when it shows the hydrotrope action. It will show the hydrotrope‐solubilization action when the solubilized amount of n‐C5H11OH in O/W microemulsion and that of water in W/O microemulsion are increased. The mechanism of the hydrotrope‐solubilization action of urea is the increase of the stability of W/O and O/W microemulsion and structural transition from the lamellar liquid crystalline phase to the bicontinuous structure.  相似文献   

5.
郭荣  魏逊  刘天晴 《中国化学》2005,23(4):393-399
In the system of SDS/n-C5H11OH/n-C7H16/H2O with the weight ratio of SDS/n-C5H11OH/H2O system at5.0/47.5/47.5, the upper phase of the system was W/O microemulsion, and the lower phase was the bicontinuous microemulsion. When the n-heptane content was less than 1%, with the increase of the n-heptane content, the capacitance (Co, Cod) in the upper phase (W/O) dropped, the capacitance (CB1, CBld) in the lower phase (BI) raised. At the same time, the W/O-BI inteffacial potential (ΔE), capacitance (Ci), and charge-transfer current (ict) decreased.After the n-heptane content reached 1%, with the increase of the n-heptane content, ΔE, Ci and ict demonstrated no significant change.  相似文献   

6.
DCD/C4H9OH/LP/H2O体系溶致液晶的润滑性能   总被引:2,自引:0,他引:2  
测定了月桂酸二乙醇胺(DCD)/正丁醇(C4H9OH)/石腊油(LP)/水体系的相图.在液晶区选取两个样品点,用2HNMR技术和偏振光显微镜测定了这两样品点的微相结构.把液晶涂在铝合金表面测定不同负荷量下的磨痕宽度,并与十二羟基锂基脂等体系进行比较.结果表明,层状液晶在铝合金表面平行方向上的摩擦阻力较小,而在垂直方向上的负荷量高,是一种理想的润滑剂  相似文献   

7.
The diffusion coefficient of methylene blue (MB) is determined by the method of non-probe microelectrode voltammetry in sodium dodecyl sulfate (SDS)/n-C5H11OH/H2O lyotropic liquid crystal system. The results obtained show that the diffusion coefficient of MB increases with water and n-pentanol contents in the microemulsions and the lyotropic liquid crystal but decreases with SDS content. The diffusion coefficient of SDS droplet in the microemulsions and the diffusion coefficient of SDS molecule in the lyotropic liquid crystal with MB all are less than those without MB. The magnitude order of the diffusion coefficient of MB is as follows: the coefficient in the oil-in-water (O/W) microemulsion is greater than the coefficient in the water-in-oil (W/O) microemulsion which is greater than the coefficient in the lamellar liquid crystal (LLC), which is also greater than the coefficient in the Hex.  相似文献   

8.
Polymerization in highly ordered lyotropic liquid‐crystalline (LLC) media enables controllable synthesis of polymers possessing interesting nanostructure and physical properties. This study investigates the radical polymerization rate and molecular weight (MW) development of monoacrylates of different aliphatic tail length in a range of LLC phases. Polymerization rate data were acquired using photodifferential scanning calorimetry, and linear polymer MW was determined with gel permeation chromatography. Polymerization occurs much more rapidly, and higher MW is attained in the ordered LLC phases relative to isotropic solutions and neat polymerization. These properties change significantly as a function of LLC phase and monomer structure. A direct relationship is observed between polymer MW formation and the polymerization rate. Definitive changes in rate and MW were observed at phase boundaries, indicating the important role of solvent order. This study demonstrates how solvent ordering effects can be used to control polymer MW and rate of polymerization. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 144–154  相似文献   

