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 surface tensions of a mixed system of sodium tetradecyl sulfate (STS) and a series of polyoxyethyleneglycol-n-dodecyl ethers (nED) were measured at temperatures of 30 and 40 C, respectively. When the STS concentration was fixed, the surface tension curves of the mixed system showed a flat portion in the range of which the mixed micelles would be formed. With a decrease of the added STS concentration, the length of the flat portion in these curves decreased gradually, and finally disappeared in the presence of 0.05 mM STS. Furthermore, when the concentration of STS was fixed at 1.0 mM, which is below the critical micelle concentration (CMC), the flat portion length in the surface tension curves de-creased with decreasing ethylene oxide chain length in nED and with increasing temperature. The influence of increment of temper-ature was comparable to that for reducing the length of ethylene oxide chain in nED.  相似文献   

2.
Using equilibrium and dynamic surface tension measurements, we have studied the effect of the addition of poly(sodium 4-styrenesulfonate) (PSS), sodium chloride, and 1,4,7,10,13,16-hexaoxyacyclooctadecane (18-crown-6) on the surface properties of sodium bis(2-ethyl hexyl sulfosuccinate) (AOT). The addition of PSS or NaCl weakly increases the maximum packing of AOT, whereas the presence of 18-crown-6 slightly decreases the maximum surface coverage. The surfactant adsorption kinetics on the interface is a diffusion-controlled process. The two asymptotic solutions at long times and at short times to the classic Ward Tordai equation were used to fit dynamic results. At long times there is evidence of the existence of an electrostatic barrier at high surfactant concentration when using pure AOT and AOT mixed with PSS. In binary mixtures of AOT with sodium chloride or 18-crown-6, the electrostatic barrier is not observed over the surfactant concentration range studied.  相似文献   

3.
The poor collecting performance of fatty acids at low temperatures is a problem in mineral flotation. In this study, the floatability of scheelite at 283 K was studied using sodium oleate and binary mixtures of sodium oleate and lauryl alcohol polyoxyethylene ether (MOA-9) as the collector, and interactions between the two surfactants at 283 K were investigated by the means of surface tension and steady state fluorescence measurements. The flotation experiment results show that the collecting performance of the mixed anonic/nonionic collector is stronger than that of single surfactant, and the best molar ratio of sodium oleate to MOA-9 is 10:1 at pH 10. The surface tension results show that the critical micelle concentration value of sodium oleate, MOA-9 and the binary solution is about 2 × 10?4, 2.5 × 10?4, and 1.2 × 10?4 mol/l, respectively. Compared with single surfactants, the mixture exhibits better surface activities since it is more efficient at decreasing the air–water surface tension. The steady state fluorescence results indicate that the hydrophobic parts of MOA-9 molecules insert into the hydrophobic region of sodium oleate micelles to form larger and less compact mixed micelles. The decrease in zeta potential of scheelite treated with the mixed collector indicates the presence of MOA-9 can enhance the chemical adsorption of sodium oleate on the scheelite surface. Therefore, the mixed anonic/nonionic collector can enhance the flotation behavior of scheelite at 283 K.  相似文献   

4.
The effects of various detergent builders on surface properties of sodium dodecyl sulfate (SDS) solutions have been investigated by measuring the equilibrium surface tension and hysteresis curves of surface pressure vs surface area under dynamic conditions. While the adsorbed layer of SDS under static conditions is in the liquid-expanded state, a two-dimensional phase transition of the adsorbed layer from liquid-expanded to condensed-film can be observed on compression at higher rates beyond the saturated adsorption. For the SDS solution containing excess NaCl (1×10−1 M), the adsorbed layer of SDS is already in the condensed state under static condition due to a depression in the dissociation of SDS. It has been found that some polyelectrolytes, as well as conventional ionic builders, contribute to increasing the surface excess concentration of SDS together with promoting micelle formation and also to stabilizing the adsorbed layer of SDS. On the other hand, poly(vinyl alcohol) (PVA) as a nonionic polymer is adsorbed at the air-water interface in preference to SDS. In this case, the hysteresis of the dynamic surface behavior of the mixed solution can be attributed to the structural change of the adsorbed layer of PVA rather than SDS, which is different from the other ionic builders.  相似文献   

