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 46 毫秒
1.
The semiequilibrium dialysis method has been used to determine solubilization equilibrium constants and activity coefficients of benzoic, phenylacetic, and hydrocinnamic acids (solubilizate) in micelles of the cationic surfactant hexadecylpyridinum chloride (cetylpyridinium chloride) in 0.1M HCl aqueous solutions. Methods described previously were employed to infer the concentrations of monomeric organic solute and surfactant on both sides of the dialysis cell. Values of the apparent solubilization constant K of the neutral acids have been correlated with mole fractions of the acid in the micelle XA, where K=XA/[monomeric acid]. The activity coefficients of both acid and surfactant were obtained, consistent with the Gibbs-Duhem equation. The solubilization constants of all three acids are nearly the same, indicating that there is no significant effect owing to the presence of one or more methylene groups between the carboxylate and the phenyl groups of benzoic acid. The solubilization constants also decrease appreciably, and the activity coefficients of the acids increase, as the mole fraction of the acid in the micelle increases.  相似文献   

2.
The semi-equilibrium dialysis method has been used to infer solubilization equilibrium constants or, alternatively, activity coefficients of solutes solubilized into micelles of aqueous surfactant solutions. Methods are described for inferring the concentrationa of monomers of the organic solute and of the surfactant on both sides of the dialysis membrane, under conditions where the organic solute is in equilibrium with both the high-concentration (retentate) and low-concentration (permeate) solutions. By using a form of the Gibbs-Duhem equation, activity coefficients of both phenol (the solubilizate) and n-hexadecylpyridinium chloride (the surfactant) are obtained for aqueous solutions at 25°C throughout a wide range of relative compositions of surfactant and solubilizate within the micelle. The apparent solubilization constant, K=[solubilized phenol]/([monomeric phenol][micellar surfactant]), is found to decrease significantly as the mole fraction of phenol in the micelle increases.  相似文献   

3.
Aqueous solutions of -cyclodextrin (-CD) or 2,6-di-o-methyl--cyclodextrin (DM--CD) and dodecylethyldimethylammonium bromide (D12EDMAB) have been studied from speed of sound (u) data at 298.15 K, using a pulse-echo-overlap technique. The molecular encapsulation process of the surfactant monomer into the cyclodextrin cavity and its effect in the micellization process of the surfactant have been analyzed from theu measurements: I) as a function of [D12EDMAB] in the presence of several initial cyclodextrin concentrations (-CD or.DM--CD); II) as a function of [cyclodextrin] (-CD or DM--CD), for an initial micellar solution of D12EDMAB and; III) as a function of the [cyclodextrin]/[surfactant] stoichiometric concentrations. Both inclusion complexes formed (-CDD12EDMAB) and (DM--CDD12EDMAB) have stoichiometries of 11, and their association constantK have been determined using a model proposed in this work, based on the additivity of the different contributions of the involved species to the speed of sound. The apparent critical micellar concentration, cmc*, of D12EDMAB is found to increase linearly upon the addition of cyclodextrin (-CD or DM--CD). The free surfactant concentration in the micellar region, [D12EDMAB]f, decreases in the presence of -CD and slightly increases in the presence of DM--CD. The influence of the parcial methylation of the -cyclodextrin (-CDDM--CD) and of the polar head of the surfactant (D12TAB D12EDMAB) on the complextion and micellar parameters are also discussed.Supplementary material available: Tables of speed of sound (14 pages) are available from the authors.  相似文献   

