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.
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   

2.
《Analytical letters》2012,45(11):2091-2105
Abstract

The quenching of the fluorescence intensity of pyrene by KBr has been measured in premicellar solutions of cetyltrimethylammonium bromide (CTAB), sodium lauryl sulfate (SLS) and polyoxyethylene 23-lauryl ether (Brij 35). The association of pyrene with premicellar aggregates is thought to bring the bromide ion closer to the fluorophore and hence results in a greater quenching effect. In 1.0 × 10?4 M CTAB solutions there is the beginning of “protective” premicellar aggregation. At this stage, 13 times more bromide ion is needed to effect the same degree of quenching as in pure water.  相似文献   

3.
Solutions of Mn(THF)2Br2 and MnI2 and tertiary phosphines in tetrahydrofuran at 0°C are oxidised by dioxygen or electrochemically giving deep blue-purple solutions which have identical electronic spectra. The principal band with λmax at 570 nm has ?>9000 1 cm?1 mol?1. Tertiary phosphines have been shown to be oxidised to the corresponding tertiary phosphine oxides. No evidence was found for the reversible formation of dioxygen-manganese complexes.  相似文献   

4.
Freshly prepared solutions of biliverdin dimethyl ester ( 2 ) in ethanol showed fluorescence maxima at 710 and 770 nm [ΦF = 1.1. 10?4 (room temperature) and 5.0 10?4 (77 K)]. The maxima of monoprotonated 2 at 77 K were shifted to 725 and 806 nm and the quantum yield was increased to 2.6. 10?2. This acid effect was reversible by neutralization with base. When a neutral solution was kept standing in the dark at room temperature, or when an acidic solution was neutralized by base, an additional fluorescence maximum at 500 nm with a mirror image excitation spectrum with λmax = 470 nm developed, which disappeared on addition of acid and which is attributed to a chemical change of 2 .  相似文献   

5.
The effect of the addition of 2-methoxyethanol on the critical micelle concentration (cmc) and on the degree of counterion dissociation (??) of butanediyl-1,4-bis(tetradecyldimethylammonium bromide) gemini surfactant, [C14H29N+(CH3)2?C(CH2)4?CN+(CH3)2C14H29,2Br?] (referred as 14?C4?C14,2Br?), has been studied by varying the compositions of the 2-methoxyethanol + water mixed solvent media (0 to 50?%). To determine various thermodynamic parameters of micellization, on the basis of the mass?Caction model for micelle formation, the experiments were performed at selected compositions of the mixed solvent at four temperatures ranging between 25?°C and 50?°C. Furthermore, the air/bulk surface tensions of the pure and mixed media were determined, and a successful attempt was made to correlate the cohesive energy density described through the Gordon parameter with the values of Gibbs energy of micellization.  相似文献   

6.
A new class of surfactant–cobalt(III) complexes of the type trans-[Co(DH)2(OA)X], where DH = dimethylglyoxime, OA = octadecylamine, X = Cl?, Br?, I?, N3 ?, NO2 ?, SCN? or OA, were synthesized and characterized by physicochemical and spectroscopic methods. The critical micelle concentration (CMC) values of these surfactant–cobalt(III) complexes in ethanol solution were obtained by measuring absorption at ~250 nm. Specific conductivity data (at 303–313 K) served for the evaluation of the temperature-dependent CMC and the thermodynamics of micellization (ΔG m 0 , ΔH m 0 and ΔS m 0 ). Steady-state photolysis and cyclic voltammetry of the complexes were studied. The surfactant–cobalt(III) complexes were screened for their antibacterial and antifungal activities against various microorganisms.  相似文献   

