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 15 毫秒
1.
The behavior of the methyl radical adduct of six β‐phosphorylated nitrones in the N‐benzylidene‐1‐diethoxyphosphoryl‐1‐methylethylamine N‐oxide series in the presence of sodium dodecyl sulfate (SDS) micelles was followed by electron paramagnetic resonance spectroscopy. Except when the highly hydrophilic trap 4‐PyOPN (2) was used, all the adducts were found to partition significantly between micelles and the bulk aqueous phase. The average correlation time τ of the exchange of spin adducts between SDS micelles and water was found to be in the range 5 × 10?8—4 × 10?7 s, which is in the region of the life time of an SDS monomer in the micelle structure. In each case, the adduct affinity for the micelles has been quantified by evaluating its micelle–water distribution coefficient Kd. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

2.
The partitioning behavior of four newly synthesized chalcones between aqueous and micellar phases of ionic surfactants (SDS and CTAB) was investigated using ultraviolet-visible spectroscopy. The simple absorption spectra were recorded to study the interaction between these drugs and surfactants (in the concentration range below critical micelle concentration to above critical micelle concentration). The absorption data is also used to determine the number of additive molecules incorporated per micelle of the surfactant. The partition coefficient (Kx) of additives between bulk water phase and the micellar phase was determined in the range of 5.52 × 10+4 to 5.06 × 10+5 at 298 K by differential spectroscopic method. The corresponding standard free energy of partition ΔG°p obtained was in the range of ?27.05 kJmol?1 to ?32.54 kJmol?1. The relative solubility of additives between aqueous and micellar phases in different micellar concentrations was also estimated. The results showed that the chalcones are preferably soluble in cationic surfactant micelles.  相似文献   

3.
Effect of anionic surfactant on the optical absorption spectra and redox reaction of basic fuchsin, a cationic dye, has been studied. Increase in the absorbance of the dye band at 546 nm with sodium dodecyl sulfate (SDS) is assigned to the incorporation of the dye in the surfactant micelles with critical micellar concentration (CMC) of 7.3 × 10?3 mol dm?3. At low surfactant concentration (<5 × 10?3 mol dm?3) decrease in the absorbance of the dye band at 546 nm is attributed to the formation of a dye–surfactant complex (1:1). The environment, in terms of dielectric constant, experienced by basic fuchsin inside the surfactant micelles has been estimated. The association constant (KA) for the formation of dye–SDS complex and the binding constant (KB) for the micellization of dye are determined. Stopped‐flow studies, in the premicellar region, indicated simultaneous depletion of dye absorption and formation of new band at 490 nm with a distinct isosbestic point at 520 nm and the rate constant for this region increased with increasing SDS concentration. The reaction of hydrated electron with the dye and the decay of the semireduced dye are observed to be slowed down in the presence of SDS. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 629–636, 2003  相似文献   

4.
The formation of micelles of hexadecyltrimethylammonium chloride (CTAC) and sodium dodecylsulfate (SDS) in aqueous solutions containing charged polysaccharides was studied by steady-state and time-resolved fluorescence measurements using pyrene as a photophysical probe. Micropolarity studies using the I1/I3 ratio of the vibronic emission bands of pyrene and the behaviour of the IE/IM ratio between the excimer and monomer emissions show the formation of hydrophobic domains. The interactions between the polyelectrolytes and surfactants of opposite charge lead to the formation of induced pre-micelles at surfactant concentrations lower than the critical micellar concentration (cmc) of the surfactants. At similar concentrations, the IE/IM ratio shows a peak. This aggregation process is assumed to be due to electrostatic attractions. At higher surfactant concentrations, near the critical micellar concentration, micelles with the same properties as those found in pure aqueous solution are formed. On the other hand, systems containing polyelectrolytes and surfactants of the same charge do not show this behaviour at low concentrations. The presence of long alkyl chains bound to the polyelectrolytes also induces the formation of free micelles at concentrations somewhat below the aqueous cmc.  相似文献   