4.
The mechanisms governing the subsolubilizing and solubilizing interaction of sodium dodecyl sulphate (SDS)/Triton X-100 mixtures and phosphatidylcholine liposomes were investigated. Permeability alterations were detected as a change in 5(6)-carboxy-fluorescein (CF) released from the interior of vesicles and bilayer solubilization as a decrease in the static light-scattered by liposome suspensions. Three parameters were described as the effective surfactant/lipid molar ratios (Re) at which the surfactant system a) resulted in 50% of CF release (Re 50%CF); b) saturated the liposomes (Re SAT;c) led to a complete solubilization of these structures (Re SOL). From these parameters the corresponding surfactant partition coefficientsK 50%CF,K SAT andK SOL were determined. The free surfactant concentrationsS W were lower than the mixed surfactant CMCs at subsolubilizing level, whereas they remained similar to these values during saturation and solubilization of bilayers in all cases. Although theRe increased as the mole fraction of the SDS rose (X SDS), theK parameters showed a maximum atX SDS values of about 0.6, 0.4 and 0.2 forK 50%CF,K SAT andK SOL respectively. Thus, the higher the surfactant contribution in surfactant/lipid system, the lower theX SDS at which a maximum bilayer/water partitioning of mixed surfactant systems added took place and, consequently, the lower the influence of the SDS in this maximum bilayer/water partitioning.Abbreviations PC Phosphatidylcholine - PIPES piperazine-1,4 bis (2-ethanesulphonic acid) - SDS sodium dodecyl sulphate - X SDS mole fraction of sodium dodecyl sulphate in the mixed system - CF 5(6)-carboxyfluorescein - Re effective surfactant/lipid molar ratio - Re 50%CF effective surfactant/lipid molar ratio for 50% CF release - Re SAT effective surfactant/lipid molar ratio for bilayer saturation - Re SOL effective surfactant/lipid molar ratio for bilayer solubilization - S W surfactant concentration in the aqueous medium - S W, 50%CF surfactant concentration in the aqueous medium for 50% CF release - S W, SAT surfactant concentration in the aqueous medium for bilayer saturation - S W, SOL surfactant concentration in the aqueous medium for bilayer solubilization - S B surfactant concentration in the bilayers - K bilayer/aqueous phase surfactant partition coefficient - K 50%CF bilayer/aqueous phase surfactant partition coefficient for 50% CF release - K SAT bilayer/aqueous phase surfactant partition coefficient for bilayer saturation - K SOL bilayer/aqueous phase surfactant partition coefficient for bilayer solubilization - PL phospholipid TLC-FID, thin-layer chromatography/flame ionization detection system - PI polydispersity index - CMC critical micellar concentration - r 2 regression coefficient  相似文献   

5.
The kinetics of formation of 11 complexes of nickel(II) and copper(II) ions with some azophenol derivatives in aqueous and micellar solution of a nonionic surfactant, Triton X-100, have been studied by a stopped-flow spectrophotometric method. Second order rate constants for the reactions were determined at 298 K and ionic strength 0.1 (NaClO4) in aqueous solution. In the surfactant solution, the pseudo-first-order rate constants for the complexation reactions,kobs, decreased with increasing the concentration of Triton X-100. This observation was explained by the assumption that the chelating reagents distribute between the micelle of the surfactant and bulk aqueous phase and rate-controlling reactions occur only in the bulk aqueous phase. On the basis of the relation betweenkobs and the concentration of the surfactant, the partition constants of the reagents between micellar and aqueous phases were determined.  相似文献   

6.
The assembly properties of the nonionic surfactant Triton X-100 and phosphatidylcholine (PC) aggregates during the overall solubilization process of PC liposome were investigated. Permeability alterations were detected as a change in 5(6)-carboxyfluorescein (CF) released from the interior of vesicles and bilayer solubilization as a decrease in the static light scattered by liposome suspensions. A direct dependence was established between the bilayer/aqueous phase surfactant partition coefficients (K), the growth of vesicles and the leakage of entrapped CF in the initial interaction steps (surfactant to phospholipid molar ratioRe up to 0.2). These changes may be related to the increasing presence of surfactant molecules in the outer monolayer of vesicles. In theRe range 0.2–0.35 the coexistence of a low vesicle growth with a constant increase of CF release may be correlated with the decrease inK (increased rate of flip-flop of surfactant molecules). Furthermore, in theRe range between 0.64 and 2.0 (lytic levels) almost a linear dependence was detected between the composition of these aggregates (Re) and the decrease in both the surfactant-PC aggregate size and the static light scattered by the system. This dependence was not observed in the last solubilization steps (Re range 2.0–2.60) possibly due to the increased formation of mixed micelles in this interval. The fact that the free Triton X-100 concentration at sublytic and lytic levels showed respectively lower and similar values than its critical micelle concentration confirms that permeability alterations and solubilization were determined respectively by the action of surfactant monomer and by the formation of mixed micelles.Abbreviations PC phosphatidylcholine - PIPES piperazine-1,4 bis(2-ethanesulphonic acid) - TX-100 Triton X-100 - CF 5(6)-carboxyflucrescein - Re enective surfactant/lipid molar ratio - Re SAT effective surfactant/lipid molar ratio for bilayer saturation - Re SOL enective surfactant/lipid molar ratio for bilayer solubilization - S W surfanctant concentration in the aqueous medium - S B surfactant concentration in the bilayers - S T total surfactant concentration - K bilayer/aqueous phase surfactant partition coefficient - K SAT bilayer/aqneous phase surfactant partition coefficient for bilayer saturation - K SOL bilayer/aqueous phase surfactant partition coefficient for bilayer solubilization - PL phospholipid - TLC-FID thinlayer chromatography/flame ionization detection system - PI polydispersity index - CMC critical micellar concentration - r 2 regression coefficient  相似文献   

