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 93 毫秒
1.
We review several key elements of alkyl polyglucoside (CmGn) microemulsion phase behavior. The low solubility of CmGn surfactants in oils such as alkanes makes producing CmGn microemulsions and subsequent study of their phase behavior difficult. Increasing the solubility of CmGn in oil is therefore helpful for the systematic study of CmGn-based microemulsion formulations. To this end, the role of cosurfactants in producing microemulsions with water, alkanes, and n-alkyl β-d-glucopyranosides is first discussed. Adding C10βG1 to mixtures of water–alkane–ethoxylated alcohol surfactants (CiEj) produces a region of the three-phase body (a ‘chimney’) that is independent of temperature; thus CmβG1 are not completely soluble in the co-oil formed of alkane and CiEj at higher temperatures. Then, through a novel approach using oxygenated ether oils (CkOC2OCk), microemulsions are formed with water, CkOC2OCk, and CmβG1 and the phase behavior studied as a function of temperature and composition. Increased CmβG1 solubility in the more hydrophilic ether oils produces patterns of phase behavior in water–CkOC2OCk–CmβG1 mixtures that are identical to those observed in water–alkane–CiEj mixtures. Using the water–ether oil–CmβG1 mixtures as a base case, the role of CmGn surfactant structure in setting CmGn microemulsion phase behavior is explored. The solubility of the α-d anomer (n-alkyl α-d-glucopyranosides, CmαG1) in water is much less than that of the β-d surfactant, and these solubility boundaries extend to high surfactant and oil concentrations in water–CkOC2OCk–CmαG1 mixtures. Adding CmG2 compounds to water–CkOC2OCk–CmβG1 mixtures shifts the phase behavior to high temperatures, again demonstrating the extreme hydrophilic nature of the sugar headgroup. Finally, adding small amounts of ionic alkyl sulfate surfactants to water–CkOC2OCk–CmβG1 mixtures dramatically reduces the total amount of surfactant needed to form a single-phase microemulsion.  相似文献   

2.
The influence of newly synthesized N-alkanoyl-N-methyllactitolamines (decanoyl [C10MELA], lauoroyl [C12MELA] and miristoyl [C14MELA]) on the thermotropic phase transition of phosphatidylcholine bilayer was compared with common sugar-based surfactants N-dodecyl-β-d-glucopyranoside [C12G1] and decanoyl-N-methyl glucamide [MEGA-10]. The results indicate that CnMELA are very active at the membrane surface and disturb the phospholipid bilayer structure less than commercially used MEGA-10 and C12G1.  相似文献   

3.
The ionic liquid 1-hexadecyl-3-methylimidazolium bromide ([C16MIm]Br) has been used as a novel cationic surfactant for separation of phenolic compounds, including quinol, phloroglucinol, resorcinol, phenol, p-cresol, and m-nitrophenol, by micellar electrokinetic capillary chromatography (MEKC). The effects of buffer concentration and pH, concentration of [C16MIm]Br, and applied potential were studied. Use of the optimized buffer (25 mmol L?1 NaH2PO4), 10 mmol L?1 [C16MIm]Br, and an applied potential of ?15 kV enables optimum separation with regard to resolution and migration time. The phenolic compounds were detected at 214 nm. The micelle of this long-alkyl-chain imidazolium ionic liquid acts as a pseudo-stationary phase in this MEKC separation.  相似文献   

4.
Thermodynamics on inclusion complexation of β-cyclodextrin (β-CD) with n-alkylpyridinium chlorides (C n PC, n = 12, 14, 16) were measured by conductivity technique to evaluate the effects of chain length of C n PC and temperature. The data obtained indicate that inclusion complexes S(CD) and S(CD)2 had formed between surfactant and β-CD in aqueous solution. Investigation showed that the K 1 (first equilibrium constant) for S(CD) formation is greater than K 2 (second equilibrium constant) for S(CD)2 formation. It has been found that C12PC forms only the 1:1 complex, while C14PC and C16PC form 1:1 and 1:2 complexes. Thermodynamic parameters of the complexation, i.e. ΔG°, ΔH° and ΔS° have been also calculated. The large values of ΔG° indicate that complexation between surfactant and β-CD is very favorable.  相似文献   

