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.
彭向东  周祖康 《化学学报》1986,44(6):613-615
属于ABA型嵌段共聚物(A为聚氧乙烯,B为聚氧丙烯)的聚醚型表面活性剂分子在水中是否生成胶团,文献报道的结果迥异,临界胶团浓度值随测定方法不同出入很大.某些AB型或ABA型嵌段共聚物在选择性有机溶剂中表现出异常胶团化行为,但对于水体系迄今未见报道.本文以多种实验手段研究了典型的聚醚型表面活性剂Pluronic L-64水溶液的胶体与表面性质.结果表明,与通常的表面活性剂不同,随着温度或浓度的变化,L-64溶液中生成单分子胶团或者缔合胶团.在两者的转变间,首次观察到了水体系中的异常胶团化现象.  相似文献   

2.
We have investigated the formation of threadlike micelles consisting of anionic surfactants and certain additives in aqueous solution. Threadlike micelles long enough to be entangled with each other were formed in a clear aqueous solution of two anionic surfactants, sodium hexadecyl sulfate and sodium tetradecyl sulfate. These solutions also contained pentylammonium bromides or p-toluidine halides and exhibited remarkable viscoelasticity. Because the molar ratio of surfactants to cationic additives in these micelles seemed close to unity, they formed 1:1 stoichiometric complexes between surfactant anions and additive cations, as previously found in systems of cationic surfactants such as hexadecyltrimethylammonium bromide and sodium salicylate. The viscoelastic behavior of these anionic threadlike micellar systems was adequately described by a simple Maxwell element with a single relaxation time and strength, as in many similar cationic systems.  相似文献   

3.
The micellar properties of dilute solutions of two ionic detergents are investigated by quasielastic scattering of laser light. The interpretation of the results in terms of Tanford's theory shows that micelles have oblate ellipsoidal shape near the critical micelle concentration. Evidence for micelle polydispersity is presented, and an approximate evaluation of the spread of the size distribution function is given.  相似文献   

4.
Aggregation and micelle formation of ionic liquids in aqueous solution   总被引:1,自引:0,他引:1  
Association of ionic liquids possessing n-octyl moiety either in the cation or in the anion has been studied in aqueous solution with conductivity and turbidity measurements as well as using 2-hydroxy-substituted Nile Red solvatochromic probe. 1-Butyl-3-methylimidazolium octyl sulfate was found to act as a surfactant above 0.031 M critical micelle concentration. In contrast, 1-methyl-3-octylimidazolium chloride produced inhomogeneous solution of larger aggregates, which were dissolved on the addition of more than 2:1 molar excess of sodium dodecyl sulfate (SDS) due to mixed micelle formation. Even small amount (<10 mM) of ionic liquids could markedly reduce the polarity of the Stern layer of SDS micelle.  相似文献   

5.
Aggregation behavior of dodecyldimethyl-N-2-phenoxyethylammonium bromide commonly called domiphen bromide (DB) was studied in aqueous solution. The Krafft temperature of the surfactant was measured. The surfactant has been shown to form micellar structures in a wide concentration range. The critical micelle concentration was determined by surface tension, conductivity, and fluorescence methods. The conductivity data were also employed to determine the degree of surfactant counterion dissociation. The changes in Gibb's free energy, enthalpy, and entropy of the micellization process were determined at different temperature. The steady-state fluorescence quenching measurements with pyrene and N-phenyl-1-naphthylamine as fluorescence probes were performed to obtain micellar aggregation number. The results were compared with those of dodecyltrimethylammonium bromide (DTAB) surfactant. The micelle formation is energetically more favored in DB compared to that in DTAB. The 1H-NMR spectra were used to show that the 2-phenoxyethyl group, which folds back onto the micellar surface facilitates aggregate formation in DB.  相似文献   

6.
The concentration dependences of the Fourier transform infrared spectra of aqueous solutions of sodiumn-pentanesulfonate,n-hexanesulfonate,n-heptanesulfonate, andn-octanesulfonate were studied in concentration ranges encompassing the critical micellization concentrations (c.m.c). Changes in wavenumber, bandwidth, and intensity of infrared absorption bands were used to monitor the changes in molecular association with concentration. The premicellar aggregation below the c.m.c. may accompany the association of the alkyl chain, but not the counter-ion binding with the SO 3 group. Above the c.m.c, the front location of the counter-ion against the SO 3 group at the micelle surface is demonstrated.  相似文献   