7.
Density measurements of water-dodecyltrimethylammonium bromide (DTAB)-alcohol ternary systems as a function of alcohol and surfactant concentrations were carried out at 25°C. The alcohols were propanol (PrOH), 2-propanol (2-PrOH) and hexanol (HexOH). The apparent molar volume V,R of alcohols have been calculated and the standard (infinite dilution) partial molar volumes of alcohols V R at each surfactant concentration were obtained by means of a least squares fit of V,R vs. the alcohol concentration. The V R vs. surfactant concentration curves have been rationalized in terms of the partial molar volume of alcohol in the aqueous V f and the micellar V b phases and the distribution constant of alcohol between the aqueous and the micellar phases K. The V b values for PrOH and HexOH together with those of butanol and pentanol previously reported satisfy the additivity rule giving a methylene group contribution of 16.7 cm3-mol–1 which is identical to that reported in the literature from the study of pure liquid alcohols. No difference between V b for PrOH and 2-PrOH has been found. From density data of water-alcohol and water-surfactant binary systems and of water-surfactant-alcohol ternary system, the apparent molar volume of the surfactant in the water-alcohol mixed solvent V,S have been calculated as a function of the surfactant concentration and of the mixed solvent composition. The effect of the alkyl chain length of the alcohols and the effect of isomerization of the alcohols on the V,S vs. surfactant concentration trends have been analyzed.  相似文献   

8.
Speed of sound and density properties of ternary water-tetradecyltrimethylammonium bromide-pentanol system at 15, 25 and 35°C and of water-hexadecyltrimethylammonium bromide-pentanol system at 25, 35 and 45°C were measured at fixed alcohol concentration as a function of surfactant concentration. The apparent molar volumes V,R and isentropic compressibilities K ,R S of pentanol in micellar solutions as a function of the surfactant concentration show irregular behavior which depends on the alkyl chain length of the surfactant and tends to disappear with increasing temperature. These anomalies are ascribed to micellar transitions. For both surfactants at high concentrations, V,R decrease and the magnitude of the change seems to depend on the type of densimeter used. This observation is tentatively explained in terms of a correlation between the micellar structure and features of the densimeter. From this work and literature data, the apparent molar isothermal compressibilities K ,R T of the alcohol in micellar solutions were calculated at 25°C. V,R , K ,R S and K ,R T are interpreted in terms of the distribution constant of the alcohol between the aqueous and the micellar phases and of the apparent molar property of the alcohol in the micellar and the aqueous phases. For a given surfactant increasing the temperature increases V,R and K ,R S in the micellar phase while the distribution constant is weakly dependent. At a given temperature, an increase in the alkyl chain length of the surfactant increases the apparent molar volume and slightly changes the apparent molar compressibility of the alcohol in the micellar phase.  相似文献   

9.
Solubilization of different zwitterionic phospholipid vesicles structures such as L--phosphatidylcholine (PC) and 1,2-didecanoyl-sn-glycero-3-phosphocholine (DPC) have been studied in aqueous bulk by using zwitterionic surfactant dimethylhexadecylammoniopropanesulfonate (HPS). This has been done by studying the aggregation of HPS in pure water and in the presence of 7–36 M of fixed concentrations of each lipid with the help of pyrene fluorescence intensity (I 1/I 3) measurements. The fluorescence measurements showed that HPS monomers undergo two kinds of aggregation process, which were identified by the three breaks in a plot of pyrene fluorescence versus HPS concentration. The first two breaks, C 1 and C 2, indicate the onset and completion of vesicle solubilization respectively, upon incorporation of HPS monomers into the vesicles and led to solubilization in the form of mixed micelles. This process was not clearly visible at low lipid concentration. We evaluated the partition coefficient (K), which defines the degree of partitioning of surfactant monomers into the vesicles with respect to the aqueous medium. A high K value of HPS-lipid aggregates indicates the stronger interactions between surfactant and lipid vesicles. The K values evaluated for PC and DPC are quite close to each other, which indicates that K values were independent of phospholipid chain length.  相似文献   

