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.
The influence of adding alkyldimethylamine oxide (CnDMAO) with varying alkyl chain lengths (nc) on the acid soap formation of oleic acid was investigated. The solutions of equimolar mixtures of CnDMAO and sodium oleate (Na+Ol), each 25 mmol kg–1, became turbid at a certain critical pH (pHc) on decreasing pH. Values of the pHc depended on nc and showed the minimum at C10DMAO/NaOl mixture. The presence of the minimum was interpreted in terms of two different kinds of the complex formed in the micelles depending on nc: the catanionic complex (CnDMAOH+/Ol) in the mixed micelles of nc=16, 14, 12 and 10, and the acid soap of oleic acid for C6DMAO/NaOl and C8DMAO/NaOl mixtures. At pHc where the amounts of these complexes of double-chain nature reached certain critical values in the mixed micelles, a phase separation (most probably lamella formation) took place. It was expected that the critical amount of the catanionic complex was smaller for the mixtures of higher nc values and hence pHc increased with nc for the mixtures nc10. For the mixtures of nc<10, it was expected that the amount of the acid soap in the mixed micelles increased with decreasing nc at a given pH and the pHc increased with decreasing nc. Micelle compositions at cmc were evaluated on the basis of the regular solution theory coupled with the pseudo phase approximation. The micelle compositions at 100 mmol kg–1 were examined with 13C-NMR. The results showed the mixed micelle formation for nc=16–10, while the micelles mostly consisting of oleic acid for the mixtures of nc=8 and 6. The assumption of two different complexes for the two groups of the mixture was thus supported. The cmc range of mixed micelles was evaluated and it was well correlated with the observed concentration range of pyrene fluorescence change.  相似文献   

2.
A Mg2+-induced vesicle phase was prepared from a mixture of tetradecyldimethylamine oxide (C14DMAO) and magnesium dodecyl sulfate [Mg(DS)2] in aqueous solution. Study of the phase behavior shows that at the appropriate mixing ratios, Mg2+–ligand coordination between C14DMAO and Mg(DS)2 results in the formation of molecular bilayers, in which Mg2+ can firmly bind to the head groups of the two surfactants. The area of the head group can be reduced because of the complexation. In this case, no counterions exist in aqueous solution because of the fixation of Mg2+ ions to the bilayer membranes. Therefore, the charges of the bilayer membranes are not shielded by salts. The birefringent solutions of Mg(DS)2 and C14DMAO mixtures consist of vesicles which were determined by transmission electron microscopy (TEM) images and rheological measurements. Magnesium oxide (MgO) nanoplates were obtained via the decomposition of Mg(OH)2 which were synthesized in Mg2+-induced vesicle phase which was used as the microreactor under the existence of ammonia hydroxide. The morphologies and structures of the obtained MgO nanoplates have been characterized by X-ray diffraction (XRD) and scanning electron microscopy (SEM). The results indicate that the crystal growth is along the (1 1 1) direction which can be affected by the presence of a vesicle phase having a fixation of Mg2+ ions to the bilayer membranes.  相似文献   

3.
We have investigated the phase behavior and self-assembled structures of diglycerol monolaurate-and monomyristate (abbreviated as C12G2 and C14G2, respectivley) in olive oil over a wide range of temperatures and compositions. At lower temperatures, both the surfactants appear in solid state (α-solid), which does not swell with olive oil. The α-solid transforms into lamellar liquid crystal (Lα) phase upon heating and the solid melting temperature is practically constant at all surfactant/oil compositions, but the C12G2 melts earlier than the C14G2. There appear the dispersions of Lα phase and α-solid in the dilute regions of the C12G2/olive oil and the C14G2/olive oil systems, respectively, at 25°C. The Lα phase can solubilize some amount of olive oil, but as the oil concentration increases the excess oil separates out from the Lα phase, and there appears Lα dispersion in the dilute surfactant concentration region. The Lα phase eventually transforms into isotropic solutions (reverse micelles) with further heating. The structures (shape and size) of the reverse micelles have been characterized by small-angle x-ray scattering technique. It has found that the C12G2 and C14G2 surfactants form reverse rod-like micelles in olive oil above the Lα melting temperature and the micellar size increases with surfactant concentration, but decreases with temperatures.  相似文献   