7.
Aqueous solutions of three kinds of surface active ionic liquids composed of the 1-alkyl-3-methylimidazolium cation have been investigated by means of surface tension and electrical conductivity measurements at room temperature (298 K). The surface tension measurements provided a series of parameters, including critical micelle concentration (cmc), surface tension at the cmc (gammacmc), adsorption efficiency (pC20), and effectiveness of surface tension reduction (Picmc). In addition, with application of the Gibbs adsorption isotherm, maximum surface excess concentration (Gammamax) and minimum surface area/molecule (Amin) at the air-water interface were estimated. The effect of sodium halides, NaCl, NaBr, and NaI, on the surface activity was also investigated. It was found that both the pC20 and the Picmc were rather larger than those reported for traditional ionic surfactants and the cmc values were somewhat lower than those for typical cationic surfactants, alkyltrimethylammonium bromides, and comparable to typical anionic surfactants, sodium alkyl sulfates. These results demonstrate that the surface activity of long-chained imidazolium IL is somewhat superior to that of conventional ionic surfactants.  相似文献   

8.
Electrical conductivity was measured for aqueous solutions of long-chain imidazolium ionic liquids (IL), 1-alkyl-3-methylimidazolium bromides with C(12)-C(16) alkyl chains. The break points appeared in specific conductivity (kappa) vs concentration (c) plot indicates that the molecular aggregates, i.e., micelles, are formed in aqueous solutions of these IL species. The critical micelle concentration (cmc) determined from the kappa vs c plot is somewhat lower than those for typical cationic surfactants, alkyltrimethylammonium bromides with the same hydrocarbon chain length. The electrical conductivity data were analyzed according to the mixed electrolyte model of micellar solution, and the aggregation number, n, and the degree of counter ion binding, beta, were estimated. The n values of the present ILs are somewhat smaller than those reported for alkyltrimethylammonium bromides, which may be attributed to bulkiness of the cationic head group of the IL species. The thermodynamic parameters for micelle formation of the present ILs were estimated using the values of cmc and beta as a function of temperature. The contribution of entropy term to the micelle formation is superior to that of enthalpy term below about 30 degrees C, and it becomes opposite at higher temperature. This coincides with the picture drawn for the micelle formation of conventional ionic surfactants.  相似文献   

9.
The self-assembly into wormlike micelles of a poly(ethylene oxide)-b-poly(propylene oxide)-b-poly(ethylene oxide) triblock copolymer Pluronic P84 in aqueous salt solution (2 M NaCl) has been studied by rheology, small-angle X-ray and neutron scattering (SAXS/SANS), and light scattering. Measurements of the flow curves by controlled stress rheometry indicated phase separation under flow. SAXS on solutions subjected to capillary flow showed alignment of micelles at intermediate shear rates, although loss of alignment was observed for high shear rates. For dilute solutions, SAXS and static light scattering data on unaligned samples could be superposed over three decades in scattering vector, providing unique information on the wormlike micelle structure over several length scales. SANS data provided information on even shorter length scales, in particular, concerning "blob" scattering from the micelle corona. The data could be modeled based on a system of semiflexible self-avoiding cylinders with a circular cross-section, as described by the wormlike chain model with excluded volume interactions. The micelle structure was compared at two temperatures close to the cloud point (47 degrees C). The micellar radius was found not to vary with temperature in this region, although the contour length increased with increasing temperature, whereas the Kuhn length decreased. These variations result in an increase of the low-concentration radius of gyration with increasing temperature. This was consistent with dynamic light scattering results, and, applying theoretical results from the literature, this is in agreement with an increase in endcap energy due to changes in hydration of the poly(ethylene oxide) blocks as the temperature is increased.  相似文献   

10.
Both the critical solution temperature (CST, or the Krafft temperature) and the critical solution pressure (CSP, or the Tanaka pressure) were determined for sodium perfluorodecanoate (NaPFDe) in water, and the result shows that the Krafft temperature is raised with the increase in the Tanaka pressure. A thermodynamic analysis has been made on the data for the critical micellization concentration (cmc) and of the solubility at various temperatures and pressures. The estimated change in the partial molal volume, resulting from micelle formation from the singly dispersed state and from the hydrated solid state, was found to be conspicuously higher for NaPFDe compared to hydrocarbon surfactants. This has been ascribed to the more pronounced role of carbon chain-water interactions and water structure effects of the fluorocarbon surfactants.  相似文献   

