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.
Foaming properties of aqueous solutions of the nonionic surfactant polyoxyethylene dodecyl ether (C12EO n ) were studied at 298 K. Four different EO chain lengths, namely C12EO3, C12EO5, C12EO7, and C12EO9, were considered. The foams obtained from C12EO3 or C12EO5 were extraordinary stable retaining a constant volume for more than 20 h. The presence of lamellar liquid crystalline phases was mainly responsible for the super-stable aqueous foams.  相似文献   

2.
The dynamic interfacial tensions (IFTs) of enhanced oil recovery (EOR) surfactant/polymer systems against n-decane have been investigated using a spinning drop interfacial tensiometer in this paper. Two anionic–nonionic surfactants with different hydrophilic groups, C8PO6EO3S (6-3) and C8PO6EO6S (6-6), were selected as model surfactants. Partially hydrolyzed polyacrylamide (HPAM) and hydrophobically modified polyacrylamide (HMPAM) were employed. The influences of surfactant concentration, temperature, polymer concentration, and oleic acid in the oil on IFTs have been studied. The experimental results show that anionic–nonionic surfactants can form compact adsorption films and reach ultralow IFT (10?3 mN/m) under optimum conditions. The addition of polymer has great influence on dynamic IFTs between surfactant solutions and n-decane mainly by the formation of looser mixed films resulting from the penetration of polymer chains into the interface. The compact surfactant film will also be weakened by the competitive adsorption of oleic acid, which results in the increase of IFT. Moreover, the penetration of polymer chains will be further destroyed surfactant/polymer mixed layer and lead to the obvious increase of IFT. On the other hand, polymers show little effect on the IFTs of 6-6 systems than those of 6-3 because of the hindrance of longer EO chain of 6-6 at the interface.  相似文献   

3.
Cuiying Lin  Li Song  Jianxi Zhao   《Acta Physico》2007,23(12):1846-1850
With p-N,N-dimethylaminobenzonitrile (DMABN) as a probe, the variations of the intensity of its second fluorescence emission (Ia) and the corresponding characteristic wavelength (λa) with the surfactant concentration (c), here the examined surfactants (C12TABr, SDS, C12E23, and C12-3-C12·2Br), were measured by Hitachi F4500 fluorescence spectrophotometer. The results showed that both the break points on the Iac curve and the minimum of the derivative variation corresponding to the λac curve agreed very well with the critical micelle concentration (cmc) of the surfactant in aqueous solution as measured by surface tension technique. Due to strong aggregation of C12-3-C12·2Br in aqueous solution, the information about loose micellar structure could be obtained by its λac curve.  相似文献   

4.
Formation of wormlike micelles (WLMs) in an aqueous mixture of polyoxyethylene cholesteryl ether (ChEOn; where n = 20 and 30) and polyoxyethylene dodecyl ether (C12EOm; where m = 3 and 4) has been reported; rheological and small angle X-ray scattering (SAXS) measurements have been performed in the micellar solutions of ChEOn as a function of C12EOm for the structural elucidation. When lipophilic cosurfactant, C12EOm is added to the micellar solutions of ChEOn, it favors the sphere-to-cylinder transition due to the penetration of C12EOm in the palisade layer of ChEOn micelle accompanying an increase in viscosity. When the concentration of C12EOm is increased, entangled network of WLMs is formed. A strong shear thinning has been observed in highly viscous samples indicating the presence of transient networks. Such samples exhibited viscoelastic behavior and could be described by the Maxwell model with a single stress relaxation mode. A maximum is observed in zero-shear viscosity-C12EOm plot. With further addition of C12EOm, viscosity declines and ultimately a phase separation occurs with the formation of turbid solution of vesicular dispersion. This decline has been interpreted in terms of micellar branching induced by an increase in endcap energy, E c (which is compensated by the formation of branch points, having a mean curvature opposite to that of endcaps). The C12EOm induced one-dimensional micellar growth has been confirmed by SAXS.  相似文献   