4.
Solubilization of cholesterol, differential scanning calorimetric (DSC), nuclear magnetic resonance (NMR) and dynamic light scattering (DLS) measurements were performed in order to reveal the dispersion mechanisms of stratum corneum (SC) into each intact corneocytes in the following systems: (1) in the aqueous mixed solutions of sodium dodecyl sulfate (SDS) and N,N-dimethyldodecylamine oxide (C12DMAO); (2) in the aqueous micellar solutions of C12DMAO containing solubilized α-terpineol (α-T); and (3) in the aqueous micellar solutions of C12DMAO containing solubilized limonene. The intercellular lamellar structure of SC was revealed to be disrupted and/or removed in all these solutions. However, considering the micellar sizes and the interaction among molecules in micelle, the dispersion mechanisms in these three systems were different each other. The three dispersion mechanisms of SC were estimated and discussed on the basis of the results of solubilization, DSC, NMR and DLS, respectively.  相似文献   

5.
Binding and distribution properties of trimethoprim (TMP) in the presence of various anionic surfactants; sodium octyl sulfate (C8SO4Na), sodium decyl sulfate (C10SO4Na), sodium lauryl sulfate (C12SO4Na), and sodium tetradecyl sulfate (C14SO4Na) has been studied by conductivity, spectrophotometry and surface tension measurements. The surface properties of anionic surfactants, that is, maximum surface excess concentration (Γ max ) and minimum area per surfactant molecule (A min ) at the air/water interface have been evaluated in the absence and presence of TMP using Gibbs adsorption isotherm. From conductivity data the ionization degree and counterion binding parameter have been obtained. Spectrophotometric experiments were used to determine binding constants of TMP to anionic micelles. With the increasing alkyl chain of surfactants, the interaction becomes stronger, which shows the importance of hydrophobic forces and incorporation of TMP molecules to the pure micelles of anionic surfactants increased. The results obtained from the surface tension and conductometric studies have been correlated with those obtained from the spectroscopic studies and binding tendency of TMP to anionic micelles followed the order as: C14SO4Na > C12SO4Na > C10SO4Na > C8SO4Na. From these results, the study of the interaction TMP in different anionic micellar solutions provided information about the characteristics of binding properties of poorly soluble drugs.  相似文献   

6.
A flexibility parameter, the persistence length, has been evaluated from the radii of gyration and the contour lengths for rodlike micelles of heptaoxyethylene alkyl ethers (C n E7,n=12, 14, 16) and tetradecyldimethylammonium chloride (C14DAC) and bromide (C14DAB) at the observed crossover concentrations between dilute and semidilute regimes. The persistence length range is 43–73 nm, except for C12E7, for which it is 32 nm. The crossover concentrations between dilute and semidilute regimes for the semiflexible rodlike micelles calculated according to Ying and Chu as a function of the molecular weight, the contour length, and the persistence length are consistent with the observed values. The crossover concentration between semidilute and concentrated regimes was, on the other hand, calculated by using the same micelle parameters, including the value of thickness of cross-section of the rodlike micelles. The obtained values are at variance with the observed values. This means that rodlike micelles in semidilute and concentrated solutions might differ in size and/or flexibility from those in dilute solution.  相似文献   

7.
The results of the synthesis and characteristics of the new lyotropic lanthanide-containing liquid-crystalline systems possessing the nematic phase based on the zwitterionic surfactant, N,N-dimethyldodecylamine oxide (C12DMAO), in the aqueous-decanol environment are presented. The phase diagrams are constructed. The formation of the La-O coordination bonds in the liquid-crystalline complex C12DMAO/LaIII is confirmed by IR spectroscopy.  相似文献   

8.
This paper reports a study on the aggregation and rheological behavior of the family of O, O’-bis(sodium 2-alkylcarboxylate)-p-dibenzenediol (referred to as Cm?2Cm, m?=?10, 12, 14, respectively) in aqueous solution using dynamic light scattering, 1H NMR and rheology measurements. The results showed that all three surfactants formed large network-like aggregates at low concentrations. However, C10?2C10 formed small compact micelles simultaneously but neither C12?2C12 nor C14?2C14 did. These network-like aggregates were transformed into the wormlike micelles with increasing the surfactant concentration. The length of alkyl tails was found to strongly affect the viscoelasticity of wormlike micellar solutions. From C10?2C10, C12?2C12 to C14?2C14 in turn, the system developed rapidly from the viscous fluid to typically viscoelastic solution and then to a solid-like gel. The scaling exponents of the concentration dependence of both zero-shear viscosity (η 0) and plateau elastic modulus (G) greatly exceeded the theoretic predictions, showing fast micellar growth and strong entanglements between the wormlike micelles. For C14?2C14 that had the longest alkyl tails in this series, the wormlike micelles formed at 140?mmol L?1 were quite long and the micellar reptation dominated over the scission and recombination. This system yielded a viscosity as high as 2.20?×?104 Pa?s at 25 °C.  相似文献   

