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.
Surfactants are widely spread in nature and are increasingly used in industry as wetting, cleaning and disinfecting agents. Recently, there are newly discovered trisiloxanes and other silicone based surfactants which show very unusual spreading behaviour. Although a number of experimental and theoretical investigations have been carried out, the underlying spreading mechanism remains unclear. Experiments using trisiloxanes and comparison with 3 different poly(ethylene glycol) monododecyl ethers (C12E4, C12E5, and C12E6) surfactants were performed to understand the influence of Marangoni force as the driving force for the spreading. We then compared our experimental results to available theoretical predictions in the literature. The obtained experimental data showed the opposite trend as compared with the theoretical predictions developed for regular surfactants. The latter is assumed to be a special feature of “superspreaders”. The article is published in the original.  相似文献   

2.
The densities and the ultrasonic speeds of the aqueous solutions of 2-(2-hexyloxyethoxy)ethanol (C6E2) were measured over the entire range of mole fractions at 5°C. Excess molar volumes V E were readily calculated from the densities. The densities, in combination with the ultrasonic speeds, furnish estimates of the molar (and excess molar) isentropic compressibilities K S and the deviations u D of the ultrasonic speeds from the values calculated for ideal mixtures. Radical changes in the mole fraction derivatives of the excess molar properties of the (C6E2 + water) system, in the vicinity of an amphiphile mole fraction of 0.003, indicate that C6E2 like C6E3 is capable of micelle formation. Our data have been compared with those reported earlier for (C4E2 +, C2E2 +, and C6E3 + water). We have employed both mass action and pseudophase approaches to data analysis, together with the four-segment model approach.  相似文献   

3.
The excess molar volumes of eight binary systems formed of propyl or butyl formate with four n-alkanes (from C6 to C9) have been determined at 25°C and atmospheric pressure. The data were obtained indirectly from densities measured experimentally with a vibrating-tube densimeter and then compared with those estimated using the Nitta model. This method yields good predictions of the symmetry of the v E curves given an average overall error for all the systems analyzed here smaller than 6 percent.  相似文献   

4.
Density and viscosity of binary systems water-nonionic surfactants poly(ethylen-oxide) alkyl alcohols type, [CnH2n+1(OCH2CH2)mOH, CnEm], have been studied. The partial molar volumes in the dilute solution range and the viscosity B-coefficients were calculated. The nonionic surfactants partial molar volumes were compared with those of ethylene glycol and poly(ethylenglycol) (PEG). The comparison shows that the ethoxy unit volume, (OCH2CH2), seems to be independent of the particular system. The consequences of this are discussed. A model for interpreting the experimental B values has been proposed. The model treats the macroscopic viscosity as the superimposition of different local effects. The following surfactants have been considered: C5E1, C5E2, C6E1, C6E2, C6E3, C6E4.  相似文献   

5.
 Monodisperse vesicular structures have been produced spontaneously from an almost structure-less aqueous amphiphilic system: water/di(ethylene glycol) mono-hexyl ether (C6E2)/sodium cholate. Individually, neither C6E2 nor the bile acid produces noticeable aggregate structures in an aqueous solution. However, the presence of small amounts of bile acid in the C6E2/water binary system is found to produce large microstructures, besides pushing the miscibility gap to higher temperatures. Dynamic light scattering studies indicate the presence of very monodisperse structures of sizes ranging from 15 to 50 nm in radii. The radii of these structures show strong dependence on the concentrations of the components and on the distance in temperature from the cloud point. For a given set of conditions the sizes are very stable and reproducible. Electron microscopy and conductivity measurements confirm these structures as vesicles. These may be formed due to the association of bile acid with C6E2 producing the geometrical parameters necessary for the formation of vesicles. Received: 23 January 1998 Accepted: 11 June 1998  相似文献   

6.
Abstract

The liquid-crystalline phases of the system hexaethylene glycol n-dodecyl methyl ether (C12E6C1)/water are studied by deuteron NMR spectroscopy. Information about the molecular orientation is derived from the quadrupole splittings of two selectively deuteriated derivatives of C12E6C1, one deuteriated at the α-position of the alkyl chain, the other at the methoxy group. The temperature dependence of the quadrupole splittings reveals a continuous decrease of orientational order on approach to the macroscopic phase transformations. This behaviour is explained by an increase of defects and fluctuations in the microstructure of the mesophases.  相似文献   

