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.
Abstract

The liquid-crystalline phases of the system hexaethylene glycol n-dodecyl methyl ether (C12E6C1)/water are studied by deuteron NMR spectroscopy. Information about the molecular orientation is derived from the quadrupole splittings of two selectively deuteriated derivatives of C12E6C1, one deuteriated at the α-position of the alkyl chain, the other at the methoxy group. The temperature dependence of the quadrupole splittings reveals a continuous decrease of orientational order on approach to the macroscopic phase transformations. This behaviour is explained by an increase of defects and fluctuations in the microstructure of the mesophases.  相似文献   

2.
The properties and phase behaviors of the catanionic mixtures consisting of tetradecyltrimetylammonium bromide (TTABr) and different anionic surfactants (i.e., sodium docanoate, C10HOONa; sodium laurate, C12HOONa; sodium perfluorodecanoate, C10FOONa) were examined, in particular when the molar mixing ratio in the aqueous solution was exactly 1:1. Although the three inspected systems have identical head groups and counterions, they exhibited very different lamellar (Lα) phases. When using the hydrogenated surfactants, the C10HOONa–TTABr system formed domain-like Lα/L1 two phases and the C12HOONa–TTABr system formed cream-like Lα/L1 two phases, respectively. In the case of the perfluorinated surfactant, the C10FOONa–TTABr system formed interdigitated and tilted Lα gel. The microstructures of the three Lα phases were characterized by polarized microscope, freeze-fracture transmission electron microscope, small angle X-ray scattering, and X-ray diffraction. The phase transition of the lamellar gel at different temperature was studied by differential scanning calorimetry and rheological measurements. The results elucidated the formation of the Lα phase in catanionic mixtures containing hydrogenated or fluorinated anionic surfactants with molar mixing ratio of 1:1.  相似文献   

3.
The liquid crystals and other phases formed when the mixed surfactant system sodium dodecyl-p-benzene sulphonate (NaLAS) and octa-oxyethylene glycol hexadecyl ether (C16E8, 1:1 by weight) is dispersed in water have been investigated using optical microscopy, X-ray diffraction and differential scanning calorimetry. Despite the fact that neat LAS is a multi-phase solid and C16E8 is a crystalline solid, when the two are mixed at temperatures above the melting temperature of the C16E8 with no water present, what appears to be a metastable gel phase formed containing only a small volume fraction of un-dissolved LAS (ca. 5%). Moreover, when water is added to the system, the phase behaviour of the mixture considerably differs to that of either of the individual components. We report a detailed phase study on this mixture here particularly focussing on the ‘neat’ mixture. The phase behaviour when water is added is also discussed. Particularly interesting is the presence of a micellar phase between the hexagonal and lamellar phases thought to be due to weak interactions between micelles during the transition from rods to discs. In addition, the presence of a low temperature intermediate phase is discussed.  相似文献   

4.
The molecular order and phase transitions of two smectic poly(ester imide)s based on aminobenzoic acid trimellitimide (PEI 1) or aminocinnamic acid trimellitimide (PEI 2) and α,ω-dihydroxydodecane were investigated by X-ray scattering. During cooling, the polymers pass through monotropic smectic liquid-crystalline (LC) phases (SA, SC), which transform into higher-ordered smectic-crystalline phases (SE, SH). The smectic layer structure of about 3 nm gives rise to a sharp reflection at 2θ ≅ 3°. Peak shape analysis and analysis of the interface distribution function revealed long-range longitudinal correlation among the mesogens in the LC phase but short-range lateral correlation. The development of a broad reflection in the small-angle X-ray scattering (SAXS, 2θ < 1°) indicates the formation of a lamellar two-phase system. The long-period changes reversibly between 10 and 30 nm with increasing temperature. The crystalline lamellae comprise a number of smectic-crystalline layers with packed mesogens, while the noncrystalline interlamellar regions keep their smectic-LC order. In the metastable SB phase, formed during annealing of quenched PEI 1, the diffuse SAXS indicates a random distribution of small, probably fringed, crystals with hexagonal-packed mesogens. In the lamellar SE and SH phases, tie molecules play an important role, but chain folding cannot be excluded. Received: 16 July 1999/Accepted: 28 April 2000  相似文献   