10.
Summary The foaminesses of bovine serum albumin solutions, (BSA) with and without buffer, salt and alcohol additives were measured by bubbling and their surface tensions were obtained as function of the timet. The timet DG , which is necessary to attain the equilibrium surface tension is long (15 h). The area requirement of a single surface adsorbed moleculeA j0 was obtained fromd/dc, wherec the surfactant concentration, by the Gibbs relation. By assuming the existence of a hydration complex, which consists of the surfactant andX water molecules, the,coordination numbers were estimated toX=12±1 fromA j0 and the surface requirement of the hydrophilic group of the surfactant. The dependences of CMC andX on the additives are discussed. A good relation prevails between andt DG . By applying the phase change model of Avrami for the adsorption and surface denaturation of theBS A, simple relation was found between the dimensionless surface tensionV = (0 - s) (st-s) and the timet:log (2.3 logV) =n logt + logb The good relations between andn as well as between andnb indicate the applicability of this model.
Zusammenfassung Das Schaumbildungsvermögen von Rinderserumalbumin-Lösungen (BSA), mit und ohne Zusätze wurde gemessen und ihre Oberflächenspannung als Funktion der Meßzeit ermittelt. Die Einstellzeitt DG der Gleichgewichtsoberflächenspannung eq ist lang. Zwischen undt DG läßt sich eine einfache Beziehung aufstellen. Aus der Änderung von eq mit der bedarf des adsorbierten Molekel ermitteln. Durch die Anwendung des Modells von Steinbach kann gezeigt werden, daß sich der Flächenbedarf durch die Bildung eines Assoziates erklären läßt, bei dem das Wasser komplexartig um das Protein-Molekül lagert. Für die Koordinationszahl der WassermoleküleX ergab sichX=12±1.Durch das Kristallwachstumsmodell von Avrami, das von überreiter auf Polymeradsorption angewendet wurde, läßt sich die Zeitabhängigkeit der Oberflächenspannung beschreiben. Das Schaumbildungsvermögen kann man als einfache Funktion der Konstanten dieser Avrami-Gleichung darstellen.
  相似文献   

11.
Summary Crystals of Co2(X 2O7)·2H2O,X=P/As were synthesized under hydrothermal conditions. Their crystal structures were determined by single crystal X-ray diffraction:a=6.334(1)/6.531(2),b=13.997(2)/14.206(4),c=7.637(1)/7.615(2)Å, =94.77(2)/94.74(2)°, space group P21/n,R=0.032/0.046,R w=0.028/0.034 for 2423/2042 reflections and 131/119 variables. Within the twoXO4 tetrahedra connected via a common corner to anX 2O7 group the average P-O bond lengths are approximately equal (1.540 and 1.543 Å), but As-O differs significantly (1.685 and 1.696 Å). A comparison with the isotypic Mn and Mg pyrophosphates shows a correlation between the ratio Me-O/X-O and the angle O-X-O.
Vergleich der Kristallstrukturen von Co2(X 2O7)·2H2O,X=P und As
Zusammenfassung Kristalle von Co2(X 2O7)·2H2O,X=P/As wurden unter Hydrothermalbedingungen synthetisiert. Ihre Kristallstrukturen wurden mittels Röntgenbeugung an Einkristallen bestimmt:a=6.334(1)/6.531(2),b=13.997(2)/14.206(4),c=7.637(1)/7.615(2) Å, =94.77(2)/97.74(2)°, Raumgruppe P21/n,R=0.032/0.046,R w=0.028/0.034 für 2423/2042 Reflexe und 131/119 Variable. In den beiden über eine gemeinsame Ecke zuX 2O7-Gruppen verknüpftenXO4-Tetraedern sind die mittleren P-O-Abstände ungefähr gleich (1.540 und 1.543 Å), hingegen differiert As-O signifikant (1.685 und 1.696 Å). Ein Vergleich mit den isotypen Mn- und Mg-Pyrophosphaten zeigt eine Korrelation zwischen dem Quotienten Me-O/X-O und dem WinkelX-O-X.
  相似文献   