5.
The DNA damage caused by TiO2 under in vitro conditions by UV-A exposure in the presence of anionic vesicles of Aerosol OT (AOT) was investigated. The supercoiled form (S) in DNA plasmids was converted to the linear form (L) via the relaxed form (R). The DNA damage was slower in the presence of AOT vesicles prepared in aqueous NaCl solution. Moreover, the presence of AOT vesicles in solution after 6 h of UV irradiation was confirmed with an optical microscope. Probably, a fraction of the DNA was protected by random trapping during sonication. However, the addition of NaCl needed for the vesicle formation can decrease the TiO2 activity. On the other hand, in the absence of vesicles the NaCl concentration led to a profound influence on the adsorption of DNA onto the TiO2 surface. During UV irradiation, the degradation rate of DNA increased with increasing the salt concentration. Solutions containing vesicles were prepared at various NaCl concentrations between 10 mM and 75 mM. Consequently, the salt concentration had no significant effect on the DNA damage. The presence of NaCl can play a deleterious role during the photoinduced process. However, the encapsulation of a fraction of DNA is not excluded. In such conditions, the DNA could be protected against the reactive oxygen species.  相似文献   

6.
《分离科学与技术》2012,47(4):999-1011
Abstract

Three sulfonic acid extractants, dinonylnaphthalene sulfonic acid (HDNNS), didodecylnaphthalene sulfonic acid (HDDNS) and di-(2-ethyl-hexyl) sodium sulfosuccinate (Aerosol OT, AOT) are compared as to their effects on the extraction of nickel with LIX63. The acidic extractants interact synergistically with the oxime. Interfacial tension results are presented which demonstrate that the sulfonates form reversed micelles in non-polar organic solvents. It is proposed that the reversed micelles catalyze the extraction by specific solubilization of both the metal and the extractant, resulting in an increase in the interfacial concentration of the reacting species. The ability of LIX63 to chelate with nickel without deprotonating permits the synergism to occur at low pH.  相似文献   

7.
The surface tension of sodium dodecyl sulfate was determined in the presence of nonionic surfactant. The nonionic surfactant used was homogeneous pentaethyleneglycol-n-dodecylether which has no poisson distribution of added mole numbers of ethyleneoxide. The concentration of sodium dodecyl sulfate was changed at various fixed concentrations of homogeneous pentaethyleneglycol-n-dodecylether. Two inflection points were observed on each surface tension curve, and the surface tension maintained a constant value between the inflections, regardless of the concentrations of added homogeneous pentaethyleneglycol-n-dodecylether. However, the concentrations of sodium dodecyl sulfate at the inflections were affected at the concentrations of homogeneous pentaethyleneglycol-n-dodecylether, respectively. When the homogeneous pentaethyleneglycol-n-dodecylether concentrations were fixed below the critical micelle concentration, the concentration of sodium dodecyl sulfate at the lower inflection point decreased with increasing concentration of homogeneous pentaethyleneglycol-n-dodecylether, while that of the higher was hardly influenced. If the homogeneous pentaethyleneglycol-n-dodecylether concentrations were fixed above the critical micelle concentration, the sodium dodecyl sulfate concentrations at the two inflections both increased with an increment of homogeneous pentaethyleneglycol-n-dodecylether concentration. These results were interpreted in terms of mixed surface layer and mixed micelles consisting of sodium dodecyl sulfate and homogeneous pentaethyleneglycol-n-dodecylether. Also, the surface tension curves of homogeneous pentaethyleneglycol-n-dodecylether in the presence of sodium dodecyl sulfate, as well as those of sodium dodecyl sulfate with given concentrations of homogeneous pentaethyleneglycol-n-dodecylether, were studied.  相似文献   

