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 effect of cationic micelles of cetyltrimethyl ammonium bromide (CTAB) on the observed pseudo-first-order rate constant for the interaction of nickel dipeptide complex [Ni(II)-Gly-Gly]+ with ninhydrin has been studied spectrophotometrically. At constant temperature and pH, increase in the [CTAB] from 0.0 to 60.0 × 10?3 mol dm?3 caused nearly three-fold increase of the rate constant. The micellar catalysis is explained in terms of the pseudophase model. From the observed kinetic data, binding constants of micelle–[Ni(II)-Gly-Gly]+ (K S), and micelle–ninhydrin (K N) are evaluated, respectively, to be 5.3 mol?1 dm3 and 84.0 mol?1 dm3. The role of added inorganic (NaCl, NaBr, Na2SO4) and organic salts (NaBenz, NaSal) on the reaction rate has also been examined.  相似文献   

2.
Aggregation behavior of didodecyldimethylammonium bromide (DDAB) in aqueous solution was investigated using negative-staining and freeze-fracture transmission electron microscopy (TEM) methods. With the concentration increase, the vesicle size enlarged from the range of 100–200 nm to 500–3000 nm and the structure transform from unilamella to multilamella. Rheological and viscosity measurement results indicated that the system exhibited a gel-like material characteristic and shearing thinning in property, as shown that the apparent viscosity decreased gradually during the shear rate ascended from 1 × 10?4–1 × 104 s?1. Furthermore, the relationship between aggregate conformation and solution rheological properties was discussed.  相似文献   

3.
《Analytical letters》2012,45(10):1783-1795
The behavior of seven hydroxy anthraquinone derivatives was studied by capillary zone electrophoresis and micellar electrokinetic chromatography. The effects of buffer pH (6.5–10.8) and sodium dodecyl sulfate concentration (10–20 mmol L?1) on the effective mobilities of the analytes and their separation were tested. A comparison of the two optimized separation systems showed that micellar electrokinetic chromatography was superior as it permits separation of all the seven analytes within 15 min, using 15 mmol L?1 sodium dodecyl sulfate in 10 mmol L?1 tetraborate buffer, pH 8.5, at a voltage of 20 kV. The calibration curves were linear in the concentration range from 5.0 · 10?7 to 5.0 · 10?4 mol L?1 for most of the analytes, at a detection wavelength of 254 nm. LOD and LOQ values of the analytes were in the ranges of 2.10 · 10?7–1.28 · 10?6 mol L?1and 6.99 · 10?7–4.25 · 10?6 mol L?1, respectively. The proposed separation conditions were applied to determination of 1,2-dihydroxy anthraquinone (alizarin) and 1,2,4-trihydroxy anthraquinone (purpurin) in Rubia tinctorum aglycone and of the recently described 1-acetyl-2,4,5,7-tetrahydroxy-9,10-anthraquinone and 1-acetyl-2,4,5,7,8-pentahydroxy-9,10-anthraquinone in the mycelium of fungi Geosmithia lavendula.  相似文献   

4.
The interaction of ginkgolic acid (15:1, GA) with human serum albumin (HSA) was investigated by FT–IR, CD and fluorescence spectroscopic methods as well as molecular modeling. FT–IR and CD spectroscopic showed that complexation with the drug alters the protein’s conformation by a major reduction of α-helix from 54 % (free HSA) to 46–31 % (drug–complex), inducing a partial protein destabilization. Fluorescence emission spectra demonstrated that the fluorescence quenching of HSA by GA was by a static quenching process with binding constants on the order of 105 L·mol?1. The thermodynamic parameters (ΔH = ?28.26 kJ·mol?1, ΔS = 11.55 J·mol?1·K?1) indicate that hydrophobic forces play a leading role in the formation of the GA–HSA complex. The ratio of GA and HSA in the complex is 1:1 and the binding distance between them was calculated as 2.2 nm based on the Förster theory, which indicates that the energy transfer from the tryptophan residue in HSA to GA occurs with high probability. On the other hand, molecular docking studies reveal that GA binds to Site II of HSA (sub-domain IIIA), and it also shows that several amino acids participate in drug–protein complexation, which is stabilized by H-bonding.  相似文献   

5.
Time-resolved EPR has been used to study the photoinitiated reactions of the excited triplet of duroquinone (3DQ*) quenched by the antioxidant vitamin C (VC) in homogeneous solutions of ethylene glycol/water (EG/H2O) and micelle solutions of aerosol OT (AOT) and sodium dodecyl sulphate (SDS). During the photolysis reactions of DQ and VC in homogeneous solutions of EG/H2O, 3DQ* abstracts hydrogen atoms from the solvent EG and the antioxidant VC. The rate constant for the quenching of 3DQ* by VC is 4.90 × 107 L mol?1 s?1, close to being diffusion-controlled. In AOT and SDS micelle solutions, rate constants for the quenching of 3DQ* by VC are 3.28 × 107 and 3.15 × 107 L mol?1 s?1, respectively. Lipid-soluble 3DQ* and water-soluble VC need to diffuse to the w/o interface to react, which reduces the reaction rate between 3DQ* and VC. The charge repulsion interaction between the anionic shell of AOT (SDS) micelles and the VC monoanions AsH? also slows the reaction.  相似文献   