9.
IntroductionMicellesolutioncomposedofsurfactantscanaffect,adjustandcontrolmanychemicalreactionsaccordingtoitseffectsoflocalconcentration ,polarity ,charge ,microvis cosity,electrostatics ,etc ..1,2 Recently ,ithasarousedmuchattentiontoconductchemicalreactionsusingmi croemulsionasmicroreactor .3 7Theeffectofmicelleonthechemicalreactionstemsfromitsstaticelectricityandhy drophobicity .Usually ,cationicsurfactantcancatalyzethereactionbetweennucleophileandneutralmolecule ,whileanionicsurfactantsusp…  相似文献   

10.
It has been found that some transition metal salts considerably decrease the induction period of the chemical polymerization of aniline. The efficiency of the promoting agents decreases in the series (N)2IrQ6> MoCl5>NiCl2.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1500–1502, August, 1994.  相似文献   

11.
A series of triarylbismuth(V) di(Np‐toluenesulfonyl)aminoacetates with the formula (4‐CH3C6H4SO2NHCH2CO2)2BiAr3 (Ar?C6H5, 4‐CH3C6H4, 4‐ClC6H4, 4‐BrC6H4) were synthesized and characterized by elemental analysis, IR, 1H NMR and mass spectra. The crystal structure of (4‐CH3C6H4SO2NHCH2CO2)2Bi(C6H4Cl‐4)3 was determined and shows the bismuth to exist in a distorted trigonal bipyramidal geometry. Four human neoplastic cell lines (HL‐60, PC‐3MIE8, BGC‐823 and MDA‐MB‐435) were used to screen these compounds. The results indicate that these compounds at 10 μM show cytotoxicity. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

12.
Yuan CHEN  Rong GUO 《中国化学》2007,25(12):1790-1794
At a weight ratio of n‐C5H11OH/H2O=50/50, when the total content of sodium dodecyl sulfate (SDS) was less than 6.0%, the ternary mixture of SDS/n‐C5H11OH/H2O coexisted in two immiscible microemulsions. The distribution and transfer of gatifloxacin (GTFX) between the two phases were studied using UV‐Vis and electrochemistry AC impedance spectra. The results show that GTFX transferred from the upper phase (W/O) to the lower phase (O/W or bicontinuous microemulsion), but a small amount of SDS transferred from the lower phase to the upper phase correspondingly with the increase of the total SDS content at a total GTFX concentration of 1.0×10?5 mol/L. The addition of GTFX did not change the structures of the two different phases fundamentally, but resulted in the transfer and redistribution of GTFX and SDS, so the electric properties of the system were changed correspondingly.  相似文献   

13.
A series of arylantimony ferrocenecarboxylates with the formula (C5H5FeC5H4CO2)nSbAr(5?n) (n = 1, 2; Ar = C6H5, 4‐CH3C6H4, 3‐CH3C6H4, 2‐CH3C6H4, 4‐ClC6H4, 4‐FC6H4) were synthesized and characterized by elemental analysis, IR, 1H NMR and mass spectra. The crystal structures of (C5H5FeC5H4CO2)2Sb(4‐CH3C6H4)3 and C5H5FeC5H4CO2SbPh4 were determined by X‐ray diffraction. Four human neoplastic cell lines (HL‐60, Bel‐7402, KB and Hela) were used to screen these compounds. The results indicate that these compounds at 10 µM show certain in vitro antitumor activities. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

14.
Urea is found to show the hydrotrope action when the aqueous solubility of surfactant CTAB is enhanced while it will show the hydrotrope-solubilization action when the solubilized amount of n-C5H11OH in O/W microemulsion and that of water in W/O microemulsion are increased. The mechanism of the hydrotrope-sotubilization action of urea is in fact the increase of the stability of W/O and O/W microemulsion and structural transition from the lamellar liquid crystal phase to the bicontinuous structure.  相似文献   