12.
The stability constants (1(F)) of the monofluoro complex of Lu(III) and those (1(Cl)) of the monochloride solvent-shared ion-pair of Lu(III) have been determined in mixed solvents of methanol and water at 0.10 and 1.00 mol·dm–3 ionic strengths, respectively. The variation in ln1(F) with an increase in the mole fraction of methanol (X s) in the mixed solvent system showed an acute-angled convex inflection point at X s 0.12, an acute-angled concave inflection point at X s 0.22, and another acute-angled convex inflection point at X s 0.27. It was concluded that the first and the second convex inflection points denoted the CN of Lu3+ from CN = 8 to a mixture of CN = 8 and 7 and from CN = 8 and 7 to a mixture containing CN = 6, respectively. The concave point is the starting point of a change in the CN of Lu(III) in LuF2+ from CN = 8 to a mixture of CN = 8 and 7. The values of two inflection points of the CN around Lu3+ are consistent with the inflection points of the variation in the values of ln1(Cl) versus the dielectric constant of the mixed solvent.  相似文献   

13.
The enthalpies of transfer from water to aqueous surfactant solutions, H(WW+S), of polar additives have been determined as a function of the surfactant concentration at fixed additive concentration. The surfactants used are sodium dodecylsulfate (NaDS), dodecyltrimethylammonium bromide and dodecyldimethylamine oxide (DDAO). The additives used are iso-butanol t-butanol, butoxyethanol, phenol, benzene, tributylphosphine oxide (TBPO), octyldimethylphosphine oxide (ODPO), octydimethylamine oxide (ODAO), DDAO and NaDS. A maximum was observed in the plots of H(WW+S) vs. fsms curves for ODPO and ODAO in NaDS while a small minimum was observed for TBPO. The experimental data are rationalized on the basis of the pseudo-phase transition model for the micellization process and a mass action model for the distribution of the additive between aqueous and micellar phases. The standard free energies, enthalpies and entropies of transfer of the additives from the aqueous to the micellar phases are reported. The effect of different butanol isomers on the thermodynamics of solubilization in the micellar phase has been derived. The enthalpies of transfer of benzene are always negligible with respect to those of phenol while the free energies of transfer are always comparable. Studies of symmetrical and asymmetrical additives show that asymmetry causes an increase of the free energy of transfer due to the decrease of the entropy. The thermodynamics of transfer of NaDS from the aqueous to the DDAO micellar phases and of DDAO from the aqueous to the NaDS micellar phases are compared to the thermodynamics of micellization of the two surfactants; the formation of mixed micelles seems to be energetically unfavored with respect to the pure micelles.  相似文献   

14.
A multiple regression model was generated, which can satisfactorily estimate the association constants (K a ) for the inclusion complexation of -cyclodextrin with mono- and 1,4-disubstituted benzenes. It was found that lnK K a was correlated with the substituent molar refraction (R m ), hydrophobic constant () and Hammett constant ) of the guest compounds with a correlation coefficient of 0.95. The main driving forces for -cyclodextrin complexation was concluded to consist of van der Waals forces and hydrophobic interactions, while the influence of electronic effects was small.  相似文献   

15.
The aggregation of aqueous dodecylphosphonic acid (DPA) and dodecyltrimethylammonium hydroxide (DTAOH) mixtures was studied by several methods. The behavior of DPA-rich mixtures is close to that of pure DPA. This is probably due to the preservation of the hydrogen-bonded structure of the micellar headgroup layer. The behavior is almost ideal. Betweeny DPA =0.5 and 0.33 (y DPA being the mole fraction of DPA in the surfactant mixture), the hydrogen-bonded structure of the micellar headgroup layer is destroyed. A sort of micellar azeotrope is formed, and the maximum of non-ideal interaction between the two surfactants is attained aty DPA =0.4. Fory DPA <0.33 the system behaves as a common mixture of a cationic surfactant and a non-ionic one (DPA.2LTA). There is a phenomenon of counterion condensation on aggregates at concentrations over the CMC.  相似文献   

16.
The mean ionic activity coefficients of HCl () in the system HCl-H2SO4-H2O at 298 K were calculated by the Mikulin and MacKay-Perring methods and were then used for calculating the mixed thermo- dynamic dissociation constant of HCl (K m). The mean value of the constant proved to be equal to that found previously for aqueous solution of HCl, and deviations from the mean value are most likely due to the fact that, when calculating K m, incompleteness of dissociation of both electrolytes was neglected. The values calculated by the MacKay-Perring and Mikulin methods virtually coincide, within the determination and calculation errors, with the published data. This result confirms the suitability of the previously suggested procedure for determining the strictly thermodynamic mixed dissociation constants from the experimental data on the vapor pressure in combination with the mean ionic activity coefficients.  相似文献   