11.
A series of gradient copolymers of methacrylic acid (MAA)/methyl methacrylate (MMA) with four end-to-end composition profiles (uniform, linear gradient, triblock with linear gradient midblock, and diblock) but all having an average chain composition of ?F(MMA) ≈ 0.5 and an average chain length of 200 were synthesized via model-based, computer-programmed, semibatch atom-transfer radical copolymerization (ATRcoP). These samples allowed us to investigate systematically the effects of the gradient composition profile on the pH responsivity and micelle formation of the copolymers in an aqueous solution. Measurements included light transmittance, TEM, AFM, DLS, (1)H NMR, and pH titration. It was found that linear gradient, triblock, and diblock copolymers formed spherical micelles at high pH. The micelles of the linear gradient copolymer contained MMA units in their hydrophilic shells, and those of the triblock and diblock copolymers had all of their MMA units residing in their cores. The composition profile showed a strong effect on the degree of acid dissociation at a given pH. The conformational transition of the copolymer chains was determined by both the pH value and composition profile. Copolymers having sharper gradients required a lower pH to trigger the conformational transition and a narrower pH range to complete the transition.  相似文献   

12.
A thermodynamic treatment of the volumetric behavior of surfactant mixtures in water have been developed on the basis of the thermodynamic treatment of mixed micelle by Motomura et al. Densities of aqueous solutions of mixtures of decyltrimethylammonium bromide (DeTAB) and dodecyltrimethylammonium bromide (DTAB) have been measured as a function of total molality at constant compositions. The apparent molar volumes of the mixtures have been derived from the density data and the mean partial molar volume of monomeric surfactant mixture V t w , the molar volume of mixed micelle VM/N t M , the voluem of formation of mixed micelle W M V, and the composition of surfactant in the mixed micelle have been evaluated. The V t W , VM/N t M , and W M V have been observed to depend on the composition. The linear dependence of V t W and VM/N t M on the composition indicates that the mixing of DeTAB and DTAB is ideal both in the monomeric and micellar states. This has been confirmed further by the shape of the critical micelle concentration vs. composition curves.  相似文献   

13.
Critical micelle concentrations were determined by conductometry for the homologous series of sodium monoalkyl sulfosuccinates ROOCCH2CH(SO3Na)COONa (R is an alkyl radical, C10–C15) in the presence of sodium chloride. Coefficients of the Corrin-Harkins equation were calculated. It was shown that the formation of micelles began when a certain value of the mean ionic activity of the surface active substance was reached.  相似文献   

14.
本文研究了盐存在时不同比例的十二烷基硫酸钠(简称12CH)和溴化正辛基三甲基铵(简称C8NBr)混合物的表面活性、表面吸附以及胶团形成等性质,结果表明:(1)正、负离子表面活性剂混合物具有很高的表面活性,不论其混合比例如何,临界胶团浓度(cmc)及cmc时溶液的表面张力(γcmc)皆较任何单一组分时小;(2)不论体相中比例如何,表面层中12CH和C8NBr的饱和吸附量的摩尔比皆~1.7且总饱和吸附量亦皆~5.2x10[-10]mol.cm[-2].由此求得表面层中分子截面积为32A[2],与由分子结构计算的数据相近,说明正、负表面活性离子排列紧密;(3)与碳链长相同的正、负离子型表面活性剂混合水溶液比较,本体系反应离子浓度对cmc有明显影响,证实表面层带电,胶团也带电;(4)计算了离子强度相同,温度不同时和温度相同、离子强度不同时的热力学量,得出离子强度大者易形成胶团。  相似文献   

15.
Potentiometric analyses indicate that previous investigations have overestimated the stability of ferric borate complexes. The FeB(OH) 4 2+ formation constant result obtained in the present work isBβ 1 * = [FeB(OH) 4 2+ ][H+][Fe3+]-1[B(OH)3]-1 = (5.4±0.3) x 10-3 at 25.0°C and 0.7 molal ionic strength. Our result indicates that solution concentrations of FeOH2+ and FeB(OH) 4 2+ are approximately equal in aqueous solution for boric acid concentrations on the order of 0.3 molal. Fe(B(OH)4) 2 + is a minor species in solution compared to FeB(OH)4 2+ for conditions such that [B(OH)3][H+]-1≤ 350, and ferric borate complexation is insignificant in solutions such as seawater where [B(OH)3] ≤ 4× 10-4 molal.  相似文献   

16.
In a recent study, we showed that the surfactant 1,2-distearoyl-sn-glycero-3-phosphatidylethanolamine-N-[methoxy(polyethylene glycol)-2000 (DSPE-PEG2000) induced mixed micelles of either threadlike or discoidal shape when mixed with different types of lipids. In this study, we have exchanged the PEG-lipid for the more conventional surfactants octaethylene glycol monododecyl ether (C12E8), hexadecyltrimethylammonium bromide (CTAB), and sodium dodecyl sulfate (SDS). Cryo-TEM investigations show that also these surfactants are able to induce the formation of long-lived discoidal micelles. Generally, the preference for either discoidal or threadlike micelles can be tuned by the choice of lipids and environmental conditions in much the same way as observed for the lipid/PEG-lipid system. Our investigation showed, furthermore, that the choice of surfactant may influence the type of mixed micelles formed. It is argued that the formation of discoidal rather than threadlike micelles may be rationalized as an effect of increasing bending rigidity. Our detailed theoretical model calculations show that the bending rigidity becomes significantly raised for aggregates formed by an ionic rather than a nonionic surfactant.  相似文献   