5.
A smart lyotropic liquid crystal (LLC) system was prepared to control the diffusion rate of hydrophilic and hydrophobic molecules. The LLC system is composed of a nonionic surfactant (tetraethylene glycol monododecylether; C12EO4) and an anionic azobenzene surfactant (Azo‐surfactant). C12EO4 was the main component of the LLC system. The Azo‐surfactant, which can undergo photo‐isomerization, played the role of trigger in this system. LLC gels formed in a solution comprised of Azo‐surfactant (10 mm ) and C12EO4 (300 mm ). The LLC gels became broken when more Azo‐surfactant was added (e.g., up to 15 mm ) and the viscoelasticity was lost. Surprisingly, when we used UV light to irradiate the 300 mm C12EO4/15 mm Azo‐surfactant sample, the gel was recovered and high viscoelasticity was observed. However, under visible‐light irradiation, the gel became broken again. The gel formation could also be triggered by heating the sample. On heating the 300 mm C12EO4/15 mm Azo‐surfactant sample, the system thickened to a point at which typical gel behavior was registered. When the sample was cooled, the gel broke again. The LLC could be used for controlled release of hydrophilic and hydrophobic molecules, and could be considered as a versatile vehicle for the delivery of actives in systems of practical importance.  相似文献   

6.
The influence of hydrophobic chain length in nonionic surfactants on interfacial and thermodynamics properties of a binary anionic‐nonionic mixed surfactant was investigated. In this study, nonionic surfactants lauric‐monoethanolamide (C12 MEA) and myrisitic‐monoethanolamide (C14 MEA) were mixed with an anionic surfactant, α‐olefin sulfonate (AOS). The critical micelle concentration (cmc), maximum surface excess (Γmax), and minimum area per molecule (Amin) were obtained from surface tension isotherms at various temperatures. The thermodynamic parameters of micellization and adsorption were also computed. Micellar aggregation number (Nagg), micropolarity, and binding constant (Ksv) of pure and mixed surfactant system was calculated by fluorescence measurements. Rubingh's method was applied to calculate interaction parameters for the mixed surfactant systems.  相似文献   

7.
The micellization behavior of an anionic gemini surfactant, GA with nonionic surfactants C12E8 and C12E5 in presence of 0.1 M NaCl at 298 K temperature, has been studied tensiometrically in pure and mixed states, and the related physicochemical parameters (cmc, γ cmc, pC 20, Γ max, and A min) have been evaluated. Tensiometric profile (γ vs log [surfactant]), for conventional surfactants, generally consists of a single point of intersection; a gradually decreasing line (normally linear, or with slight curvature) ultimately saturates in γ at a particular [surfactant], corresponding to complete monolayer saturation. The gemini, in this report, led to two unequivocal breaks in the tensiometric isotherm. An attempt to the interpretation of the two breaks from molecular point of view is provided, depending solely on the chemical structure of the surfactant. The gemini, even in mixed state with the conventional nonionic surfactants C12E5 and C12E8, manifested the dual breaks; of course, the dominance of the feature decreases with increasing mole fraction of the nonionics in the mixture. Theories of Clint, Rosen, Rubingh, Motomura, Georgiev, Maeda, and Nagarajan have been used to determine the interaction between surfactants at the interface and micellar state of aggregation, the composition of the aggregates, the theoretical cmc in pure and mixed states, and the structural parameters according to Tanford and Israelachvili. Several thermodynamic parameters have also been predicted from those theories.  相似文献   

8.
The mutual interactions between nonionic surfactants such as polyoxyethylene cetyl ethers (C16EO n ,n=15, 20) especially their cubic lyotropic liquid crystalline phases of typeI 1 and polymer gelatin were investigated. The colloidal microstructure of such anI 1-phase consisting of close-packed globular surfactant aggregates was shown by transmission electron microscopy (TEM). The diameter of the globules found by TEM correlated well with the periodic distance of about 7.5 nm obtained by small angle x-ray diffraction (SAXD). In ternary systems consisting of surfactant, gelatin, and water cubic liquid crystalline structures were also proved by polarized light microscopy, TEM, and SAXD. The polymer did not participate in the cubic structure but formed, at least in part, anisotropic spherulites. In diluted surfactant systems however, interactions between polymer and surfactants were clearly found by polarimetry. The nonionic surfactants caused an accelerated coil-helix transition of the polypeptid gelatin.  相似文献   