9.
Cationic micelles of alkyltrimethylammonium chloride and bromide (alkyl = n? C12H25, n? C14H29, and n? C16H33) catalyze and anionic micelles of sodium dodecyl sulfate inhibit the reaction of hydroxide ion with 2-phenoxyquinoxaline (1). Inert anions such as chloride, nitrate, mesylate, and n-butanosulfonate inhibit the reaction in CTABr by competing with OH? at the micellar surface. The overall micellar effects on rate in cationic micelles and dilute electrolyte can be treated quantitatively in terms of the pseudo-phase ion-exchange model. The determined second-order rate constants in the micellar pseudo-phase are smaller than the second-order constants in water. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
The ternary phase diagrams of zwitterionic single-chain hydrocarbon surfactant (tetradecyldimethylaminoxide, C14DMAO)—a perfluoro cosurfactant (1,1-H-dihydroperfluorooctanol, C7F15CH2OH)-H2O, and C14DMAO-C7F15CH2OH-H2O-HCl have been studied at 25°C. The identification of the phases was done by means of electrical conductivity, optical polarizing microscopy, and 2H-NMR techniques. In this system the originally uncharged zwitterionic surfactant was increasingly charged by protonation through addition of HCl. The sequence of the phases observed is similar to that observed for hydrocarbon surfactant–cosurfactant mixtures, namely, L1, L1/Lα, Lα and L2, as relative volume fraction of apolar compound increases over that of polar compound. The absence of sponge phase (L3) is a result of the high bending constant of the mixed bilayers in this perfluoro system.  相似文献   

11.
Alkyl 2-acetamido-2-deoxy-β-d-glucopyranosides were incorporated into the phospholipid bilayer of small unilamellar vesicles. Depending on the lengths of the alkyl aglycone group (C8–C14), the carbohydrate was either found in the form of micelles (C8) or was effectively incorporated in the bilayer (C14). The use of radiolabelled lipids and/or carbohydrates enabled the quantitative determination of the amounts of lipid and carbohydrate recovered into the vesicles. An enzymatic assay has been carried out to evaluate the outer and inner distribution of the carbohydrate moieties on the liposomes.  相似文献   

12.
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   

13.
The molal formation quotients for cadmium–malonate complexes were measured potentiometrically from 5 to 75°C, at ionic strengths of 0.1, 0.3, 0.6 and 1.0 molal in aqueous sodium trifluoromethanesulfonate (NaTr) media. In addition, the stepwise dissociation quotients for malonic acid were measured in the same medium from 5 to 100°C, at ionic strengths of 0.1, 0.3, 0.6, and 1.0 molal by the same method. The dissociation quotients for malonic acid were modeled as a function of temperature and ionic strength with empirical equations formulated such that the equilibrium constants at infinite dilution were consistent, within the error estimates, with the malonic acid dissociation constants obtained in NaCl media. The equilibrium constants calculated for the dissociation of malonic acid at 25°C and infinite dilution are log K 1a=-2.86 ± 0.01 and log K 2a=-5.71 ± 0.01. A single Cd–malonate species, CdCH2C2O4, was identified from the complexation study and the formation quotients for this species were also modeled as a function of temperature and ionic strength. Thermodynamic parameters obtained by differentiating the equation with respect to temperature for the formation of CdCH2C2O4 at 25°C and infinite dilution are: K = 3.45 ± 0.09, S° = 7 ± 6 kJ-mol-1, S° = 91 ± 22 J-K--mol-1, and C p o =400±300 J­K-1­mol-1.  相似文献   