5.
The fluorescence spectra of retinol obtained in bilayer structures of two different systems with dodecyl tetraethylenglycol ether are shown. A correlation between the fluorescence intensity of retinol and the different topologies of bilayers has been found. We have tested this correlation with the C12E4/benzyl alcohol/water system, and we have also applied this idea to the study of the lamellar phases of the C12E4/PEG/water system. The highest fluorescence intensity of retinol corresponds to unilamellar vesicles, while the lowest is observed for multilamellar vesicles. The kinetic study of the degradation of vitamin A in these media is also related to the different microstructures of the bilayers.  相似文献   

6.
Phase changes in binary systems of poly(ethylene glycol) dodecyl ethers, C12H25(OC2H4)xOH (x=5, 6 and 8) with water have been studied between –40 to 100 C by differential scanning calorimetry. A number of transitions, including the liquidliquid phase separation, were seen and the transition temperatures and enthalpy changes were determined. The observed temperatures were generally in good agreement with reported phase diagrams. In the C12E5 system, three of the four three-phase lines were seen and a more complete phase diagram is suggested for the water-rich part of the system. Most of the phase changes seen above 0 C are accompanied by small endothermic enthalpy changes of 0.7 to 0.9 kJ (mol C12Ex)–1, independent of system studied, type of transition and transition temperature. Water-rich solutions and liquid crystalline phases separate upon freezing into ice and crystals of hydrated amphiphile. Eutectics are developed at the following temperatures and compositions: C12E5 –3.0 C and 54 wt % C12E5; Q2E6 –4.5 C and 54 wt % C12E6, C12E8 –3.8 C and 49 wt % C12E8.  相似文献   

7.
Differences in the system constants of the solvation parameter model and retention factor correlation plots for varied solutes are used to study the retention mechanism on XBridge C8, XBridge Phenyl and XTerra Phenyl stationary phases with acetonitrile–water and methanol–water mobile phases containing from 10 to 70% (v/v) organic solvent. These stationary phases are compared with XBridge C18 and XBridge Shield RP18 characterized in an earlier report using the same protocol. The XBridge stationary phases are all quite similar in their retention properties with larger difference in absolute retention explained by differences in cohesion and the phase ratio, mainly, and smaller changes in relative retention (selectivity) by the differences in individual system constants and their variation with mobile phase type and composition. None of the XBridge stationary phases are selectivity equivalent but XBridge C18 and XBridge Shield RP18 have similar separation properties, likewise so do XBridge C8 and XBridge Phenyl, while the differences between the two groups of two stationary phases is greater than the difference within either group. The limited range of changes in selectivity is demonstrated by the high coefficient of determination (>0.98) for plots of the retention factors for varied compounds on the different XBridge phases with the same mobile phase composition.  相似文献   

8.
Group III-nitrides are of great interest in both fundamental sciences and technical application. Most of the common nitrides are well known as hard and wide band gap semiconductor materials. In general they have been studied in zinc-blende and wurtzite phases. In this paper, we focus our attention to structural, electronic, phase transition and elastic properties of aluminum nitride (AlN) in zinc-blende and rock-salt phases. A little work has been reported either theoretically or experimentally on elastic and electronic properties of AlN; especially in RS phase. All the calculations are performed using the full-potential linearized augmented plane-wave approach plus local orbitals within the framework of density functional theory as implemented in the Wien2k code. The generalized gradient approximation based on the Perdew–Burke–Ernzerhof is used for the exchange and correlation functional. We determine the full set of first order elastic constants, C11, C12 and C44 at zero pressure to confirm the mechanical stability and hardness, which have not been established either experimentally or theoretically for RS phase. In the study obvious phase transition from ZB phase to RS phase due to pressure effect has been obtained at 12.75 GPa.  相似文献   