7.
《Analytical letters》2012,45(8):1407-1412
Abstract

A spectrophotometric method was developed to determine nitrite using safranin as color reagent. The reaction between nitrite and safranin produces a safranin-HNO2 species, which exhibits absorption peaks at 280, 349, 420(shoulder) and 610 nm. The peak at 610 nm was chosen as the analysis wavelength because nitrite ion and safranin do not present absorption bands in this region. The Lambert-Beer law was obeyed in the concentration range 7.0 × 10?6 - 5.0 × 10?5M. The effects of various ions on absorbance of the safranin-HNO2 species were studied; the nitrite analysis can be performed without interference in the presence of the ions SCN?, Br?, CH3COO?, Cl? (≤ 1.0 × 10?3 M) and NO3 ? (< 1.0 × 10?5 M). The SO4 = does not interfere even at a concentration of 0.25M.  相似文献   

8.
In this study, we have investigated the micellization characteristics of n-cetylpyridinium bromide (CPB) and lysozyme–CPB system using isothermal titration calorimetry (ITC) technique. The ITC operates in a stepwise addition mode, providing an excellent method of determination of critical micelle concentration (CMC) and enthalpy of demicellization (and hence micellization). The micellization characteristics of CPB have been investigated by microcalorimetric technique at 25, 30, 35, and 40 °C in a buffer solution of Tris(hydroxymethyl)aminomethane and pH of 7.3. A scheme to describe lysozyme–CPB interaction has also been proposed on the basis of ITC technique in a buffer solution of Tris(hydroxymethyl)aminomethane at 30 °C, pH of 7.3, 0 mM NaCl, and 1 mM NaCl. The enthalpy changes associated with micelle dissociation were temperature and lysozyme concentration dependent, indicating the importance of hydrophobic interactions. The ΔG mic was found to be negative, implying that, micellization, as expected, occurs spontaneously once the CMC has been reached. The values of ΔG mic were found to become more negative with increasing temperature (in case of micellization of CPB) and with increasing the lysozyme concentration (in the case of lysozyme–CPB). The ΔS mic was also found to decrease with increasing temperature in both cases. The presence of NaCl (1 mM) in the solutions decreased the CMC of CPB.  相似文献   

9.
Nanosecond spectroscopic and kinetic studies of 4-nitronaphthylamine (4-NO2NA) in aerated and deaerated nonpolar solvents at room temperature show a transient species with absorption maxima at 470 and 665 nm. The rate constant for the decay of this species in deaerated benzene is 6.7 × 105 sec?1, while in aerated benzene solutions the species is quenched by oxygen with arate constant k = 2.0 × 109M?1·sec?1. The transient absorption at 470and 665 nm is assigned to the lowest triplet excited state of 4-NO2NA. In polar solvents, however, electronic excitation of 4-NO2NA does not lead to any detectable transient absorption between 400 and 800 nm for the temperature range of 25 to ?150°C. This is attributed to lack of intersystem crossing of 4-NO2NA in polar solvents.  相似文献   

10.
《Analytical letters》2012,45(17):2105-2126
Abstract

Native low-temperature phosphorescence of mebendazole and flubendazole in ethanol is used for the determination of these imidazoles in anthelmintic preparations with wavelength maxima and detection limits of λEXC = 322 nm, λEM = 454 nm; 10 ng ml?1 and λEXC = 325 nm, = λEM = 455 nm; 5 ng ml?1, respectively, with linear response up to 8 μg ml?1 and 9 μg ml?1, respectively. The structural basis of these phenomena is discussed for both compounds and for related imidazoles and benzimidazoles. Apart from good sensitivity and excellent specificity offered by the technique, the use of cryogenic equipment (liquid nitrogen, special cuvettes, expensive dewar cells) implies some disadvantages for routine analyses.  相似文献   