14.
New [(N?,N,N?)ZrR2] dialkyl complexes (N?,N,N?=pyrrolyl‐pyridyl‐amido or indolyl‐pyridyl‐amido; R=Me or CH2Ph) have been synthesised and tested as pre‐catalysts for ethene and propene polymerisation in combination with different activators, such as B(C6F5)3, [Ph3C][B(C6F5)4], [HNMe2Ph][B(C6F5)4] or solid AlMe3‐depleted methylaluminoxane (DMAO). Polyethylene (Mw>2 MDa and Mw/Mn = 1.3–1.6) has been produced if pre‐catalysts were activated with 1000 equivalents of DMAO (based on Al) [activity >1000 kgPE (mol[Zr] h mol atm)?1] or by using a higher pre‐catalyst concentration and a mixture of [HNPhMe2][B(C6F5)4] (1 equiv) and AliBu2H (60 equiv). In the case of propene polymerisation, activity has been observed only if pre‐catalysts were treated with an excess of AliBu2H prior to addition of DMAO, which led to highly isotactic polypropylene ([mmmm]>95 %). Neutral pre‐catalysts and ion pairs derived from their activation have been characterised in solution by using advanced 1D and 2D NMR spectroscopy experiments. The detection and rationalisation of intercationic NOEs clearly showed the formation of dimeric species in which some pyrrolyl or indolyl π‐electron density of one unit is engaged in stabilising the metal centre of the other unit, which relegates the counterions in the second coordination sphere. The solid‐state structure of the dimeric indolyl‐pyridyl‐amidomethylzirconium derivative, determined by X‐ray diffraction studies, points toward a weak Zr???η3‐indolyl interaction. It can be hypothesised that the formation of dimeric cationic species hampers monomer coordination (especially of less reactive α‐olefins) and that addition of AliBu2H is crucial to split the homodimers.  相似文献   

15.
1-Alkyl-3-(2-oximinopropyl)imidazolium chlorides were prepared with different alkyl chain lengths (Alk = C12H25, C14H29). Some physicochemical indices (CMC and pK aapp) were determined. The reactivity of these compounds was studied in the dissociation of 4-nitrophenyl esters of diethylphosphonic, diethylphosphoric, and toluenesulfonic acids. The times for 50% conversion of the substrates into reaction products decrease in the series: C12H25 > C14H29 >C16H33. In selecting the direction of modification of the supernucleophilic functional surfactants, we should take into account not only their hydrophobic properties but also the efficiency of substrate solubilization as well as the reactivity of the oximate group in the surfactant micelles.  相似文献   

16.
Salt effects on the aggregation behavior of tripolar zwitterionic surfactants in aqueous solutions have been investigated using surface tension, dynamic light scattering (DLS), freeze-fracture transmission electron microscopy (FF-TEM), and 1H NMR. The tripolar zwitterionic surfactants with different inter-charge spacers are [C14H29(CH3)2N+CsN+(CH3)2CH2CH2CH2SO3 ?]Br? (C14CsTri, Cs?=?–(CH2)2–, –(CH2)6–, –(CH2)10–, and p-xylyl). It is found that the critical micelle concentration (CMC) values of the corresponding traditional zwitterionic surfactant C14H29(CH3)2N+CH2CH2CH2SO3 ? (TPS) are almost constant with the increase of the NaBr concentration. However, the CMC values of C14CsTri decrease sharply at a lower NaBr concentration and then level off at a higher NaBr concentration. Moreover, the decreasing extents of the CMC values for C14C2Tri, C14C6Tri, and C14CpxTri are very close, but more significant than that for C14C10Tri, suggesting that the self-assembly ability of the tripolar zwitterionic surfactants with a longer inter-charge spacer is less sensitive to NaBr. The DLS and FF-TEM results reveal that C14C2Tri, C14C6Tri, and C14CpxTri form micelles without NaBr and that the size slightly increases with the increase of NaBr concentration, whereas micelles and vesicles coexist for C14C10Tri and TPS without NaBr and then transfer to micelles upon the addition of NaBr. The salt-induced morphological transition for C14C10Tri is further studied using 1H NMR. The addition of NaBr reduces both the electrostatic repulsion between the same charged ammoniums and the electrostatic attraction between the oppositely charged ammonium and sulfonate. Thus, the longer inter-charge spacer of C14C10Tri tends to be more bended and the sulfonate group becomes available to contact the ammonium, which promotes micellization.  相似文献   