5.
Solutions of n-nonyl-β-D-glucoside (C9G1), n-decyl-β-D-glucoside (C10G1), n-dodecyl-β-D-maltoside (C12G2), n-tetradecyl-β-D-maltoside (C14G2) and C9G1/C10G1 mixtures have been characterised by capillary viscometry and rheology in H2O and D2O, in order to map the influence of surfactant characteristics on micellisation over a wide concentration range. For the maltosides, the micellar solutions are shear thinning with a zero-shear viscosity that scales with concentration according to a power law with an exponent of about 5.8. In contrast, solutions of the glucosides C9G1, C10G1 and their mixtures show Newtonian flow behaviour and a much lower scaling exponent (<2.4). In C9G1/C10G1 mixtures, the scaling exponent decreases monotonously with increasing C10G1 content. The flow behaviour correlates with the packing requirements of the various surfactants, and are compatible with the idea that the maltosides form worm-like micelles, whereas the glucosides form branched, interconnected micelles (C9G1) and space-filling micellar networks (C10G1).  相似文献   

6.
A new kind of surfactant, [CnH_(2n+1)OCH2CH(OH)CH2N(CH3)3]Cl (n=12, 14, 16) was synthesized. The solubility of benzyl alcohol in micellar solutions was determined by 1H NMR method. The results indicate that the length of alkyl chains of surfactant affects the solubility of ben-zyl alcohol in 2.5 × l0~(-2) mol/L micellar solutions. The solubility of benzyl alcohol per liter of micellar solution is 0.095 mole for n=12, 0.115 mole for n=14, 0.165 mole for n=16. The transfer free energy of benzyl alcohol from aqueous phase to micellar phase is -24.29 kJ/mol for n=12, -24.37 kJ/mol for n=14, -24.49 kJ/mol for n=16.  相似文献   

7.
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   

8.
The effects of the variables of head group structure and salt concentration on microemulsions formed in mixtures of water, alkyl ethylene glycol ethers (CkOC2OCk), andn-alkyl β- -glucopyranosides (CmβG1) are explored. Phase behavior of mixtures containing an anomer of the surfactant (n-alkyl α- -glucopyranoside, CmαG1), or surfactants with long head groups (n-alkyl maltopyranosides, CmG2), or NaCl or NaClO4as electrolyte are systematically reported as a function of temperature and composition. The substitution ofn-alkyl α- -glucopyranosides forn-alkyl β- -glucopyranosides causes precipitation under some conditions in all mixtures studied. These solubility boundaries begin in the water–surfactant binary mixture at the Krafft boundary, then extend to high concentrations of both surfactant and oil. Increasing the effective length of the surfactant head group by adding CmG2to water–CkOC2OCk–CmβG1mixtures moves the phase behavior dramatically up in temperature when even small amounts of CmG2are used. Adding a lyotropic electrolyte, NaCl, to water–CkOC2OCk–CmβG1mixtures moves the phase behavior down in temperature, while the hydrotropic electrolyte NaClO4moves the phase behavior up in temperature.  相似文献   

9.
The solubilization of four chalcones, between aqueous and micellar phases of ionic surfactants (SDS and CTAB), was investigated by conductivity and cyclic voltammetry (CV) techniques. From conductivity data, a decrease in the critical micellar concentration (CMC) of the surfactants, in presence of the chalcones was ascribed to the decreased charge density over the surfactants. The results were seconded by thermodynamic parameters including degree of ionization (α), counter ion binding (β), and standard Gibbs free energy of micellization (ΔG m ). The added surfactant decreased the peak current of the oxidized chalcone and shifted the peak potential either positively (in presence of SDS) or negatively (in presence of CTAB). The effect is rationalized as chalcone-surfactant interaction and quantitated as binding constant (K b) assorting values from 8.78 to 552.97 M?1. The preferred solubilization of the chalcones in the micellar phase has been inferred.  相似文献   