5.
《Analytical letters》2012,45(1):151-162
Abstract

A novel developed spectrofluorimetric method for the determination of trazodone hydrochloride in the presence of sodium dodecyl sulfate (SDS) surfactant micelles was described. Under optimal conditions, there was a good linear relationship between fluorescence intensity and trazodone hydrochloride concentration in the range of 4.0×10?9 to 8.0×10?6 mol · l?1with the detection limit of 1.3×10?9 mol · l?1 (S/N=3). This method has been used to determine trazodone hydrochloride in biological fluids.  相似文献   

6.
The onset of micelles formations critical micelle concentration, diffusion coefficients as well as particle sizes for some new synthesized anionic copolymer surfactants PSA4a, PSA4b, and PSA4c, were determined and discussed. Three different electrochemical techniques such as cyclic voltammetry (CV), rotating disk voltammetry (RDV), and chronocoulometry (CC) were used in this investigation. The voltammetry of electroactive hydrophobic probe ferrocene solubilized surfactants was investigated in aqueous buffer carbonate solutions of pH 10. The CMC for each PSA4a, PSA4b, and PSA4c, was found to be 3.20 × 10?4, 4.60 × 10?4 and 6.30 × 10?4 M, respectively, using both CV and RDV techniques. The amount of adsorption contribution of ferrocene solubilized surfactants at the glassy carbon electrode was determined from chronocoulometric measurements and it was found in the range from (1.4 to 2.7) × 10?15 M. The apparent diffusion coefficients were estimated from RDV measurements and the real micelles diffusion coefficients were obtained. Re-quilibrium considerations of ferrocene probe kinetics at the electrode surface were treated according to two different modes of slow- and fast-kinetics. The corrected diffusion coefficient values showed constancy at (5.3 ± 0.1) × 10?7, (4.8 ± 0.1) × 10?7, and (3.6 ± 0.4) × 10?7 cm2/sec for PSA4a, PSA4b, and PSA4c, respectively in the concentration range from 20 to 200 mM. The morphological features of anionic copolymeric surfactants PSA4a, PSA4b, and PSA4c, micelles showed globular self-assembled structure.  相似文献   

7.
A combined flash photolysis and pulse radiolysis experiment was carried out to produce triplet pyrene (P) molecules in micelles of cetyltrimethylammonium bromide and Br2? in the surrounding aqueous medium. The reaction 3Pmic + Br → P + 2 Br? was followed by optical absorption measurements in the 10?8?10?4–sec range. This reaction possesses a “fast” and a “slow” component with respective rate constants of 2.3 × 106 sec?1 and 1 × 109M?1 · sec?1. The fast component is related to the probability of a Br2? radical meeting a triplet pyrene containing micelle on the first encounter (only 16% of the micelles contained a triplet molecule). Reactions involving more than one Br2? radical–micelle encounter are ascribed to the slow component. The presence of two components reflects the fact that the residence time of a Br2? radical in the vicinity of a cationic micelle is substantially longer than the diffusion time of the radical between micelles. Thus the conditions met in micellar chemistry differ dramatically from those in ordinary solution kinetics where the encounter time is generally much shorter than the time between encounters. Some considerations on the energetics of this electron transfer reaction are also presented.  相似文献   

8.
The behavior of the triphenylmethane dye crystal violet in aqueous solutions containing polyoxyethylene nonionic surfactants was investigated using absorption and fluorescence spectroscopic techniques. The interactions of the dye were examined in micellar media in order to prevent dye aggregation and to ensure maximum dye and surfactant interaction. The relative fluorescence enhancements and the binding constants of the dye to the surfactant micelles were determined. The micropolarities of the micellar environment sensed by the pyrene probe were estimated from the I 1/I 3 intensity ratios of the fluorescence spectra of pyrene. The fluorescence quenching of pyrene by hexadecylpyridinium chloride was investigated in aqueous surfactant mixtures at a fixed concentration of surfactant in order to determine the aggregation numbers. Attempts were made to correlate the binding constants obtained in this investigation to various micellar parameters.  相似文献   