11.
The spectroscopic and kinetic data of the short lived intermediates obtained by the attack of H-radicals on fluoro-, chloro-, bromobenzene, benzylchloride and phenethylchloride in aqueous solutions were studied by pulse radiolysis technique. The first three yield cyclohexadienylradicals (k=1–1.5×109 dm3 mol?1 s?1) with characteristic absorption maxima in the region 220–330 nm. In the case of benzylchloride a quantitative abstraction of chlorine by the H-atoms is observed (k=9.5×108 dm3 mol?1 s?1) leading to the formation of the benzylradical (λmax=257, 303, 317.5nm). The attack of H-atoms on phenethylchloride can occur on the aromatic ring forming also a cyclohexadienylradical (k=2.0×109 dm3 mol?1 s?1, λmax=317, 323nm) as well as on the side chain (k=1.5×108 dm3 mol?1 s?1) yielding H2. The intermediates decay according to a second order reaction withk=2 to 4.6×109 dm3 mol?1 s?1. To elucidate reaction mechanisms, steady state radiolysis experiments on the same systems were performed.  相似文献   

12.
The concentration dependences of the Fourier transform infrared spectra of aqueous solutions of sodiumn-pentanesulfonate,n-hexanesulfonate,n-heptanesulfonate, andn-octanesulfonate were studied in concentration ranges encompassing the critical micellization concentrations (c.m.c). Changes in wavenumber, bandwidth, and intensity of infrared absorption bands were used to monitor the changes in molecular association with concentration. The premicellar aggregation below the c.m.c. may accompany the association of the alkyl chain, but not the counter-ion binding with the SO 3 group. Above the c.m.c, the front location of the counter-ion against the SO 3 group at the micelle surface is demonstrated.  相似文献   

13.
Photolyses of rhodium(III) complexes, Rh(NH3)5X2+ (X = Cl? and Br?), under intense magnetic fields, e.g. λexcit = 360 nm and H = 24 kG, have been investigated. The magnetic field quenches the photoaquation of the ammonia and enhances the photoaquation of the acido ligand, X = Cl? or Br? by 10% for H = 24 kG. The implication of two different precursors in the formation of the photoproducts is discussed.  相似文献   

14.
Abstract— Excitation by a Q-switched giant ruby laser (1.2 J output at 694 nm ?50 ns flash) of 2–10 µM solutions of methylene blue in water, 30% ethanol in water or 50 v/v% water-CH3CN at pH values in the range 2.0–9.3 converted the dye essentially completely to its T1 state. The absorption spectrum of T1 dye was measured in different media at pH 2.0 and 8.2 by kinetic spectrophotometry. Previously reported T-T absorption in the violet in acidic and alkaline solutions and in the near infrared in alkaline solution was confirmed. Values found for these absorptions in the present work with 30% ethanol in water as solvent are λmax - 370nm, εmax, - 13,200 M-1 cm-1 at pH 2 and λmax,?420nm, εmax 9000 M-1 cm-1, λmax, - 840 nm, εmax - 20,000 m -1 cm-1 at pH 8.2. Long-wavelength T-T absorption in acidic solution is reported here for the first time: λmax, ? 680 nm, emax? 19,000 M cm-1 in 30% ethanol in water at pH 2. Observation of a pH-independent isobestic point ? 720 nm confirms that the long-wavelength absorptions are due to different protonated states of the same species, MB+(T1) and MBH2+(T1). The pKa of MBH2+(T1) in water was determined from the dependence on pH of absorption at 700 and 825 nm to be 7.14± 0.1 and from the kinetics of decay of triplet absorption to be 7.2. The specific rate of protonation of MB+(T1) by H2PO4 in water at pH 4.4 was found to be 4.5 ± 0.4 times 108M-1s-1.  相似文献   