9.
The solution properties of homogeneous hexaethylene and octaethylene glycol mono(n-dodecyl) ethers, C12E6 and C12E8, respectively, and octaethylene glycol mono(n-decyl) ether, C10E8, with poly(methacrylic acid) (PMA) were investigated by dye solubilization, surface tension, fluorescence, viscosity, and pH measurements. The data were discussed regarding non-cooperative and cooperative binding of surfactant to polymer. Whereas in the interaction with poly(acrylic acid) (PAA), the critical aggregation concentrations (cac or T 1) of these surfactants were lower than the respective critical micelle concentration (cmc), in that with the more hydrophobic PMA, T 1’s of C12E6 and C12E8 were higher than the respective cmc, but that of C10E8 was lower than its cmc. These may be ascribed to the hydrophobic microdomains (HMD) of the PMA coil in water, probably in its inside. It is considered that some surfactants are bound first to the HMD non-cooperatively and then they are abruptly bound cooperatively at T 1. This raises T 1 higher than cmc when the cmc is low, and the amount bound by the HMD is relatively large and vice versa. T 1 of C12E6 or C12E8 is the former case, and that of C10E8 is the latter. Thus, different from PAA, T 1 for PMA + nonionic surfactant system consists of the amount of non-cooperative binding and the cac of the cooperative binding in equilibrium. Therefore, this T 1 has a different meaning from that for PAA and should be called apparent T 1. As the binding to the HMD is dependent on PMA concentration and cac is not, which is like in the PAA system, separation of apparent T 1 from the HMD binding was achieved by extrapolating T 1’s to zero PMA concentration (denoted intrinsic T 1). This value for C12E8 was found to be lower than the respective cmc and also lower than the respective T 1 for PAA. With increase in surfactant concentration, the pH of PMA solution rose and demonstrated a peak. This pH rise and fall may be induced by loosening of the HMD coil due to binding increase and by rearrangement of PMA + surfactant complex in high surfactant concentrations region. By raising the initial pH, the HMD were loosened; consequently, T 1 rose a little, and at higher pH, no surfactant binding took place.  相似文献   

10.
Micellar particles can solubilize lipophilic extractants similarly to the organic phase in classical biphasic extraction. This analogy is used here to investigate the kinetics of complex formation between Ni2+ ions and long chain 5-alkoxypicolinic acids (Cn-PIC, withn=12, 15, 18) solubilized in different types of micelles, namely cetyl trimethylammonium bromide (CTAB), hexaethyleneglycol-dodecylether (C12EO6) and CTAB/C12EO6 mixed micelles. In the case of CTAB micelles, the interaction between the carboxylic function of the extractant and the polar head of surfactant molecules was expected to decrease the rate of complex formation so as to make possible kinetic separation of mixtures of metal ions. The observed rate constants for complex formation at pH 4.5 or 7.0 are indeed much smaller in CTAB micelles than in C12EO6 or mixed micelles, but they still remain too high for the previous purpose, although the influence of the surfactant concentration demonstrates, as expected, a much stronger partitioning in the case of CTAB in comparison to C12EO6. On the other hand, it is shown that, once complex formation has occurred the removal of Ni2+ ions can be achieved using ultrafiltration. The yield of extraction increases withn, with the mole fraction of C12EO6, and with the ligand to metal ratio.Institut Nancéien de Chimie Moléculaire (I.N.C.M.)  相似文献   

11.
Interactions between three triblock copolymers of poly (ethylene oxide)‐poly (propylene oxide)‐poly (ethylene oxide), EOmPOnEOm, and the ionic surfactant sodium dodecyl trioxyethylenated sulfonate, C12E3S, in aqueous solutions were investigated with titration microcalorimetry at 293.15 K. Values of enthalpies, entropies, and free energies of interaction have been derived. The thermodynamic data indicate that interactions between EOmPOnEOm and C12E3S decrease with the increase of m/n.  相似文献   