9.
The binding of mixed surfactants of cationic cetyltrimethylammonium bromide (CTAB) and nonionic octaethylene glycol monododecyl ether (C 12E 8) on anionic polyelectrolyte poly[2-acrylamido-2-methylpropanesulfonic acid (PAMPS)] and fluorophore-labeled copolymers containing about 40 mol% of AMPS was investigated at different mole fractions, Y , of CTAB in the surfactant mixture. The excimer emission of the cationic probe 1-pyrenemethylamine hydrochloride (PyMeA·HCl), nonradiative energy transfer (NRET) between pyrene and naphthalene labels and I 1/ I 3 of the pyrene label were determined by varying the total surfactant concentration, c Surf. The I E/ I M value of PyMeA·HCl firstly increases and then decreases to 0 with c Surf, showing a maximum on every curve. The critical aggregation concentration of the mixed surfactants determined from the I E/ I M maximum decreased from 5×10 -5 to 1×10 -5 mol/l as Y increased from 0.1 to 0.50, and then leveled off as Y increased up to unity. And at least 5×10 -6 mol/l CTAB was required for the mixed surfactants to bind on the PAMPS cooperatively. Equimolar binding of CTAB on AMPS was formed at I E/ I M=0 when Y =0.25, while at Y =0.1 some CTAB molecules in the mixed micelle were directed to the water phase without binding with AMPS. Both the intramolecular and the intermolecular NRET increased and then decreased with c Surf, having a maximum on each curve corresponding to the equimolar binding of CTAB and AMPS so long as Y >0, indicating the coiling of the chain and interchain aggregation upon bound surfactants. The I Py/ I Np value at the maximum decreased with decreasing Y because more nonionic surfactant C 12E 8 participated into the polyelectrolyte-mixed surfactant complexes together with bound CTAB.  相似文献   

10.
Zwitterionic diazeniumdiolates of the form RN[N(O)NO?](CH2)2NH2+R, where R=CH3 ( 1 ), (CH2)3CH3 ( 2 ), (CH2)5CH3 ( 3 ), and (CH2)7CH3 ( 4 ) were synthesized by reaction of the corresponding diamines with nitric oxide. Spectrophotometrically determined pKa(O) values, attributed to protonation at the terminal oxygen of the diazeniumdiolate group, show shifts to higher values in dependence of the chain lengths of R. The pH dependence of the decomposition of NO donors 1 – 3 was studied in buffered solution between pH 5 and 8 at 22 °C, from which pKa(N) values for protonation at the amino nitrogen, leading to release of NO, were estimated. It is shown that the decomposition of these diazeniumdiolates is markedly catalyzed by anionic SDS micelles. First‐order rate constants for the decay of 1 – 4 were determined in phosphate buffer pH 7.4 at 22 °C as a function of SDS concentration. Micellar binding constants, KSM, for the association of diazeniumdiolates 1 – 3 with the SDS micelles were also determined, again showing a significant increase with increasing length of the alkyl side chains. The decomposition of 1 – 3 in micellar solution is quantitatively described by using the pseudo‐phase ion‐exchange (PIE) model, in which the degree of micellar catalysis is taken into account through the ratio of the second‐order rate constants (k2m/k2w) for decay in the micelles and in the bulk aqueous phase. The decay kinetics of 1 – 3 were further studied in the presence of cosolvents and nonionic surfactants, but no effect on the rate of NO release was observed. The kinetic data are discussed in terms of association to the micelle–aqueous phase interface of the negatively charged micelles. The apparent interfacial pH value of SDS micelles was evaluated from comparison of the pH dependence of the first‐order decay rate constants of 2 and 3 in neat buffer and the rate data obtained for the surfactant‐mediated decay. For a bulk phase of pH 7.4, an interfacial pH of 5.7–5.8 was determined, consistent with the distribution of H+ in the vicinity of the negatively charged micelles. The data demonstrate the utility of 2 and 3 as probes for the determination of the apparent pH value in the Stern region of anionic micelles.  相似文献   

11.
Abstract

Interaction between dye (ECAB), nonionic surfactant (TX-100) micelle in aqueous solution and TX-100 hemimicelie at solid (SiO2)/liquid interface has been investigated quantitatively. There are linear relationships between concentrations of free ECAB(Ca), ECAB bound with TX-100 micelles in solution(Cm) and ECAB bound with TX-100 hemimicelles at interface of solid/liquid(Chm). The slopes of the three straight lines are 0.32 for Chm~Ca -1.32 for Cm~Ca and -1.00 for (Cm+Chm~Ca respectively. The linear relationships can be described by three linear equations as follows: Chm=0.32 (Ca?O.88×10?5),Cm.=4.0×10?5-l.33 Ca and Chm+Cm=3.742×l0?5-Ca,. It is inferred that the interaction between ECAB, TX-100 micelles and TX-100 hemimicelles is essentially partition of ECAB molecules in solution, TX-100 micelles and hemimicelles. The concentration of ECAB bound with TX-100 micelles well as electronic repulsion. Additionally, A quantitative method to determine adsorbance of surfactant TX-100 on silica gel by spectroscopy in coadsorption conditions of dye (ECAB) and TX-100 was proposed.  相似文献   