15.
We studied the kinetics of the oxidative chemical homopolymerization of 2‐methoxyaniline (OMA) in aqueous acid solutions by monitoring OMA depletion with 1H NMR spectroscopy. We used the same semiempirical kinetic model used for aniline (ANI) homopolymerization to evaluate the experimental data. The reaction kinetics of OMA homopolymerization was similar to that of ANI, although we obtained longer induction and propagation times for OMA. This was attributed to steric hindrance of the bulky methoxy substituent during the coupling reaction. Furthermore, it was suggested that a lower OMA polymerization rate could also be related to a lower concentration of nonprotonated OMA molecules in the reaction solution due to a higher pKa value for OMA than for ANI. This may also explain the lower OMA end conversion (90%) compared with that of ANI (96%). The OMA end conversion was not influenced substantially by reaction conditions; it was lower than 90% only when high acid or low oxidant (oxidant‐deficient oxidant/OMA ratio) concentrations were applied. Because the oxidant took an active part in polymerization, it markedly influenced the polymerization rate, especially the initiation rate. The OMA initiation and propagation rates increased with increasing oxidant and initial monomer concentrations and with the reaction temperature, but there was no uniform trend in the correlation between the homopolymerization rate and acid concentration. The activation energies of the OMA initiation and propagation were 57 and 10 kJ/mol, respectively. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 2471–2481, 2001  相似文献   

16.
The reaction of VO(acac)2 with 2-hydroxyl-1-naphthaldehyde isonicotinyl hydrazone and amines (ethylenediamine or diethylenetriamine) in CH3OH yields crystals of novel vanadium compounds characterized by IR, NMR spectroscopic methods and X-ray single-crystal structure determination. Two different vanadium units exist in the crystal cell of [VO2(C17H11N3O2)][VO- (C4H13N3)(C6H5N3O)](C2H5OH) which crystallizes in the triclinic system, space group P1 with a = 8.0104(17), b = 13.898(3), c = 14.955(3)A, α = 89.103(4), β = 79.551(4), γ = 78.352(4)°, V = 1603.3(6)A^3, Mr = 723.54, Dc = 1.499 g/cm^3, Z = 2, λ(MoKα) = 0.71073 ]A,μ= 0.644 mm^-1, F(000) = 748, the final R = 0.0547 and wR = 0.0997 for 8920 observed reflections with I 〉 2σ(I). According to structure analysis, two different molecules are arranged in the lattice and the two vanadium atoms adopt octahedral and square pyramidal coordination geometries, respectively. The interactions between DNA and vanadium complexes have been investigated by UV-Vis absorption spectro- photometry.  相似文献   

17.
A new nickel(Ⅱ) coordination complex [Ni(2,2'-bipy)2(ClC6H4COO)(H2O)](ClO4) was synthesized by self-assembly reaction of m-chlorobenzoic acid, 2,2'-bipy and Ni(ClO4)2(6H2O. The crystal of the title compound belongs to monoclinic system, space group P21/n with a = 0.77764(14), b = 2.6572(5), c = 1.3637(2) nm, β = 96.456(3)°, V = 2.8000(9) nm3, Dc = 1.528 g/cm3, Z = 4, Mr = 644.10, μ(MoKα) = 0.937 mm-1, F(000) =1320, R = 0.0549 and wR = 0.1164. Structure analysis shows that the nickel(Ⅱ) ion is coordinated with four nitrogen atoms from two bipys as well as two oxygen atoms from m-chlorobenzoic acid and water, respectively, giving a distorted octahedral coordination geometry. The cyclic voltametric behavior of the complex is also presented.  相似文献   