9.
The micellization process of binary surfactant mixtures containing cationic surfactants viz. dodecyl pyridinium halide (C12PyX; X=Cl, Br, I), tetradecyl pyridium bromide (C14PyBr), and hexadecyl pyridium halide (C16PyX; X=Cl, Br) and a nonionic surfactants viz. dodecyl nonapolyethylene glycol ether (C12E9), dodecyl decapolyethylene glycol ether (C12E10), dodecyl dodecapolyethylene glycol ether (C12E12), and dodecyl pentadecapolyethylene glycol ether (C12E15) in water at different mole fractions (0–1) were studied by surface tension method. The composition of mixed micelles and the interaction parameter, β evaluated from the CMC data obtained by surface tension for different systems using Rubingh's theory were discussed. Activity coefficient (f1 and f2) of cationic surfactant (CnPyBr)/C12Em (n=12, 14, 16 and m=10, 12, 15) mixed surfactant systems were evaluated, which shows extent of ideality of individual surfactant in mixed system. The stability factors for mixed micelles were also discussed by Maeda's approach, which was justified on the basis of steric factor due to difference in head group of nonionic surfactant.  相似文献   

10.
The adsorption of a series of sodium polystyrene sulfonates with a narrow distribution of molecular weight and ethoxylated non-ionic surfactants ( NP-12, NP-40, NP-100 and C16E20 ) was investigated at carbon black-water interface. Adsorption isotherms of the Langmuir type were obtained for all the samples with negative adsorption free energies. The dependence of maximum of adsorption on molecular weight tor PSSNa was used to obtain information on the macromolecular conformation at the interface; surface geometry effects on adsorption conformation were analyzed using a simplified fractal approach. PSSNa does not adopt a flat conformation at the interface. Non-ionic surfactants ( NP-12, NP-40, NP-100 and C16E20 ) are adsorbed at the interface with the EO chains in a coil conformation.  相似文献   

11.
Phase behaviors of dodecane–hexadecane (n-C12H26–C16H34, C12–C16) binary mixtures in bulk and confined in SBA-15 (pore diameters 3.8, 9.5, and 17.2 nm) are investigated using differential scanning calorimetry. According to the thermal analysis, the bulk mixtures belong to a system of partial miscibility with two solid solutions and a eutectoid in the range of mole fraction $ x_{{{\text{C}}_{ 1 6} }} $  = 0.1–0.8. Under confinement, phase behavior of C12–C16 mixtures is distinct from the bulk. Inside pores of SBA-15 (3.8 nm), the solid mixtures has only a melting boundary. In the pores larger than 9.5 nm, phase behaviors of the mixtures show some resemblance to the bulk system. The growth of the phase diagram with the pore diameter clearly shows the size effect on the phase behavior of the confined mixtures. In comparison with those of chain length difference of pure components of two carbon atoms or less, C12–C16 mixtures exhibit different phase behaviors not only in the bulk but also in the confined state.  相似文献   

12.
The kinetics of alkaline hydrolysis of tris(1,10–phenanthroline)iron(II) has been studied in the presence of nonionic and mixed nonionic–ionic micellar media at 308 K. The effects of mixed-micellar environments of nonionic with ionic surfactants (C12E23/ATABs and C12E23/SDS) on the hydrolytic rate have been studied. The rate decreases monotonically with an increment of [C12E23]T (total Brij 35 concentration) at constant [?OH]0 and has been discussed with the pseudo-phase micellar model. The rate also decreases with [C12E23]T at a continuous addition of ionic surfactants (ATABs and SDS). The observed rate constant kobs follows the empirical relation: kobs = (k0 + θK [C12E23]T)/(1 + K [C12E23]T) (where θ and K are empirical constants). The values of θ remain unaffected, whereas K decreases nonlinearly with [ATABs]T in a mixed C12E23?ATAB micellar system. But the kobs in a mixed C12E23–SDS micellar system is much lower than that of the C12E23–ATAB system and do not comply with any micellar kinetic models.  相似文献   