8.
The new triple chain surfactant was firstly prepared by ring-opening and a quaternization reaction with 2-(chloromethyl)-oxirane, tertiary amine and propane-1,2,3-triol as starting materials. The surfactant was characterized by MS and elementary analysis. Air–Liquid surface tension was measured by using a tensiometer at temperatures from 298 to 318 K respectively and in various salt solutions (NaX). The results show that with the increase of temperature, the values of critical micelle concentration (cmc) can achieve a minimum at 308 K. The maximum surface adsorption capacity Γ decreases, while the minimum molecule A min area increases. With the increase in temperature and the decrease in counterionic diameter, the values of cmc decrease from 8.09 × 10−5 to 5.52 × 10−5 mol/L, and the adsorption capacity Γ increase from 1.80 × 10−10 to 2.74 × 10−10 mol/m2, while there are almost no influences to surface tension (γcmc). The free energy of micelle formation is negative (−85.2 to −92.0 kJ/mol).  相似文献   

9.
The properties of some well-characterized sodium linear decyldiphenylether (C10DPE)sulfonates have been studied. Among the properties investigated are dynamic and equilibrium surface tension, critical micelle concentration (CMC), area per molecule at the aqueous solution/air interface, wetting time by the Draves technique, foaming by the Ross-Miles method, solubilization, and hydrotropy. The decyldiphenylether moiety appears to be equivalent to a terminally substituted straight alkyl chain of 16 carbon atoms. The trialkyl- and dialkyl-mono-sulfonates have solubilities of < 0.01 g/dm3 in water, but are readily soluble in hexane. The didecyldiphenyl ether disulfonate (DADS) has a very low CMC value (1.0 × 10−5 mol dm−3) in aqueous 0.1 N Na+ solution (NaCl), characteristic of surfactants with two hydrophilic and two hydrophobic groups. It also has a much larger area per molecule at the aqueous solution/air interface than the monodecyldiphenyl-ether monosulfonate (MAMS) and a much higher surface tension at the CMC. MAMS has a much lower surface tension at a surface age of 1 second (γ1s) than either DADS or the monodecyldiphenylether disulfonate (MADS). In agreement with γ1s and γeq values, wetting times increase in the order: MAMS < DADS < MADS and initial foam heights decrease in the order: MAMS > DADS > MADS. Solubilization for three water-insoluble surfactants decreases in the order: DADS > MAMS > MADS, while hydrotropy is most pronounced with the disulfonates.  相似文献   

10.
Measurements of the surface tension of aqueous solution of mixtures of sodium dodecyl sulfate (SDDS) with methanol and ethanol in SDDS concentration range from 10−5 to 10−2 M and mixtures of sodium hexadecyl sulfonate (SHS) with methanol and ethanol at SHS concentration from 10−5 to 8 × 10−4 M and for methanol and ethanol from 0 to 21.1 and, 11.97 M, respectively, were carried out at 293 K. Moreover, the surface tension of aqueous solution mixtures of SDDS with propanol in the concentration range from 0 to 6.67 M taken from the literature was also considered. The results obtained indicate that it is possible to describe the relationship between the surface tension and molar concentration or molar fraction of alcohol by Szyszkowski and Connors equations. However, the Fainerman and Miller equation allows us to predict the isotherm of the surfactant tension at constant anionic surfactants concentration at which their molecules are present in the solution in the monomeric form if the molar area of surfactants and alcohols can be determined. Based on the surface tension isotherms, the Gibbs surface excess of anionic surfactants and alcohols concentration at water–air interface was determined and then recalculated for Guggenheim-Adam surface excess concentration of these substrates, and next the molar fraction of alcohols and surfactants in the surface layer was determined. These molar fractions were discussed with regard to surfactant and alcohol standard free energy of adsorption at the water–air interface determined from Langmuir and Aronson and Rosen equations. The standard free energy of adsorption determined in these ways was compared to that deduced on the basis of pC20 and Lifshitz van der Waals-components of the anionic surfactant and alcohol tails.  相似文献   