7.
Values for 〈ΔEdown〉, the average downward energy transferred from the reactant to the bath gas upon collision, have been obtained for highly vibrationally excited undeuterated and per-deuterated isopropyl bromide with the bath gases Ne, Xe, C2H4, and C2D4, at ca. 870 K. The technique of pressure-dependent very low-pressure pyrolysis (VLPP) was used to obtain the data. For C3H7Br, the 〈ΔEdown〉 values (cm?1) are 490 (Ne), 540 (Xe), 820 (C2H4), and 740 (C2D4), and for C3D7Br, 440 (Ne), 570 (Xe), 730 (C2H4), and 810 (C2D4). The uncertainties in these values are ca. ±10%. The 〈ΔEdown〉 values for the inert bath gases Ne and Xe show excellent agreement with the theoretical predictions of the semi-empirical biased random walk model for monatomic/substrate collisional energy exchange [J. Chem. Phys., 80 , 5501 (1984)]. The relative effects of deuteration of the reactant molecule on 〈ΔEdown〉 also compare favorably with the predictions of this theoretical model. Extrapolated high-pressure rate coefficients (s?1) for the thermal decomposition of reactant are 1013.6±0.3 exp(?200 ± 8 kJ mol?1/RT) for C3H7Br and 1013.9±0.3 exp(?207 ± 8 kJ mol±1/RT) for C3D7Br, which are consistent with previous studies and the expected isotope effect.  相似文献   

8.
The characterization of binary and ternary complexes of benzoate, lauryl hexaoxyethylene (C12E6) and -CD is presented. The complexation equilibrium was characterized by UV-Vis spectrophotometry, titration microcalorimetry, capillary electrophoresis, and 2D ROESY 1H-NMR. Results suggested that -CD forms one complex with C12E6in the stoichiometric ratio of -CD : C12E61.5 : 1, with a stability constant 1.3 × 105 M-1.5. The 2-D ROESY 1H-NMR spectrum indicated that C12E6is included inside the -CD cavity. The primary binding site of C12E6 is on the lauryl subunit of this molecule. Analogous to a previously reported study of -CD, the combination of -CD and C12E6precipitated from the solution. Addition of benzoate seemed to dissolve the precipitate and nearly doubled the apparent stability constant of the complex. Results from the various techniques supported formation of ternary complexes between -CD, C12E6, and benzoate.  相似文献   

9.
Phase changes in binary systems of poly(ethylene glycol) dodecyl ethers, C12H25(OC2H4)xOH (x=5, 6 and 8) with water have been studied between –40 to 100 C by differential scanning calorimetry. A number of transitions, including the liquidliquid phase separation, were seen and the transition temperatures and enthalpy changes were determined. The observed temperatures were generally in good agreement with reported phase diagrams. In the C12E5 system, three of the four three-phase lines were seen and a more complete phase diagram is suggested for the water-rich part of the system. Most of the phase changes seen above 0 C are accompanied by small endothermic enthalpy changes of 0.7 to 0.9 kJ (mol C12Ex)–1, independent of system studied, type of transition and transition temperature. Water-rich solutions and liquid crystalline phases separate upon freezing into ice and crystals of hydrated amphiphile. Eutectics are developed at the following temperatures and compositions: C12E5 –3.0 C and 54 wt % C12E5; Q2E6 –4.5 C and 54 wt % C12E6, C12E8 –3.8 C and 49 wt % C12E8.  相似文献   

10.
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   

11.
The inversion of the reactivity order of R2C6H4SC1 in AdE-reactions with R1C6H4CHCH2 owing to the variations in R1 and R2 has been found and interpreted in terms of two kinetically distinguishable AdE-2 mechanisms.  相似文献   

12.
The zero-field splitting parameters of C6H6 and C6D6 in a borazole crystal host have been measured. The absolute value of the parameter E is three times as large as for C6H6 in C6D6. The relative rate constants for radiative decay in some vibronic bands have been obtained by MIDP methods. In the OO band of the phsophorescence spectrum the radiation stems from the upper zf level indicating a quinoid electronic structure for benzene in borazole, i.e. opposite to that for C6H6 in C6D6.  相似文献   