10.
The interactions between an anionic surfactant, viz., sodium dodecylbenzenesulfonate and nonionic surfactants with different secondary ethoxylated chain length, viz., Tergitol 15-S-12, Tergitol 15-S-9, and Tergitol 15-S-7 have been studied in the present article. An attempt has also been made to investigate the effect of ethoxylated chain length on the micellar and the thermodynamic properties of the mixed surfactant systems. The micellar properties like critical micelle concentration (CMC), micellar composition (XA), interaction parameter (β), and the activity coefficients (fA and fNI) have been evaluated using Rubingh's regular solution theory. In addition to micellar studies, thermodynamic parameters like the surface pressure (ΠCMC), surface excess values (ΓCMC), average area of the monomers at the air–water interface (Aavg), free energy of micellization (ΔGm), minimum energy at the air–water interface (Gmin), etc., have also been calculated. It has been found that in mixtures of anionic and nonionic secondary ethoxylated surfactants, a surfactant containing a smaller ethoxylated chain is favored thermodynamically. Additionally, the adsorption of nonionic species on air/water interface and micelle increases with decreasing secondary ethoxylated chain length. Dynamic light scattering and viscometric studies have also been performed to study the interactions between anionic and nonionic surfactants used.  相似文献   

11.
This paper describes a sweeping–micellar electrokinetic chromatography (sweeping–MEKC) technique for the determination of seven benzodiazepines, using, as sweeping carriers, the ionic liquid-type cationic surfactants 1-cetyl-3-methylimidazolium bromide (C16MIMBr) and N-cetyl-N-methylpyrrolidinium bromide (C16MPYB). These surfactants resemble the commonly employed cationic surfactant cetyltrimethylammonium bromide (CTAB), but they provide different separation efficiencies. We optimized the separation and sweeping conditions, including the pH, the concentrations of organic modifier and surfactant, and the sample injection volume. Adding C16MIMBr or C16MPYB to the background electrolyte enhanced the separation efficiency and detection sensitivity during the sweeping–MEKC analyses of the benzodiazepines. C16MIMBr enhanced the sensitivity for each benzodiazepine 31–59-fold; C16MPYB, 86–165-fold. In the presence of C16MPYB, the limits of detection for the seven analytes ranged from 4.68 to 9.75 ng/mL. We adopted the sweeping–MEKC conditions optimized for C16MPYB to satisfactorily analyze a human urine sample spiked with the seven benzodiazepines. To minimize the matrix effects, we subjected this urine sample to off-line solid phase extraction (SPE) prior to analysis. The recoveries of the analytes after SPE were satisfactory (ca. 77.0–88.3%). Our experimental results reveal that the cationic surfactant C16MPYB exhibits superior sweeping power relative to those of C16MIMBr and CTAB and that it can be applied in sweeping–MEKC analyses for the on-line concentrating and analyzing of benzodiazepines present in real samples at nanogram-per-milliliter concentrations.  相似文献   

12.
In the preceding paper of this series, we studied the interactions of copolymers with the ionic liquids, 1-alkyl-3-methylimidazolium bromide (C n mimBr, n?=?8, 10, 12, 14, 16) and N-alkyl-N-methylpyrrolidinium bromide (C n MPB, n?=?12, 14, 16). An obvious difference was detected between the interaction mechanism and the alkyl chain length of the surfactant. In the present study, we performed a systematic study on the interaction of sodium carboxymethylcellulose (NaCMC) with ionic liquids in aqueous solution by isothermal titration microcalorimetry (ITC), conductivity, turbidity, and dynamic light scattering (DLS) measurements. The existence of electrostatic attraction between NaCMC and ILs could increase the complexity of these systems. The results show that the monomers of C8mimBr can bind to the NaCMC chains and form free surfactant micelles in the solution, while no micelle-like C8mimBr/NaCMC cluster is detected. For other surfactants, the formation of surfactant/NaCMC clusters in the solution is driven by electrostatic and hydrophobic interactions, which could be divided into two types. One type is the polymer-induced surfactant/NaCMC complexes that form in the solution for the surfactant of C n mimBr (n?=?10, 12, 14) or C n MPB (n?=?12, 14). The other type is that the surfactant-induced surfactant/NaCMC complexes come into being for the surfactant of C16mimBr or C16MPB. Finally, the different modes of complex formation proposed have a good interpretation of the experiment results, unraveling the details of the effect of surfactant alkyl chain length and headgroup on the surfactant–NaCMC interactions.  相似文献   