18.
The kinetics of propylene polymerization initiated by ansa‐metallocene diamide compound rac‐Me2Si(CMB)2Zr(NMe2)2 (rac‐1, CMB = 1‐C5H2‐2‐Me‐4‐tBu)/methylaluminoxane (MAO) catalyst were investigated. The formation of cationic active species has been studied by the sequential NMR‐scale reactions of rac‐1 with MAO. The rac‐1 is first transformed to rac‐Me2Si(CMB)2ZrMe2 (rac‐2) through the alkylation mainly by free AlMe3 contained in MAO. The methylzirconium cations are then formed by the reaction of rac‐2 and MAO. Small amount of MAO ([Al]/[Zr] = 40) is enough to completely activate rac‐1 to afford methylzirconium cations that can polymerize propylene. In the lab‐scale polymerizations carried out at 30°C in toluene, the rate of polymerization (Rp) shows maximum at [Al]/[Zr] = 6,250. The Rp increases as the polymerization temperature (Tp) increases in the range of Tp between 10 and 70°C and as the catalyst concentration increases in the range between 21.9 and 109.6 μM. The activation energies evaluated by simple kinetic scheme are 4.7 kcal/mol during the acceleration period of polymerization and 12.2 kcal/mol for an overall reaction. The introduction of additional free AlMe3 before activating rac‐1 with MAO during polymerization deeply influences the polymerization behavior. The iPPs obtained at various conditions are characterized by high melting point (approximately 155°C), high stereoregularity (almost 100% [mmmm] pentad), low molecular weight (MW), and narrow molecular weight distribution (below 2.0). The fractionation results by various solvents show that iPPs produced at Tp below 30°C are compositionally homogeneous, but those obtained at Tp above 40°C are separated into many fractions. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 737–750, 1999  相似文献   

19.
本文通过单晶X-射线衍射法测定了EtEDTB1.4C2H5OH5H2O 1和H4EtEDTB(ClO4)4 C2H5OH 2的晶体结构。晶体学数据如下:1的分子式为C44.8H66.4N10O6.4, Mr = 847.48, 属三斜晶系, 空间群P, a = 11.489 (2), b = 11.866(3), c = 12.002(3) , = 97.47(2), ?= 114.564(13), ?= 114.11(2)? V = 1266.6(5) 3, Z = 1, Dc = 1.111 g/cm3, F(000) = 456, m(MoK? = 0.076 mm-1。共收集衍射数据5207条, 其中独立衍射数据4323条(Rint = 0.0257), 1318条可观测衍射数据(I > 2(I))用于结构计算。结构由直接法解出, 并用全矩阵最小二乘法修正, 最终偏离因子R = 0.0706, wR = 0.1802。分子具有对称中心, 4个苯并咪唑基团围绕中心呈螺旋桨状均匀排布。在1的晶体中, EtEDTB分子通过水和乙醇的氢键相连形成二维网状结构。2的分子式为C44H58Cl4N10O17, Mr = 1140.80, 属单斜晶系, 空间群C2/c, a = 24.260(5), b = 13.040(3), c = 17.680(4) , ?= 97.50(3)? V = 5545.2(2) 3, Z = 4, Dc = 1.366 g/cm3, F(000) = 2384, m(MoK? = 0.289 mm-1。共收集衍射数据12055条, 其中独立衍射数据6360条(Rint = 0.0408), 2875条可观测衍射数据(I > 2(I))用于结构计算。结构由直接法解出, 并用全矩阵最小二乘法修正, 最终偏离因子R = 0.0692  相似文献   

20.
The o‐substituted hybrid phenylphosphines, PPh2(o‐C6H4NH2) and PPh2(o‐C6H4OH), could be deprotonated with LDA or n‐BuLi to yield PPh2(o‐C6H4NHLi) and PPh2(o‐C6H4OLi), respectively. When added to a solution of (η5‐C5H5)Fe(CO)2I at room temperature, these two lithiated reagents produce a chelated neutral complex 1 (η5‐C5H5)Fe(CO)[C(O)NH(o‐C6H4)PPh2C,P‐η2] for the former and mainly a zwitterionic complex 2 , (η5‐C5H5)Fe+(CO)2[PPh2(o‐C6H4O?)] for the latter. Complex 1 could easily be protonated and then decarbonylated to give 4 [(η5‐C5H5)Fe(CO){NH2(o‐C6H4)PPh2N,P‐η2}+]. Complexes 1 and 4‐I have been crystallographically characterized with X‐ray diffraction.  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