15.
High‐regioregular poly{3‐[6‐(1‐methylimidazolium‐3‐yl)hexyl]thiophene‐2,5‐diyl bromide}, PMHT‐Br, has been prepared by reaction of high‐regioregular (above 92%) poly[3‐(6‐bromohexyl)thiophene‐2,5‐diyl] with 1‐methylimidazole. PMHT‐Br is soluble in water and water miscible solvents such as methanol, DMSO and shows solvatochromism; λmax (nm): 423 (H2O); 435 (MeOH); 452 (DMSO). Increased absorption band broadening observed for aqueous solution as well as NMR spectra in D2O suggests a micelle‐like structure of PMHT‐Br molecules in these solutions: poly(3‐hexylthiophene) core and 1‐methylimidazolium bromide shell. Despite the disturbing effect of ionic groups, the solid‐state PMHT‐Br shows absorption maximum at 520 nm, the band edge at 660 nm (ca. 1.9 eV), and fluorescence emission with maximum at 635 nm, in a good agreement with the polymer regioregularity. Fluorescence emission maxima: λem (nm): 598 (H2O); 562 (MeOH); 574 (DMSO), occur in a vicinity of corresponding adsorption band edges. Plot of electrical conductivity of PMHT‐Br (measured under the dynamic vacuum conditions, 5 × 10?5 Pa) versus 1/T shows a break at about 70 °C same as the temperature dependence of λmax of the solid PMHT‐Br. These breaks indicate an increase in the mobility of polymer segments and ions within PMHT‐Br; however, a thermal analysis did not provide solid evidence for it. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 3073–3081, 2010  相似文献   

16.
The gas phase photodissociation spectra of four protonated β-diketones were obtained and compared with the absorption spectra of the corresponding ions in solution. Protonated 2,4-pentanedione was observed to undergo the photodissociation process [C5H9O2]+ +hν → [CH3CO]+ +C3H6O with a λmax at 276±10 nm compared with a solution absorption maximum at 286 nm. Protonated 2,4-hexanedione was observed to undergo the photodissociation processes [C6H11O2]+ +hν → [CH3CO]+ +C4H8O and [C6H11O2]+ +hν → [C2H5CO]+ +C3H6O with a λmax at 279±10 nm compared with a solution absorption maximum at 288 nm. Protonated 3-methyl-2,4-pentanedione was observed to undergo the photodissociation process [C6H11O2]+ +hν → [CH3CO]+ +C4H8O with a λmax at 295±10 nm compared with a solution absorption maximum at 305 nm. Protonated 1,1,1-trifluoro-2,4-pentanedione was observed to undergo the photodissociation process [C5H6F3O2]+ +hν → CF3H+[C4H5O2]+ with a λmax at 273±10 nm compared with a solution absorption maximum at 288 nm. The [CH3CO]+ and [C2H5CO]+ produced photochemically with the first three ions react to regenerate the protonated β-diketone leading to a photostationary state. Photodissociation of the protonated alkyl β-diketones is believed to occur from the protonated keto form, whereas photodissociation of protonated 1,1,1-trifluoro-2,4-pentanedione is believed to occur from the protonated enol form. Mechanisms for the observed photodissociation processes are proposed and comparisons with results from related techniques are presented.  相似文献   

17.
Abstract— Semimethylene blue was generated by reductive quenching of triplet methylene blue, 3MBH2+, with diphenylamine at pH 0.62–3.4. A Q-switched ruby laser flash-photolysis-kinetic spectro-photometric apparatus was used to characterize the absorption spectrum of semimethylene blue from 350 to 900 nm and a number of physical constants at 25°C with μ= 0.4 M and Cl? as the anion. The specific rate of quenching of 3MBH2+ by DPA is 2.8 × 109M?1 s?1 in 5% EtOH-95% water and 1.2×109M?1 s?1 in 50 v/v% aq. CH3CN. Corresponding efficiencies of net electron transfer are, respectively, 0.15 and 0.62. Spectral characteristics in 5% EtOH are, for MBH22±, λmax= 375 nm, ε375= 9000 M?1 cm?1; λmax= 880 nm, ε880= 12700 M?1 cm?1; for MBH±, λmax= 410 nm, ε410= 9800 M?1 cm?1, λmax= 880 nm, ε880= 33000 M?1 cm?1; for MBH± in 50 v/v% AN, λmax= 400 nm, ε400= 11000 M?1 cm?1 and λmax= 880 nm,ε880= 39000 M?1 cm?1. The pKa of MBH22ε calculated from the pH dependence of the absorption spectrum is 1.86 × 0.04 in 5% EtOH and 1.15 in 50 v/v% AN. Rate constants, kdecay, for reaction DPAH±+ with MBH22ε and MBH± in 5% EtOH are, respectively, 3.9 × 109 and 9.5 × 109M?1 s?1. The value of pKa of MBH22ε calculated from the dependence of kdecay on pH is 1.75 in 5% EtOH.  相似文献   