17.
Apparent molar heat capacities and volumes of pentanol (PentOH) 0.05m in dodecyltrimethylammonium chloride (DTAC), dodecyldimethylammonium chloride (DDAC) and dodecylamine hydrochloride (DAC) micellar solutions were measured at 25°C. They were assumed to approach the standard infinite dilution values and rationalized by means of previously reported equations. The distribution constant between the aqueous and the micellar phase and heat capacity and volume of pentanol in both phases were thus derived. The results show that the presence of methyl groups on the surfactant head group does not appreciably influence the apparent molar volume and heat capacity of pentanol in micellar phase and the free energy of transfer of pentanol from the aqueous to the micellar phase. Also, the apparent molar heat capacities of pentanol in micellar solutions as a function of surfactant concentration show evidence of two maxima for DAC and of one maximum for DTAC whereas no maxima were detected for DDAC. According to the literature data for alkyltrimethylammonium bromides these maxima can be ascribed to the presence of structural post-micellar transitions. It is shown that the C,PentOH vs. surfactant molality curve for DAC lies between that for hexadecyltrimethylammonium bromide and that for tetradecyltrimethylammonium bromide. This evidence, which is similar to that found for solubilities, agrees with the previously reported idea that the removal of a CH3 group from the head group of surfactant is equivalent to the introduction of a CH2 group in its hydrophobic moiety. By comparing data for DTAC with those for the corresponding bromide, the role of the nature of the counterion in the thermodynamics of solubilization of pentanol in micellar solutions is derived.  相似文献   

18.
Zusammenfassung Die Komplexe des Ni2+ mit o-Methylbenzamidoxim wurden in neutraler und in alkalischer Lösung spektrophotometrisch untersucht. Die Bildungskonstanten sindK 1=40 für 11 undK 2=1,7·102 für 12 in neutraler Lösung und 1 = =(3,92 ±0,2) · 104für 11 und lgK = lg 1 + lg 2 = 3,45 ± ±0,15 für 12 bei 25° und =1 in alkalischer Lösung.
Complex formation in the systemeNi 2+—o-methylbenzamide oxime
The complexes of Ni2+ with o-methylbenzamide oxime were investigated spectrophotometrically in neutral as well as in alkaline solution. The formation constants areK 1=40 for 11 andK 2=1.7·102 for 12 in the neutral solution and 1 = =(3.92 ±0.2) · 104 for 11 and lgK = lg 1 + lg 2 = 3.45 ± ±0.15 for 12 at 25° and =1 for the alkaline solution. *** DIRECT SUPPORT *** A3615139 00007
  相似文献   

19.
Fe(III) hydrolysis and fluoride complexation behavior was examined in 0.68 molal sodium perchlorate at 25°C. Our assessment of the complexation of Fe(III) by fluoride ions produced the following results: logF1 = 5.155, logF2 = 9.107, logF3 = 11.96, logF4 = 13.75, where logFn = 5.155=[FeF n (3-n)+ ][Fe3+]–1[F]–n. The stepwise fluoride complexation constants,FK n+1, obtained in our work (where logF K n+1 =logFn) indicate that K n+1/K n =0.072±0.01. Formation constants for equilibria, Fe3++nH2OFe(OH) n (3–n)+ +nH+, expressed in the form n * [Fe(OH) n (3-n)+ ][H+]n ,[Fe 3+]-1, were estimated as 1 * = –2.754, and 2 * –7. Our study indicates that the results of previous hydrolysis investigations include very large overestimates of Fe(OH) 2 + formation constants.  相似文献   

20.
In the aqueous mixtures of sodium alkylcarboxylate and alkyltrimethylammonium bromide, large unilamelar vesicles can be formed spontaneously or by sonication as the total carbon number in the HC chains is 19 (or larger). Vesicle formation can be influenced by changes of pH, molar ratio of the two surfactant components, and the polar head group of cationic surfactant. Micelles may coexist with the vesicles in these mixed systems. The larger hydrodynamic radius (200 nm) and aggregation number (800) illustrate that the shape of the micelle in 1:1 C9H19COONa–C10H21N(CH3)3Br is rod-like. In some mixed systems, the micelles can be transformed into stable vesicles by sonication — a phenomenon revealed for the first time. The surface-chemical properties of these catanionic surfactant solutions and the stabilities of vesicle have been studied systematically.  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