12.
New methylene blue (NMB) dye incorporated into AlMCM‐41 surfactant‐free and hybrid surfactant‐AlMCM‐41 mesophase. UV‐vis evidence shows that new methylene blue dye protonated in both cases of zeolites. New methylene blue is electroactive in zeolites and their electrochemical activity has been studied by cyclic voltammetry and compared to that of NMB in aqueous solutions. New methylene blue molecules are not released to the solution during CV measurements and are accessible to H3O+ ions. The presence of surfactant affects the kinetics of the redox process through proton ions diffusion. The midpoint potentials (Em) values show that new methylene blue dye incorporated into AlMCM‐41 can be reduced easily with respect to solution new methylene blue. New methylene blue interacting with surfactant polar heads and residual Br? ions as a results, it shows a couple of peaks in high potential with respect to new methylene blue solution. The electrode made with methylene blue‐AlMCM‐41 without surfactant was used for the mediated oxidation of ascorbic acid. The anodic peak current observed in cyclic voltammetry was linearly dependent on the ascorbic acid concentration. The calibration plot was linear over the ascorbic acid concentration range 1.0×10?5 to 5.0×10?4 M. The detection limit of the method is 1.0×10?5 M, low enough for trace ascorbic acid determination in various real samples.  相似文献   

13.
The redox system of potassium persulfate–thiomalic acid (I1–I2) was used to initiate the polymerization of acrylamide (M) in aqueous medium. For 20–30% conversion the rate equation is where Rp is the rate of polymerization. Activation energy is 8.34 kcal deg?1 mole?1 in the investigated range of temperature 25–45°C. Mn is directly proportional to [M] and inversely to [I1]. The range of concentrations for which these observations hold at 35°C and pH 4.2 are [I1] = (1.0–3.0) × 10?3, [I2] = (3.0–7.5) × 10?3, and [M] = 5.0 × 10?2–3.0 × 10?1 mole/liter.  相似文献   

14.
Quasielastic neutron scattering measurements have been made on some micellar aggregates. It is shown that the observed spectra arise almost exclusively from monomer motions in the aggregates. Two illustrative systems were studied, direct micelles of tetradecyl trimethyl ammonium bromide in water, and reverse micelles of aerosol OT in cyclohexane with different amounts of water solubilised inside the micellar core. For the aerosol OT micelles, the translational diffusion is fast (8×10?10 m2.sec?1) and independent of the size of the water pool. Rotational correlation times are of the order of 4×10?11 sec. At high water contents the water in these pools is similar to bulk aqueous electrolytic solutions. For the tetradecyl trimethyl ammonium bromide micelles the translational diffusion constant was found to be 5×10?10 m2.sec?1.  相似文献   

15.
《Chemphyschem》2003,4(10):1065-1072
Dielectric spectra have been measured for aqueous sodium dodecylsulfate (SDS) solutions up to 0.1 mol L?1 at 25 °C over the frequency range 0.005≤ν GHz?1≤89. The spectra exhibit two relaxation processes at approximately 0.03 GHz and 0.2 GHz associated with the presence of micelles in addition to the dominant solvent relaxation process at approximately 18 GHz and a small contribution at approximately 1.8 GHz due to H2O molecules hydrating the micelles. Detailed analysis reveals that the micelles bind 20 water molecules per SDS unit, but not as strongly as trimethylalkylammonium halide surfactants do. The relaxation times and amplitudes of both micelle relaxation processes can be simultaneously analysed with the theory of Grosse, yielding the effective volume of a SDS unit in the micelle and the lateral diffusion coefficient of the bound counterions. The findings of this investigation fully corroborate recent molecular dynamics simulations on structure and dynamics of SDS micelles.  相似文献   