11.
The surface tension of disodium hexadecyl diphenyl ether disulfonate (C16‐MADS) was measured at different NaCl concentrations (0.00–0.50 mol L?1) and temperatures (298.0–318.0 K) using the drop‐volume method. The results show that, with increasing temperature, the critical micelle concentration (CMC) of C16‐MADS increases slightly, but the maximum surface adsorption capacity (Γmax) at the air–water interface decreases. When the concentration of NaCl was increased from 0.00 to 0.50 mol L?1, the CMC of C16‐MADS decreased from 1.45 × 10?4 to 4.10 × 10?5 mol L?1, but the surface tension at the CMC (γcmc) was not affected. When the concentration of NaCl was increased at 298.0 and 303.0 K, the Γmax of C16‐MADS increased. When the temperature was increased from 308.0 to 318.0 K, the surface excess concentration (Γmax) of C16‐MADS abnormally decreased from 2.26 to 1.41 μmol m?2 with increasing NaCl concentration. The micellization free energy () decreased from ?63.98 to ?76.20 kJ mol?1 with increase of temperature and NaCl concentration. The micellar aggregation number (Nm) of disodium hexadecyl diphenyl ether disulfonate (C16‐MADS) was determined using the molecule fluorescence probe method with pyrene as probe and benzophenone as quencher. The results show that an appropriate Nm could be measured only at surfactant concentration above the CMC. The Nm increased with an increase in C16‐MADS concentration, but the micropolarity in the micelle nucleus decreased. The temperature had little effect on Nm. Compared with typical single hydrophilic headgroup surfactants, aggregates of C16‐MADS exhibit different properties.  相似文献   

12.
Linear alkanes with two pyrrolidone groups in both termini [abbreviation: 1,n-dipyrrolidonyl alkane (1,n-DPA) (n=3, 5, 9, and 12)] were synthesized. Their surface activities were evaluated in terms of surface tension, and compared with those of a few kinds of surfactants. Among 1,n-DPAs, 1,3-, 1,5-, and 1,9-DPAs were fully soluble in water and surface-active: especially 1,9-DPA showed a surface tension value of γW/A 32.8 mN/m at 5.9×10−2 mol/L or 1.7 wt% and was suggested as a nonionic surfactant with simple structure and moderate activity. Critical miceller concentration (CMC) of 1,9-DPA was 1.45×10−2 mol/L. Surface tension values of 1,5- and 1,9-DPAs were lower than those of the corresponding monofunctional pyrrolidones, N-propyl and N-pentylpyrrolidones. Minimum area per surfactant molecule (Amin) decreased with increase of methylene chain length in the series of 1,n-DPAs. The results suggested that 1,n-DPA favorably chooses a loop structure at air/water interface.  相似文献   

13.
琥珀酸二辛酯磺酸钠对海星皂苷在超临界CO2中的增溶作用   总被引:1,自引:0,他引:1  
文震  党志  宗敏华  赵金华  李靖 《日用化学工业》2006,36(3):148-150,158
利用琥珀酸二辛酯磺酸钠(AOT)/混合醇/水溶液为助溶剂,以乙醇助溶剂为参照,考察了表面活性剂对海星皂苷在超临界CO2介质中的增溶作用。探讨了表面活性剂浓度、助表面活性剂组成、复合表面活性剂配比等因素对海星皂苷增溶的影响,结果表明:摩尔比为4∶1、浓度为0.05 moL/L的AOT/辛基酚聚氧乙烯醚(OP-10)复配表面活性剂,对海星皂苷具有良好的选择性增溶作用,海星皂苷的萃取率为2.40%,提取物中海星皂苷质量分数为58.99%,分别是使用乙醇助溶剂的4.08、2.18倍。  相似文献   