13.
Pseudo-first-order rate constants, kobs, for the alkaline hydrolysis of N-hydroxyphthalimide, 1, at 0.02 M NaOH and 30°C remain essentially independent of the total concentration of C12E23, [C12E23]T, at ≤0.005 M C12E23. The increase in [C12 E23]T from 0.005 to 0.015 M causes a nonlinear decrease in kobs. The rate of hydrolysis becomes either too slow or the change in absorbance values becomes significantly small to allow a reliable observed data fit to a first-order kinetic equation at ≥0.020 M C12E23 in the absence and presence of total concentration of cetyltrimethylammonium bromide, [CTABr]T ranging from 0.003 to 0.020 M. The values of fraction of nonionized 1, FSH, obtained at reaction time t = 0 and 0.02 M NaOH, remain ~0 at ≤0.010 M C12E23 while they increase from 0.39 to 0.89 with the increase of [C12E23]T from 0.015 to 0.10 M. The values of kobs show a nonlinear decrease of ~5-fold with the increase of [C12E23]T from 0.0 to 0.010 M in the presence of 0.02 M NaOH and [CTABr]T range of 0.003 to 0.020 M. The values of FSH remain ≤~0.10 at ≤0.015 M C12E23 while they vary between 0.40 and 0.90 within a [C12E23]T range 0.02 to 0.05 M in the presence of 0.02 M NaOH and [CTABr]T ranging from 0.003 to 0.020 M. The values of FSH represent the fraction of nonionized 1 trapped almost irreversibly by pure C12E23, and mixed C12E23–CTABr micelles.  相似文献   

14.
采用界面扩张流变技术研究了季铵盐偶联表面活性剂C12-(CH2)2-C12·2Br (Gemini12-2-12)及其与离子液体表面活性剂溴化1-十二烷基-3-甲基咪唑(C12mimBr)复配体系的动态界面张力、扩张流变性质和界面弛豫过程等, 探讨了C12mimBr 对C12mimBr/Gemini12-2-12 混合体系界面性质的影响及C12mimBr 对Gemini12-2-12界面聚集行为影响的机制. 结果表明, 随着离子液体表面活性剂的不断引入, 体系界面吸附达到平衡所需的时间逐渐缩短, 扩张模量和相角明显降低, 界面吸附膜由粘弹性膜转变为近似纯弹性膜; 同时, 界面及其附近的弛豫过程也发生显著变化, 慢弛豫过程消失, 快弛豫过程占主导地位, 且离子液体浓度越高, 快弛豫的贡献越大. 这些界面性质的变化主要归因于离子液体表面活性剂C12mimBr参与界面形成及两表面活性剂在界面竞争吸附的结果. 少量离子液体表面活性剂C12mimBr 的加入可以填补疏松的Gemini12-2-12 界面上的空位, 形成混合界面吸附膜. 随着C12mimBr 含量的增加, 嵌入界面的C12mimBr 分子数不断增多, 导致界面上相互缠绕的Gemini12-2-12烷基链“解缠”, 在体相和界面分子扩散交换的过程中“解缠”的Gemini12-2-12分子从界面上解吸回到体相, 与此同时, C12mimBr 分子相对较小的空间位阻及较强的疏水作用促使其优先扩散至界面进而取代Gemini12-2-12分子, 最终界面几乎完全被C12mimBr分子所占据.  相似文献   

15.
The crystal structures in two solid phases, i.e. phase II stable between 146 and 253 K and phase IV below 136 K, of the title compound [phenazine–chloranilic acid (1/1), C12H8N2·C6H2Cl2O4, in phase II, and phenazinium hydrogen chloranilate, C12H9N2+·C6HCl2O4, in phase IV], have been determined. Both phases crystallize in P21, and each structure was refined as an inversion twin. In phase II, the phenazine and chloranilic acid mol­ecules are arranged alternately through two kinds of O—H⋯N hydrogen bonds. In phase IV, salt formation occurs by donation of one H atom from the chloranilic acid molecule to the phenazine mol­ecule; the resulting monocation and monoanion are linked by N—H⋯O and O—H⋯N hydrogen bonds.  相似文献   