12.
A series of dissymmetric gemini imidazolium surfactants with different spacer length ([CmCsCnim]Br2, m + n = 24, m = 12, 14, 16, 18; s = 2, 4, 6) were synthesized and characterized by 1H NMR and ESI-MS spectroscopy. Their adsorption and thermodynamic properties were investigated by the surface tension and electrical conductivity methods. Consequently, the surface activity parameters (cmc, γcmc, πcmc, pC20, cmc/C20, Γmax, Amin) and thermodynamic parameters (ΔGmθ, ΔHmθ, ΔSmθ) were obtained. The effects of the dissymmetry (m/n) and the spacer length (s) on the surface activity and micellization process of surfactants have been discussed in detail.  相似文献   

13.
The interaction of nonionic triblock copolymers of poly(ethyleneoxide) (PEO) and poly(propyleneoxide) (PPO) (PEOnPPOmPEOn) with a series of cationic surface-active ionic liquids in aqueous solutions have been investigated. The cationic surface-active ionic liquids include 1-alkyl-3-methylimidazolium bromide (CnmimBr, n?=?8, 10, 12, 14, 16) and N-alkyl-N-methylpyrrolidinium bromide (CnMPB, n?=?12, 14, 16). For different polymer-surfactant systems, the critical aggregation surfactant concentration (cac), the surfactant concentration to form free micelles (C m), and the saturation concentration of surfactant on the polymer chains (C 2) were determined using isothermal titration microcalorimetry (ITC) and conductivity measurements. The structure of the formed aggregates depended strongly on the hydrophobicity of the surfactant and the ratio of polymer/surfactant concentration. For C8mimBr, there were not any micelle-like surfactant?Cpolymer clusters detected in the solution, and only micelles appeared. For other surfactants, the polymer?Csurfactant aggregates were formed in the solution, which was verified by the appearance of a broad endothermic peak in the ITC thermograms. The intensity of polymer?Csurfactant interaction increased with the hydrophobicity of the surfactants and the polymers but was not affected by the surfactant headgroups.  相似文献   

14.
The distribution of 1,10-phenanthroline (phen) in micellar solutions of the nonionic surfactants Triton X and C12E n with varying poly (ethylene oxide) chain lengths has been studied by potentiometry, calorimetry, and fluorometry at 298 K. Micelles accommodate 1,10-phenanthroline according to the reaction, phen + Ym = Ym(phen), where Ym denotes a surfactant molecule aggregated in micelles. The constant K m for the reaction of Triton X increases as a linear function of n*, the number of ethylene oxide (EO) groups, as K m = KEO n* + Kc. Nonzero K EO and K c values suggest a heterogeneous inner structure of the micelle, i.e., the hydrophobic core surrounded by a hydrophilic poly (ethylene oxide) (PEO) shell. On the basis of molar volumes, the intrinsic thermodynamic parameters of transfer of 1,10-phenantholine were extracted. The enthalpy and entropy of transfer of 1,10-phenanthroline from the PEO shell to the core are found to be small and negative. By using K EO and K c values for C12E n obtained by fluorometry, individual fluorescence spectra of 1,10-phenanthroline in the PEO shell and core were extracted. The fluorescence intensity of 1,10-phenathroline accommodated in the core, like in organic solvents, is significantly reduced relative to that in water. These facts indicate that the aromatic rings of 1,10-phenanthroline penetrate into the hydrophobic core, while its hydrophilic N site is still hydrated in the PEO shell.  相似文献   

15.
Dependences of the surface tension of aqueous solutions of ionic (dodecylpyridinium bromide, sodium dodecylsulfonate) and nonionic (Triton X‐100) surfactants and their mixtures on total surfactant concentration and solution composition were studied, and the surface tension of the mixed systems were predicted using different Miller's model. It was found that how to select the model for calculation of ω is corresponding to the degree of the deviation from the ideality during the adsorption of mixed surfactants. The compositions of micelles and adsorption layers at air‐solution interface as well as parameters (βm, βads) of headgroup‐headgroup interaction between the molecules of ionic and nonionic surfactants were calculated based on Rubingh model. The parameters (B1) of chain‐chain interaction between the molecules of ionic and nonionic surfactants were calculated based on Maeda model. The free energy of micellization calculated from the phase separation model (ΔG 2 m ), and by Maeda's method (ΔG 1 m ) agree reasonably well at high content of nonionic surfactant. The excess free energy ΔG ads E and ΔG m E (except α=0.4) for TX‐100/SDSn system are more negative than that TX‐100/DDPB system. These can be probably explained with the EO groups of TX‐100 surfactant carrying partial positive charge.  相似文献   