17.
The enthalpy of micelle formation of ionic surfactants in aqueous solutions with varying concentration of additives (i. e. NaCl. alcohols) was measured by using a very sensitive microcalorimeter. The heats of dilution were measured from concentrations below the c.m.c. to above c.m.c. This enables one to detect the aggregation process around the c.m.c. region, an analysis hitherto not extensively reported in the current literature. The effect of electrolyte addition on micellar enthalpy is discussed, as compared to the addition of short chain alcohols.  相似文献   

18.
Micelle formation of various surfactants, such as sodium caprylate, sodium laurate, sodium palmitate and sodium stearate has been studied in organic solvents of various dielectric constants and intermolecular H-bonding capability, viz. molten acetamide, N-methyl acetamide (NMA) and N,N-dimethyl acetamide (DMA), at different temperatures by electrical conductivity and surface tension methods. Both methods show that micelles are formed in acetamide, NMA and DMA. Gibbs energy changes, enthalpies and entropies of micelle formation, respectively, have been determined by studying the variation of critical micelle concentration (c.m.c.) with temperature. Micelle formation in these solvents has been explained on the basis of several factors such as dielectric constant of the medium, its intermolecular H-bonding capability including solvophobic interaction.  相似文献   

19.
Strong experimental and theoretical evidence was provided on the controlled formation of the two-dimensional J-aggregates that were assembled in the herringbone morphology. The exciton-band structure formation of 1,1',3,3'-tetraethyl-5,5',6,6'-tetrachlorobenzimidazolocarbocyanine (TTBC) J-aggregates was investigated in ionic (NaOH) aqueous solution at room temperature. The control was achieved by changing the [TTBC] at a given [NaOH], or vice versa, and was monitored through the changes in the absorption, fluorescence excitation, and emission spectra. Specific attention was paid to expose the excited-state structure and dynamics through simulations of the excitonic properties, which included diagonal energetic disorder and phonon-assisted exciton relaxation. Aggregates were characterized by an asymmetrically split Davydov pair, an H-band (approximately 500 nm, 1300 cm(-1) wide, Lorentzian-like) and a J-band (approximately 590 nm, 235 cm(-1) wide, with a band shape typical of a one-dimensional J-aggregate), whose relative intensities showed a strong dependence on the [TTBC]/[NaOH]. The H-band is favored by high [TTBC] or high [NaOH]. An explanation of the control on the aggregate formation was given by correlating the changes in the absorption with the structural modifications and the subsequent changes in the dynamics, which were induced by variations in the dye and NaOH concentrations. The J-band shape/width was attributed to disorder and disorder-induced intraband phonon-assisted exciton relaxation. The intraband processes in both bands were estimated to occur in the same time scale (about a picosecond). It has been suggested that the wide energetic gap between the Davydov split bands (3000 cm(-1)) could get bridged by the excitonic states of the loosely coupled chains, in addition to the monomeric species at low [TTBC]. Phonon-assisted interband relaxation, through the band gap states and/or directly from the H- to the J-band, are suggested for accounting the difference between the bandwidths and shapes of the two bands. Energy transfer between the H-band and the monomeric species is suggested as crucial for tuning the relative strengths of the two bands.  相似文献   

20.
Zinc complexes of 3-hydroxymethyl-131-oxo-chlorins possessing a vinyl or its trans-substituted group at the 20-position were prepared through regioselective bromination and successive palladium-catalyzed cross-coupling reactions from naturally occurring chlorophyll-a. The synthetic zinc chlorophyll-a derivatives self-aggregated in an aqueous Triton X-100 micelle solution to give red-shifted and broadened electronic absorption bands, which were similar to those of bacteriochlorophyll-c/d self-aggregates in the main light-harvesting antenna systems of green photosynthetic bacteria. Because of the sterically demanded perpendicular orientation of the (un)substituted ethenyl group to the chlorin π-system, the present model compounds had less J-aggregation ability in a slipped π–π interaction fashion than the corresponding saturated ethyl-type analogs, while halogenations at the terminal of the 20-vinyl group enhanced the ability due to their electron-withdrawing effect.  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