14.
Solubilization of water and aqueous NaCl in mixed reverse micelles (RMs) comprising sodium bis(2‐ethylhexyl) sulfosuccinate (AOT), and polyoxyethylene (20) sorbitan trioleate or polyoxyethylene (20) sorbitan monooleate has been studied at different compositions (Xnonionic = 0–1.0) at a total surfactant concentration, ST = 0.10 × 103 mol m?3 in biocompatible oils of different chemical structures; viz., ethyl oleate (EO), isopropyl myristate (IPM) and isopropyl palmitate (IPP) at 303 K. The enhancement in water solubilization (i.e., synergism) has been evidenced by the addition of nonionic surfactant to dioctyl sulfosuccinate/oil(s)/water systems. Addition of NaCl in these systems at different Xnonionic enhances their solubilization capacities further until a maximum, ωNaCl,max is reached. ωNaCl,max and [NaCl]max (concentration at which maximization of NaCl solubilization occurs) depend on type of nonionic surfactant, its content (Xnonionic) and oil. A new solubilization efficiency parameter (SP*water or SP*NaCl) has been proposed to compare solubilization phenomena in these oils. The energetic parameters of the desolubilization process of water or aqueous NaCl in single and mixed RMs have been estimated. Energetically, the water dissolution process in oil has been found to be more exothermic as well as more organized in IPP. Overall, the dissolution of water and aqueous NaCl in mixed RMs is entropically driven process. Conductance behavior of these systems in the presence of NaCl has been investigated under different [NaCl] at 303 K. An attempt has been made to give an insight to the mechanism of solubilization phenomena, percolation in conductance and microstructures vis‐à‐vis role of biocompatible oils in these systems.  相似文献   

15.
The simultaneous influence of NaCl, KCl and KI salts and well known sodium dodecyl sulfate (SDS) surfactant on the interfacial tension (IFT) of conventionally used chemical system of toluene-water was studied. The concentration range of salts was within (0.010 to 0.075) mol/dm3 and of surfactant within (1.7 to 26.0)×10?5 mol/dm3. SDS adsorption on interface is highly intensified in the presence of salts and IFT can reach to 67.1% of its initial value. Accordingly, the role of ions was investigated and the order of salts effectiveness was revealed as KI?KCl>NaCl. The obtained data, with both effects, were nicely reproduced using an equation of state, based on Gibbs adsorption equation and the Langmuir isotherm. Two relevant important adsorption parameters exhibited reasonable variations. Furthermore, the general revealed linear variation of IFT with a previously defined “effective concentration” indicates the strong influence of the surfactant counterions.  相似文献   

16.
The influence of sodium fluoride additives on the physicochemical properties of glasses in the Na2O-B2O3 systems is investigated. The introduction of sodium fluoride into the Na2O · 2B2O3 and Na2O · 3B2O3 glasses leads to an increase in the electric conductivity. The temperature-concentration dependence of the electric conductivity has been investigated. It is shown that, in glasses of the NaF-Na2O · 2B2O3 system, an increase in the volume concentration of sodium ions from 2.4 × 10−2 to ∼3 × 10−2 mol/cm3 is accompanied by an insignificant decrease in the activation energy Eσ from 1.44 to 1.38 eV and a sharp (by a factor of ∼30) increase in the electric conductivity. In glasses of the NaF-Na2O · 3B2O3 system, an increase in the concentration of sodium ions from 1.8 × 10−2 to ∼2.3 × 10−2 mol/cm3 brings about an increase in the electric conductivity by a factor of approximately 100 and an increase in Eσ from 1.6 to 1.83 eV. A further increase in the concentration of sodium ions (up to 2.5 × 10−2 mol/cm3) virtually does not affect the electric conductivity and Eσ. At the same concentration of sodium ions (∼2.46 × 10−2 mol/cm3) in the 9.8NaF · 90.2[Na2O · 2B2O3] and 57.1NaF · 42.9[Na2O · 3B2O3] glasses, the electric conductivity and the activation energy are considerably higher in the glass with a larger fluorine content. The regularities revealed are interpreted in the framework of the microinhomogeneous glass structure.  相似文献   