13.
Using the Picker flow calorimeter, excess molar enthalpies H E have been obtained at 25°C for mixtures of 1,2-, 1,3- and 1,4-cis- and trans-dimethylcyclohexane and cis- and trans-decalin with n-hexadecane and the highly branched C16 isomer, 2,2,4,4,6,8,8-heptamethylnonane. Values of H E are also obtained for cis- and trans-decalin mixed with C6, C7, and C9 isomers. Anomalously low values of H E occur for normal alkanes mixed with cycloalkanes in the di-equatorial configuration, suggesting the presence of a negative contribution in H E possibly due to a restriction of n-alkane molecular motion by the flat, plate-like cycloalkane.  相似文献   

14.
Excess molar volumes V E and excess molar heat capacities C P E at constant pressure have been measured, at 25°C, as a function of composition for the four binary liquid mixtures propylene carbonate (4-methyl-1,3-dioxolan-2-one, C4H6O3; PC) + benzene (C6H6;B), + toluene (C6H5CH3;T), + ethylbenzene (C6H5C2H5;EB), and + p-xylene (p-C6H4(CH3)2;p-X) using a vibrating-tube densimeter and a Picker flow microcalorimeter, respectively. All the excess volumes are negative and noticeably skewed towards the hydrocarbon side: V E (cm3-mol–1) at the minimum ranges from about –0.31 at x1=0.43 for {x1C4H6O3+x2p-C6H4(CH3)2}, to –0.45 at x1=0.40 for {x1C4H6O3+x2C6H5CH3}. For the systems (PC+T), (PC+EB) and (PC+p-X) the C P E s are all positive and even more skewed. For instance, for (PC+T) the maximum is at x 1,max =0.31 with C P,max E =1.91 J-K–1-mol–1. Most interestingly, C P E of {x1C4H6O3+x2C6H6} exhibits two maxima near the ends of the composition range and a minimum at x 1,min =0.71 with C P,min E =–0.23 J-K–1-mol–1. For this type of mixture, it is the first reported case of an M-shaped composition dependence of the excess molar heat capacity at constant pressure.Communicated at the Festsymposium celebrating Dr. Henry V. Kehiaian's 60th birthday, Clermont-Ferrand, France, 17–18 May 1990.  相似文献   

15.
Using an oxenoid model, we investigated dependences of carcinogenic potency of the benzenes C6H5‐X on a nature of substituents X. According to the model, a P450 enzyme breaks a dioxygen molecule and generates the oxens, which readily react with substrates. We suggest that a stability of the intermediate OC6H6‐X with tetrahedrally coordinated C atom relative to the molecule C6H5‐X determines a rate of substrate biotransformation. Using MO LCAO MNDO approach, we calculated the total energies of molecules C6H6‐X and arene oxides OC6H6‐X. A difference ΔEmin of these values determines activation energy of oxidation reaction. The compounds with the low ΔEmin values are noncarcinogenic. Benzene derivatives with high ΔEmin values belong to carcinogenic compounds series. The carcinogenicity of amino‐ and nitro‐substituted benzenes is also determined by N‐oxidation of amino and reduction of the nitro group. As the phenylhydroxylamines XC6H4NHOH and nitrenium ions XC6OH4NH+ are the common metabolites of the nitro‐ and amino‐substituted benzenes and nitrenium ions XC6H4NH+ are the ultimate carcinogens, we use the differences ΔEN = E(XC6H4NH+) ? E(XC6H4NHOH) as the second parameter characterizing the carcinogenic activity of amino‐ and nitro‐substituted benzenes. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