13.
Microemulsions provide a unique opportunity to tailor the polarity and liquid confinement in asymmetric catalysis via nanoscale polar and nonpolar domains separated by a surfactant film. For chiral diene Rh complexes, the influence of counterion and surfactant film on the catalytic activity and enantioselectivity remained elusive. To explore this issue chiral norbornadiene Rh(X) complexes (X=OTf, OTs, OAc, PO2F2) were synthesized and characterized by X-ray crystallography and theoretical calculations. These complexes were used in Rh-catalyzed 1,2-additions of phenylboroxine to N-tosylimine in microemulsions stabilized either exclusively by n-octyl-β-D-glucopyranoside (C8G1) or a C8G1-film doped with anionic or cationic surfactants (AOT, SDS and DTAB). The Rh(OAc) complex showed the largest dependence on the composition of the microemulsion, yielding up to 59 % (90 %ee) for the surfactant film doped with 5 wt% of AOT as compared to 52 % (58 %ee) for neat C8G1 at constant surfactant concentration. Larger domains, determined by SAXS analysis, enabled further increase in yield and selectivity while the reaction rate almost remained constant according to kinetic studies.  相似文献   

14.
A group of novel fluorescent surfactants, N-n-alkyl-4-(1-methylpiperazine)-1,8-naphthalimide iodine [Cnndi]I (n?=?8, 10, and 12), have been synthesized and their aggregation behavior in aqueous solution have been explored by surface tension, electric conductivity, hydrogen-1 NMR spectra, absorption, and fluorescence spectra. Compared with traditional cationic surfactants, the [Cnndi]I have a rather lower critical micelle concentration and higher surface activity. Absorption and fluorescence spectra were proved to be facile method to monitor directly the aggregation states of fluorescent surfactant molecules in solution and revealed clearly the formation of face-to-face stacked structure of the [Cnndi]I molecules driven by the π–π interactions. The micelle formation process for [Cnndi]I was demonstrated to be enthalpy-driven in the temperature range investigated. Possible aggregation process was given based on the experimental results. The combination of dye and surfactant provides a way for monitoring the formation process of micelle directly by fluorescence spectra.  相似文献   

15.
Novel organometallic surfactants were synthesized from monoquaternized 1,4-diazabicyclo[2.2.2]-octanes of varied hydrophobicity [Alk = C n H2n+1, n = 14, 18] and lanthanum nitrates. The spectral, micelleforming, and adsorption properties of the synthesized compounds were studied by IR, 1Н NMR and electronic spectroscopy, tensiometry, conductometry, and potentiometry. The critical micelle concentrations, counterionic micellar binding constants, and adsorption parameters at the water–air interface were determined and compared with the respective characteristics of the ligands and conventional amphiphiles.  相似文献   

16.
The phase behaviors of four phytosterol ethoxylates surfactants (BPS-n, n = 5, 10, 20, and 30) with different oxyethylene units in room temperature ionic liquid, 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim]BF4), have been studied. The polarized optical microscopy and small-angle X-ray scattering techniques are used to characterize the phase structures of these binary systems at 25 °C. The structure and ordering of the liquid crystalline (LC) phases in such BPS-n/[Bmim]BF4 systems are found to be influenced by BPS-n concentration and the temperature. Due to the bulky and rigid cholesterol group, the phytosterol ethoxylates surfactants exhibit different properties and interaction mechanism from the conventional CnEOm type nonionic surfactant systems. The rheological measurements indicate a highly viscoelastic nature of these lyotropic LC phases and disclose a lamellar phase characteristic with a rather strong rigidity at high surfactant concentrations. The control experiment with Brij 97(polyoxyethylene (10) oleyl ether)/[Bmim]BF4 system and the FTIR measurements help to recognize that the solvophobic interaction combining with the hydrogen bonding are the main driving forces for the LC phases formation.  相似文献   