18.
The kinetics of dissociation of bis(2,4,6–tripyridyl-s-triazine) iron(II), ([Fe(TPTZ)2]2+) has been studied in CTAB/chloroform/hexane reverse micellar medium. In the absence of acid, the reaction is immeasurably slow and does not go to completion in conventional aqueous medium but is markedly accelerated and takes place with a rate constant equal to 55.3 × 10?3 s?1 and goes to completion in reverse micelles. The significant increase in rate is attributed to the special properties of the water pool in the reverse micelles like low dielectric constant, nucleophilic effect of Br- ion, and favorable partitioning of TPTZ in the organic phase. The rate of the reaction decreases with increase in W (=[H2O]/[CTAB]) at constant CTAB concentration and remains constant with increase in CTAB at fixed W. The results are compared with other closely related systems.  相似文献   

19.
Apparent molar volumes (? v ), apparent molar adiabatic compressions (? κ) and thermodynamic parameters of cetyltrimethyl ammonium bromide (CTAB) in 0.001, 0.01, 0.05 and 0.1 mol·dm?3 aqueous solution of glycine in the temperature range 25–40 °C (at an interval of 5 °C) have been determined from density, speed of sound and conductometric measurements, respectively. The above calculated parameters were found to be sensitive to the interactions prevailing in the glycine–CTAB–water system. The analysis of the data was found to suggest that the ? v and ? κ values decrease sharply in the pre-micellar region and thereafter the decrease is almost linear at all concentrations of glycine, showing the dominance of hydrophobic interactions and facilitating the process of micellization. The ? v and ? κ values of these surfactants are found to be completely consistent with temperature over the entire concentration range. From conductance studies, the value of critical micelle concentration has been calculated, which shows dependence on the concentration of glycine as well as on temperature. A reasonably good qualitative correlation is found to exist with regard to CTAB–glycine interactions obtained from the conductance measurements and those from density and sound velocity measurements. All these observations demonstrate that this amino acid–surfactant system behaves in a different manner as compared to amino acid–electrolyte systems.  相似文献   

20.
Quenching of Ru(bpy)32+ electrochemiluminescence (ECL) by Cl?, Br?, and I? ions was studied as a function of halide concentration in a bipolar electrochemical cell. All of the halides investigated showed similar qualitative behavior: above a critical concentration, ECL intensity was found to decrease linearly as the halide ion concentration was increased, due to dynamic quenching of Ru(bpy)32+ ECL. Stern‐Volmer slopes (KSV) of 0.111±0.003, 4.2±0.3, and 6.2±0.3 mM?1 were measured for Cl?, Br? and I?, respectively. The magnitude of KSV correlates with halide ion oxidation potential, consistent with an electron transfer quenching mechanism. Using the bipolar platform described herein, aqueous, halide‐containing solutions could be quantified rapidly using the sequential standard addition method. The lower detection limit is determined by a complex mechanism involving the competitive electrooxidation of halide ions and the ECL co‐reactants, as well as the passivation of the surface of the bipolar electrode, and was found to be 0.20±0.01, 0.08±0.01 and 10±1 mM, respectively, for I?, Br?, and Cl?. The performance of the bipolar ECL quenching assay is comparable to previously published fluorescence quenching methods for the determination of halide ions, while being much simpler and less expensive to implement.  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