17.
The interactions of bovine serum albumin (BSA) with two cationic gemini surfactants, (C n N)2Cl2 (n = 12, 14), in buffer solutions (pH = 7.0) were investigated by isothermal titration calorimetry (ITC) and circular dichroism (CD). CD spectra showed that the two surfactants change the secondary structure of BSA. The thermodynamic results suggest that there exist two binding types (high affinity/low affinity) in the interacting process of (C n N)2Cl2 micelles with BSA. The high affinity binding is an endothermic process driven by entropy, in which the synergistic effect among weak interactions plays an important role. The low affinity binding is an exothermic process accompanied by positive entropy effect, in which hydrophobic interaction is dominant in all driving forces. Furthermore, corresponding binding site number of BSA for (C14N)2Cl2 is much smaller than that for (C12N)2Cl2, indicating that the hydrophobic chain length of surfactant plays a key role in low affinity binding process.  相似文献   

18.
The complex formation between nonionic alkyldimethylamine oxide (CnDMAO, n=14, 16, and 18) and sodium palmitate (NaPa) in the solid phase of CnDMAO/NaPa mixtures and the dependence of the interaction parameter beta of the regular solution theory (RST) on the mixed micelle composition of C16DMAO/NaPa mixtures were investigated. The dissolution temperature showed a maximum at a NaPa mole fraction X(Pa)(*) of 0.3-0.4 for C16DMAO/NaPa and 0.2 for C18DMAO/NaPa. The compositions of the complexes suggested by X(Pa)(*) are C16DMAO: NaPa=3:2 or 2:1 and C18DMAO:NaPa=4:1. The composition X(Pa)(*) depended on the chain length of the amine oxides. The maximum was not observed in the case of the C14DMAO/NaPa/water system. In the range 0.7< or =X(Pa)< or =1.0, dissolution temperature depression was observed with decreasing X(Pa). The dissolution temperature depression was analyzed by taking into account the nonideal behavior in the mixed micelles and the counterion binding on the mixed micelle surface. The negative beta values were obtained for all three mixed systems. It was shown that the counterion activity remained practically constant in the range of 0.7< or =X(Pa)< or =1.0. The cmc values of C16DMAO/NaPa mixtures were determined by pyrene fluorescence measurement. For C16DMAO/NaPa mixtures, the dependence of the RST interaction parameter beta on the mixed micelle composition X(Pa) was determined for a wide range (0.2< or =X(Pa) < or =0.9). In the range 0.2< or =X(Pa)< or =0.5, the beta values were obtained from an analysis of cmc based on the RST. In the range 0.7< or=X(Pa)< or=0.9, the beta values were obtained from an analysis of the dissolution temperature depression. From the analysis of the micelle composition dependence of the beta values, a short-range attractive interaction between the headgroup of C16DMAO and palmitate anion is suggested.  相似文献   

19.
An unnatural amino acid, β-[6′-(N, N-dimethyl)amino-2′-naphthoyl]alanine (Ald) showing polarity-sen sitive fluorescence characteristics, was synthesized. A thorough Ald-scan of dynorphin A (Dyn A), the putative endogenous ligand for κ opioid receptors, was then performed. Replacement of the amino acid residues in positions 5, 8, 10, 12 or 14 of Dyn A(1-13)-NH2 with Ald resulted in compounds that had almost equal κ binding affinity compared with that of the parent compound; on the other hand, substi-tution o...  相似文献   

20.
The interactions of copper complex (CuQ2) of 1-amino-4-hydroxy-9,10-anthraquinone (QH) with calf thymus DNA, anionic surfactant sodium dodecyl sulfate (SDS), and cationic surfactant cetyltrimethylammonium bromide (CTAB) were investigated in an aqueous solution at physiological pH (7.4). Affinities of such molecule to DNA and surfactant micelles, a model for a biological membrane, are important in determining its biological action. Using different models, various binding parameters were evaluated in both of molecule–DNA interaction and molecule–surfactant interaction. The study showed that hydrophobic interaction plays a major role in the binding of CuQ2 to surfactant micelles. In addition, the hydrophobic interaction has an important role in the distribution of CuQ2 between micelle-water phases. Gibbs free energy for the binding and distribution of CuQ2 between the bulk aqueous medium and surfactant micelles were calculated. In order to correlate the physicochemical properties deciphered from the aforementioned studies with the biological property of the molecule, CuQ2 was treated with MDA-MB-231 breast adenocarcinoma cells where it was found that the molecule affects the viability of the cancer cells. Fluorescent staining of the treated cells with AO/EB and Hoechst indicated that the CuQ2 induces apoptosis, suggesting its use in the treatment of breast cancer.  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