17.
The diffusion-controlled adsorption kinetics of micellar surfactant C12E7 (heptaethylene glycol monododecyl ether) solutions was studied theoretically and experimentally. The corrected diffusion equation, which was used to describe the diffusion of the monomers in the micellar solutions, was solved under the initial and boundary conditions by means of Laplace transformation. The dynamic surface adsorption γ(t) as a function of surface lifetime t, monomer diffusion coefficient D and the demicellization constant was derived. The dynamic surface tensions γ(t) of aqueous submicellar and micellar solutions were measured via maximal bubble pressure method. By analyzing the experimental data, the determined demicellization constant of C12E7 at 25°C was between 100–116 s?1.  相似文献   

18.
The interaction between β-cyclodextrin (β-CD) and an amino acid-based anionic gemini surfactant derived from cysteine (C8Cys)2 was studied by three independent techniques: electrical conductivity, UV–Vis spectral displacement technique using phenolphthalein as probe, and 1H NMR spectroscopy. The data obtained indicated the formation of a 1:1 inclusion complex between β-CD and the gemini surfactant studied and allowed for the determination of the binding constant, K1, by considering this stoichiometry. Electrical conductivity, spectral displacement technique, and NMR chemical shift measurements, obtained for aqueous β-CD–surfactant systems, yielded consistent K1 values in the order of 102 dm3 mol?1, typical of a weakly bound β-CD–surfactant complex. The influence of the presence of the inclusion complex on the micellization process of the gemini surfactant has also been studied and the apparent critical micelle concentration (cmc1) has been obtained. Increasing β-CD concentration was found to shift the cmc1 to higher values, as complexed surfactant monomers are not available to form micelles and aggregation takes place only when all β-CD cavities are occupied.  相似文献   

19.
An easy preparation of mono-deprotected thioglucopyranosides via a selective Candica cylindracea lipase-catalyzed hydrolysis of a commercially available peracetylated precursor is described. Especially, ethyl 2,3,4-tri-O-acetyl-1-thio-β-d-glucopyranoside and ethyl 2,3,6-tri-O-acetyl-1-thio-β-d-glucopyranoside were obtained in 100% and 54% isolated yields, respectively. The influence of the ratio of [bmim]PF6/buffer toward the regioselectivity of the deacetylation step and the acyl migration is discussed.  相似文献   

20.
The phase behavior in water of pentaglycerol monostearate (C18G5) and pentaglycerol monooleate (C18:1G5) surfactants has been studied as a function of temperature and surfactant weight fraction, W s . The equilibrium phases present at each composition and temperature studied were characterized by means of visual observation under normal and polarized light, differential scanning calorimetry (DSC), and X‐ray scattering, both at small (SAXS) and at wide angle (WAXS). In the temperature range 0–46°C, C18G5 presents a thermotropic α‐gel structure. However, at higher temperatures, the α‐gel phase melts and a lamellar liquid crystalline (Lα) phase is formed. The amount of water that can be solubilized by α‐gel and Lα was determined by plotting the interlayer distance, d, as a function of the reciprocal of W s . Water is soluble in the α‐gel phase up to 21 w/w% water concentration and in the Lα phase up to 30 w/w% water concentration. At higher water concentrations, excess water appears and a dispersion of α‐gel (α‐gel+W) and lamellar liquid crystal (Lα+W) in water is formed, respectively. In contrast, C18:1G5 is liquid in the whole range of temperatures studied (0–100°C). While at low temperatures, C18:1G5 presents a Lα structure, at about 63°C Lα melts and an isotropic liquid reverse micellar solution (Om) phase is formed. The amount of water that can be solubilized by both Om and Lα increases with temperature.  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