16.
无机金属元素可与中药活性有机化合物通过配位键结合形成配合物,进而影响药物的生理活性[1]。但配合物的形成往往改变了原有机成分的溶解性能(如水溶性或脂溶性),而表面活性胶束体系可使金属有机配合物的水溶性或脂溶性得到明显改善,从而提高药物的生物利用度和改善药物的吸收。然而,胶束体系中已有的研究多是针对单纯的有机化合物[2]或生物分子如蛋白质[3]等,而涉及金属配合物胶束溶液的则很少,尤其是反胶束微环境中的中药金属配合物。反胶束亦称W/O型微乳液,是双亲物质在非极性有机溶剂中自发形成的具有纳米尺寸的含有水核的微小胶团聚集…  相似文献   

17.
18.
Abstract

We have observed the diffusion constants of a dye in several liquid crystals by forced Rayleigh scattering. In a liquid crystal which has a standard phase sequence of N-SA-SC, the diffusion anisotropy changes at the N-SA phase transition and increases with decreasing temperature in SA and SC phases. The diffusion constants exhibit a rather smooth decrease with decreasing temperature except an anomaly at the SA-Sc phase transition. In a liquid crystal which has the antiferroelectric SCA phase, however, the diffusion constants show discontinuous increase and decrease at the SA-Sc and the Sc-ScA phase transition temperatures, respectively: the diffusion constant in SC is larger than that in the higher temperature SA phase. Anomalous signal increase and profile were observed at the phase transition temperatures, and were ascribed to the complementary transient grating due to the coexistence of two phases.  相似文献   

19.
A series of cyanobiphenyl dimers attached via alkoxy spacers to a central malonate were prepared, and the mesomorphic behaviour was studied by differential scanning calorimetry (DSC), polarised optical microscopy and X-ray diffraction (Wide-angle X-ray scattering and Small-angle X-ray scattering). Depending on spacer lengths and substitution of the malonate, nematic and smectic A mesophases with pronounced odd–even effect were observed. C-2-unsubstituted malonates formed nematic phases for chain lengths C6–C14, while C12 and C14 homologues displayed additional smectic A phases. In contrast, malonates with fluorinated tails at C-2 displayed exclusively smectic A phases. Remarkably, the X-ray diffraction profile of the smectic A phase of the C-2-unsubstituted C12 malonate showed a fundamental (001) and the corresponding third-order (003) diffraction peak, but no (002) reflection. Using Fourier analysis, the diffraction pattern was converted to an electron density profile, which was in good agreement with the proposed packing model of the SmA mesophase based on a horseshow- or hairpin-like conformation of the malonate.  相似文献   

20.
A new analytical technique, spinning microtube fluorometry (SMF), was developed and applied to the study of interfacial hydrolysis of 5-dodecanoylaminofluorescein di-β-D-galactopyranoside (C12FDG) by β-galactosidase (β-gal) in the toluene–water system. The nonfluorescent lactone form of C12FDG in the toluene phase was converted at the interface to 5-dodecanoylaminofluorescein (C12F), which was fluorescent in the aqueous phase as a dianion at pH 7.3, though some part of C12F was extracted into the toluene phase as its nonfluorescent lactone form. The distribution ratios of C12FDG and C12F at pH 7.3 were determined as 1.4×102 and 1.97, respectively. The interfacial adsorption constants from the toluene phase to the interface at pH 7.3 were 4.8×10−4 and 1.7×10−2 dm for C12FDG and C12F, respectively. The kinetic experiments with the SMF method concluded that the rate-determining step of the enzymatic hydrolysis at the interface and in the aqueous phase was the 1:1 reaction of C12FDG and β-gal and that the hydrolysis reaction rate constant at the interface at pH 7.3 was 1.84×103 M−1s−1, almost equal to that in the aqueous solution, 1.76×103 M−1s−1. Finally, the SMF method revealed that the contribution of the interfacial reaction to the overall hydrolysis reaction rate of the toluene–water system was as high as 97%.  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