17.
The influence of sodium fluoride additives on the physicochemical properties of glasses in the Na2O-B2O3 systems is investigated. The introduction of sodium fluoride into the Na2O · 2B2O3 and Na2O · 3B2O3 glasses leads to an increase in the electric conductivity. The temperature-concentration dependence of the electric conductivity has been investigated. It is shown that, in glasses of the NaF-Na2O · 2B2O3 system, an increase in the volume concentration of sodium ions from 2.4 × 10−2 to ∼3 × 10−2 mol/cm3 is accompanied by an insignificant decrease in the activation energy Eσ from 1.44 to 1.38 eV and a sharp (by a factor of ∼30) increase in the electric conductivity. In glasses of the NaF-Na2O · 3B2O3 system, an increase in the concentration of sodium ions from 1.8 × 10−2 to ∼2.3 × 10−2 mol/cm3 brings about an increase in the electric conductivity by a factor of approximately 100 and an increase in Eσ from 1.6 to 1.83 eV. A further increase in the concentration of sodium ions (up to 2.5 × 10−2 mol/cm3) virtually does not affect the electric conductivity and Eσ. At the same concentration of sodium ions (∼2.46 × 10−2 mol/cm3) in the 9.8NaF · 90.2[Na2O · 2B2O3] and 57.1NaF · 42.9[Na2O · 3B2O3] glasses, the electric conductivity and the activation energy are considerably higher in the glass with a larger fluorine content. The regularities revealed are interpreted in the framework of the microinhomogeneous glass structure.  相似文献   

18.
The effects of an anionic surfactant, sodium dodecyl sulfate (SDS), on the micellar properties of a nonionic surfactant such as homogeneous heptaethylene glycol n-dodecyl ether (7ED) have been studied by the charge transfer solubilization of 7,7,8,8-tetracyanoquinodimethane, pNa, and electric conductivity measurements. Attention has been paid to changes in packing of polyethylene oxide chains in the mixed micelle and to binding of the counterions onto the micelle surface. All measurements were made on solutions ranging in 7ED concentration from 1 × 10−6 to 1×10−1 M, while the SDS concentration was maintained constant. It has been shown that the binding of Na+ ions to the mixed micelle occurs in the 7ED concentration region where the packing of polyethylene oxide chains in the micelle is loose, while release of Na+ ions is observed when the packing is compact. The results of electric conductivity correspond well with those mentioned above. However, in the region of high 7ED concentration, the decreasing mobility of the mixed micelles affects the electric conductivity more than the increasing degree of ionic dissociation of the micelle.  相似文献   

19.
通过表面张力的测定,研究了浓度、温度对溶液表面张力的影响,同时考察无机盐浓度对溶液表面张力的影响.结果表明:十二烷基苯磺酸钠溶液表面张力随浓度的增加、温度的升高和无机盐浓度逐渐的增大而降低,最终确定其CMC值为1.47×10-3mol/L,由于键能和水合离子半径的原因,NaCl对十二烷基苯磺酸钠表面张力的影响明显于KCl.  相似文献   

20.
《分离科学与技术》2012,47(16):2117-2129
Abstract

The transport behavior of basic amino acids (BAA), such as arginine (Arg), histidine (His), and ornithine (Orn), through an organic liquid membrane system (LMS) was investigated. The LMS was composed of two aqueous phases (Phases I and II) separated by an organic phase of chloroform containing sodium di-(2-ethylhexyl) sulfosuccinate (Aerosol OT, AOT). The amount of BAA that moved from Phase I at pH 3 into the organic phase increased with increasing AOT concentration (2–10 mM). The relative amount of extracted BAA was in the following order: Arg > His > Orn. On the other hand, the release of BAA from the organic phase into Phase II at pH 10 did not depend upon their amount in the organic phase. Arg was difficult to release. The relative amount of released BAA was in the following order: Arg = His > Orn. BAA were extracted from Phase I at pH 5 into the organic phase containing 4 mM AOT because they exist as cationic species. Other amino acids possessing nonionic residues were untransportable under these conditions except leucine, tryptophan, and phenylalanin, which have highly hydrophobic residues. However, they were transportable in their cationic forms at pH 1. These transport phenomena are essentially controlled by the interaction of the anionic group of AOT and a cationic form. These results suggested that BAA can be separated from most amino acids under an appropriate pH by using AOT.  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