16.
The interactions between nonionic surfactants such as polyoxyethylene cetyl ethers (C16EO20) and the polymer gelatin were investigated by rheological methods. Capillary viscosity measurements of diluted aqueous solutions confirmed the results of previous investigations [1]. C16EO20 caused an accelerated renaturation of the polypeptide gelatin.A strong variation of viscoelastic data was measured by vibration rheometry when gelatin was added to the lyotropic cubic liquid crystalline phases formed by the surfactant used. Loss module (E) and loss factor (tan ) increased with increasing content of the polymer indicating an obstruction of the coupling between the globular surfactant associates.  相似文献   

17.
A potentiometric technique based on surfactant ion selective electrode has been used for various cationic and anionic surfactants. The data obtained contain m 1 (surfactant monomer concentration); m 2 (free counterion concentration) and α (degree of dissociation of micelle) were used for determination of aggregation number at and above cmc (critical micelle concentration). Data fitting show a relationship between aggregation number with such parameters. The correlation equation obtained shows that size of ionic micelle vary sharply after cmc. Also, the equation obtained shows size of micelle growth with increase in counterion concentration.  相似文献   

18.
We have studied the phase behavior and rheological property of the cubic phase and related gel emulsions in water/nonionic/dodecane systems. In the phase behavior study, it is pointed out that the formation of the discontinuous cubic phase (I1) is not common in all nonionic surfactant systems; however, a cubic phase (I1) with oil-swollen micelles or a cubic phase microemulsion is found in the water/C16EO6/dodecane system, which can solubilize large amount of oil. It was also observed that water/C16EO6/dodecane system forms stable gel emulsion. In the rheological study we have found an anomalous behavior of the I1 phase in the water/C12EO6/dodecane and the water/C16EO6/dodecane systems. In the water/C12EO6/dodecane system, the viscoelastic nature of the I1 phase has been observed, which is shifted to the elastic nature with the addition of dodecane, whereas, highly elastic nature was observed in the water/C16EO6/dodecane system. In both the cases shear-thinning behavior were seen. The elastic modulus, G′ and complex viscosity, |η1| of the I1 phase increase with the dodecane concentration in the water/C12EO6/dodecane system, whereas, decreasing trend have been observed in the water/C16EO6/dodecane system. This anomalous behavior is suggested due to the nonspherical shape of micelles or polydispersity of the micelles in the water/C16EO6/dodecane system. The rheological behavior of the O/I1 gel emulsion was also studied in both the systems.  相似文献   

19.
IntroductionInrecentyears ,bis(quaternaryammonium)surfac tantsorgeminisurfactants ,inwhichtwocationicsurfac tantmoietiesareconnectedwiththeammoniumheadgroupbyaploymethylenechain ,namely ,aspacerhavebecomeofinterestduetotheirexceptionalsurfaceactivityandrem…  相似文献   

20.
The dentritic quaternary ammonium salt-type tetrameric surfactant (4C12tetraQ) was synthesized, and the molecular structure was confirmed by 1H NMR and FTIR. The surface activity of 4C12tetraQ was investigated by surface tension, and surface chemical parameters, such as critical micelle concentration (cmc), efficiency (pC20), effectiveness (πcmc), the surface tension value at cmc (γcmc), minimum surface area (Amin), maximum surface excess (Γmax), and cmc/C20 were obtained from the measurement results. The results show that the 4C12tetraQ surfactant has higher surface activity than the traditional monomeric surfactants (dodecyl trimethyl ammonium bromide, DTAB). The Krafft points were taken as <0°C, indicating that the synthesized tetrameric surfactants had good water solubility. Free energies of micellization and adsorption show that 4C12tetraQ display greater propensity to absorb at the interface than form micelle in the bulk of the aqueous solution, and that the two processes are spontaneous. The measurement results show that 4C12tetraQ has good emulsification power and foam performance. The corrosion efficiency was evaluated with the loss weight method in 1?mol/L HCl solution, and the results show that the 4C12tetraQ surfactant has good corrosion inhibition, and can be considered as a corrosion inhibitor.  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