16.
Measurements of the thermal expansion coefficients (TECs) of cellulose crystals in the lateral direction are reported. Oriented films of highly crystalline cellulose Iβ and IIII were prepared and then investigated with X‐ray diffraction at specific temperatures from room temperature to 250 °C during the heating process. Cellulose Iβ underwent a transition into the high‐temperature phase with the temperature increasing above 220–230 °C; cellulose IIII was transformed into cellulose Iβ when the sample was heated above 200 °C. Therefore, the TECs of Iβ and IIII below 200 °C were measured. For cellulose Iβ, the TEC of the a axis increased linearly from room temperature at αa = 4.3 × 10?5 °C?1 to 200 °C at αa = 17.0 × 10?5 °C?1, but the TEC of the b axis was constant at αb = 0.5 × 10?5 °C?1. Like cellulose Iβ, cellulose IIII also showed an anisotropic thermal expansion in the lateral direction. The TECs of the a and b axes were αa = 7.6 × 10?5 °C?1 and αb = 0.8 × 10?5 °C?1. The anisotropic thermal expansion behaviors in the lateral direction for Iβ and IIII were closely related to the intermolecular hydrogen‐bonding systems. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 1095–1102, 2002  相似文献   

17.
In this work, the interaction between the anionic surfactant sodium dodecyl sulfate (SDS) and the polyelectrolyte complex hydrolyzed polyacrylamide/poly(4-vinylpyridine) (AD37–P4VP) in aqueous solution was investigated by conductometric measurements. Three main series with SDS concentrations of 0.01, 0.25 and 1 % and in a wide range of P4VP and AD37 concentrations, from 0.1 × 10?4 to 4 × 10?4 g/ml, and from 10?4 to 10?3 g/ml, respectively, were studied. The polyelectrolyte complex interacts strongly with the SDS surfactant. These interactions are of electrostatic and hydrophobic types. Thus, the effect of salt on the critical micelle concentration of SDS, and the neutralization degree on behavior conductivity of the mixture, were quantified.  相似文献   

18.
Thermodynamic properties of sodium dodecyl sulfate (SDS) in micellar aqueous solutions of L-serine and L-threonine were determined by fluorescence spectroscopy and dynamic light scattering techniques. The values of Gibbs free energy, enthalpy and entropy of the process of micelle formation were calculated using the critical micelle concentration and degree of dissociation. Changes in critical micelle concentration of SDS with the addition of amino acids were examined by both conductivity and pyrene I 1/I 3 ratio methods at different temperatures. The pyrene fluorescence spectra were used to study the change of micropolarity produced by the interaction of SDS with amino acids. The aggregation behavior of SDS was explained in terms of structural changes in mixed solutions. The data on dynamic light scattering suggest that size of SDS micelles was influenced by the presence of amino acids.  相似文献   

19.
The equilibrium constant for the reaction CH2(COOH)2 + I3? ? CHI(COOH)2 + 2I? + H+, measured spectrophotometrically at 25°C and ionic strength 1.00M (NaClO4), is (2.79 ± 0.48) × 10?4M2. Stopped-flow kinetic measurements at 25°C and ionic strength 1.00M with [H+] = (2.09-95.0) × 10?3M and [I?] = (1.23-26.1) × 10?3M indicate that the rate of the forward reaction is given by (k1[I2] + k3[I3?]) [HOOCCH2COO?] + (k2[I2] + k4[I3?]) [CH(COOH)2] + k5[H+] [I3?] [CH2(COOH)2]. The values of the rate constants k1-k5 are (1.21 ± 0.31) × 102, (2.41 ± 0.15) × 101, (1.16 ± 0.33) × 101, (8.7 ± 4.5) × 10?1M?1·sec?1, and (3.20 ± 0.56) × 101M?2·sec?1, respectively. The rate of enolization of malonic acid, measured by the bromine scavenging technique, is given by ken[CH2(COOH)2], with ken = 2.0 × 10?3 + 1.0 × 10?2 [CH2(COOH)2]. An intramolecular mechanism, featuring a six-member cyclic transition state, is postulated to account for the results on the enolization of malonic acid. The reactions of the enol, enolate ion, and protonated enol with iodine and/or triodide ion are proposed to account for the various rate terms.  相似文献   

20.
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