16.
Liquid–liquid equilibria of systems water (A) + CiEj surfactant (B) + n-alkane (C) have been modeled by a mass-action law model previously developed and so far successfully applied to a series of binary water + CiEj systems and to the ternary system water + C4E1 + n-dodecane. These calculations provide the basis for the presented modeling. The aqueous systems give information about the association constants and the χAB-parameter of the Flory–Huggins theory and the ternary C4E1-system provides universal temperature functions for the χAC- and the χBC-parameter. The three-phase equilibrium for seven ternary CiEj systems (i = 6–12, j = 3–6) has been calculated by fitting one additional parameter for each of both temperature functions to the characteristic “fish-tail” point. The agreement with the experimental data is reasonably well. For systems with very small three-phase areas the results can considerably be improved by individual temperature functions that incorporate the experimental temperature maximum of the “fish” into the parameter fit. Based on the parameters of the system water + C8E4 + n-C8H18 the “fish-shaped” phase diagram of the system water + C8E4 + n-C14H30 was predicted reasonably well.  相似文献   

17.
We have observed temperature-dependent X-trap phosphorescence in C6H6 crystals doped with C6D6. The observed X-trap is identified as C6H6 which is perturbed by the presence of C6D6 in a nearest-neighbor crystal site. Zero-field optical-microwave double resonance measurements on the X-traps give for the zero-field anisotropy parameter |E| a value of 20 ± 5 MHz. This is an order of magnitude smaller than that observed for C6H6 in C6D6 and indicates for the X-traps a nearly hexagonal distribution of spin density in the 3B1u state. This difference shows that the electronic structure of the 3B1u state is sensitive to subtle changes in its environment.  相似文献   

18.
Stable N‐heterocyclic carbene analogues of Thiele and Chichibabin hydrocarbons, [(IPr)(C6H4)(IPr)] and [(IPr)(C6H4)2(IPr)] ( 4 and 5 , respectively; IPr=C{N(2,6‐iPr2C6H3)}2CHCH), are reported. In a nickel‐catalyzed double carbenylation of 1,4‐Br2C6H4 and 4,4′‐Br2(C6H4)2 with IPr ( 1 ), [(IPr)(C6H4)(IPr)](Br)2 ( 2 ) and [(IPr)(C6H4)2(IPr)](Br)2 ( 3 ) were generated, which respectively afforded 4 and 5 as crystalline solids upon reduction with KC8. Experimental and computational studies support the semiquinoidal nature of 5 with a small singlet?triplet energy gap ΔES?T of 10.7 kcal mol?1, whereas 4 features more quinoidal character with a rather large ΔES?T of 25.6 kcal mol?1. In view of the low ΔES?T, 4 and 5 may be described as biradicaloids. Moreover, 5 has considerable (41 %) diradical character.  相似文献   

19.
The effect of polyoxyethylene type nonionic surfactants (C12E n n = 3, 4, 5, 6, 7 and 8) on the aqueous solution of sodium dodecyl sulfate (SDS) in absence and presence of NaCl was examined using small-angle neutron scattering (SANS), dynamic light scattering (DLS), and viscosity measurements. Upon addition of C12E n , micellar size of SDS was found to increase significantly, and such micellar elongation was further enhanced in the presence of NaCl. Micellar growth is most significant in presence of shorter moieties of C12E n (e.g., n = 3, 4) as compared to higher ethereal oxygen content. The results of structural investigations with SANS and DLS to confirm this assumption are reported. The cloud point of C12E n has increased upon addition of SDS and decrease with NaCl, and a typical behavior is observed when both SDS and NaCl were present.  相似文献   

20.
The cloud point curves of polyethoxylated surfactants are established experimentally. These experimental data are preliminary to the development of the cloud point extraction process, which appears as an interesting alternative to the usual solvent extraction unit operation. Starting from the thermodynamic model developed by Flory and Huggins for phase separation of polymer aqueous solutions, this paper aims at the prediction of cloud point curves. In this work, Rupert’s approach is extended to commercial nonionic surfactants, mixtures of homologous species, namely a few alkylphenol and alcohol ethoxylates. The limit of such an approach is clearly demonstrated, provided that a fitting parameter is finally required for a successful model application to pilot-plant manufactured surfactants, like C8ΦEn (n?=?7.5, 10, 12), C9ΦEn (n?=?8, 10, 12), C12E4, C12E6 and commercial Tergitol 15-S-7 (linear C12-C14 secondary alcohol with an average of 7 ethylene oxide units).GRAPHICAL ABSTRACT  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