6.
The reduction reaction of the Cu(II)–pitn complex (pitn = 1,3-di(pyridine-2-carboxaldimino)propane) by decamethylferrocene [Fe(Cp*)2] was examined in acetonitrile. The observed pseudo-first-order rate constants exhibited saturation kinetics with increasing excess amount of [Fe(Cp*)2]. Detailed analyses revealed that the reaction is controlled by a structural change prior to the electron transfer step, rather than a conventional bimolecular electron transfer process preceded by ion pair (encounter complex) formation. The rate constant for the structural change was estimated to be 275 ± 13 s?1 at 298 K (?H* = 33.3 ± 1.0 kJ·mol?1, ?S* = 86 ± 5 J·mol?1·K?1), which is the fastest among gated reactions involving CuN4 complexes. It was confirmed by EPR measurement and Conflex calculations that the dihedral angle between the two N–N planes is significantly large (40°) in solution whereas it is merely 17.14° in the crystal.  相似文献   

7.
[CrIII(LD)(Urd)(H2O)4](NO3)2?·?3H2O (LD?=?Levodopa; Urd?=?uridine) was prepared and characterized. The product of the oxidation reaction was examined using HPLC. Kinetics of the oxidation of [CrIII(LD)(Urd)(H2O)4]2+ with N-bromosuccinimide (NBS) in an aqueous solution was studied spectrophotometrically, with 1.0–5.0?×?10?4?mol?dm?3 complex, 0.5–5.0?×?10?2?mol?dm?3 NBS, 0.2–0.3?mol?dm?3 ionic strength (I), and 30–50°C. The reaction is first order with respect to [CrIII] and [NBS], decreases as pH increases in the range 5.46–6.54 and increases with the addition of sodium dodecyl sulfate (SDS, 0.0–1.0?×?10?3?mol?dm?3). Activation parameters including enthalpy, ΔH*, and entropy, ΔS*, were calculated. The experimental rate law is consistent with a mechanism in which the protonated species is more reactive than its conjugate base. It is assumed that the two-step one-electron transfer takes place via an inner-sphere mechanism. A mechanism for this reaction is proposed and supported by an excellent isokinetic relationship between ΔH* and ΔS* for some CrIII complexes. Formation of [CrIII(LD)(Urd)(H2O)4]2+ in vivo probably occurs with patients who administer the anti-Parkinson drug (Levodopa), since CrIII is a natural food element. This work provides an opportunity to identify the nature of such interactions in vivo similar to that in vitro.  相似文献   

8.
Concentration effect of silver nanoparticles (AgNPs) on the photophysical properties of ciprofloxacin (Cip) have been investigated using optical absorption and fluorescence techniques. When performed AgNPs solution was added to the Cip solution, metal-enhanced fluorescence intensity and a blue-shift of 20 nm in the maximum emission spectra of Cip has been observed. The enhanced intensity of this system is strongly dependent on the AgNPs concentration and largest at the 6.0 × 10?6 mol L?1. With increase of AgNPs concentration, quenching of fluorescence is observed. Stern–Volmer quenching constants have been calculated at four temperatures. The results show the quenching constants are directly correlated with temperature. It indicates the quenching mechanism is the dynamic quenching in nature rather than static quenching. From which we determined the activation energy for the quenching of Cip-AgNPs to be about 31.1 kJ mol?1. In addition, in the presence of optimum AgNPs concentration, a sensitive fluorimetric method for the determination of ciprofloxacin at the range 5.0 × 10?7–3.0 × 10?5 mol L?1 and the detection limit of 2 × 10?8 mol L?1 in solution is proposed.  相似文献   

9.

Electrochemical polymerization of azure B from sulfuric acid solution was carried out by using cyclic voltammetry. The electrolytic solution consisted of 5.0 mmol · dm?3 azure B and 0.3 mol · dm?3 H2SO4. The temperature for polymerization was controlled at 20°C. A blue film, i.e., poly(azure B) was formed on a platinum foil and had a electrochemical reversibility, stability and a fast charge transfer ability in the 0.5 mol · dm?3 Na2SO4 with pH ≤4.0 solution. The currents of both anodic and cathodic peaks are proportional to υ1/2 at the scan rate (υ) region of 25 and 600 mV · s?1 on the cyclic voltammograms. The conductivity of poly(azure B) is 2.8×10?6 S · cm?1 at 20°C. The UV‐visible spectrum and Raman spectrum of the polymer are different from those of the monomer. A possible polymerization mechanism of azure B was also proposed.  相似文献   

10.
The kinetics of the oxidation of malachite green (MG+) by Fe(III) were investigated spectrophotometrically by monitoring the absorbance change at 618 nm in aqueous and micellar media at a temperature range 20–40 °C; I = 0.10 mol dm?3 for [H+] range (2.50–15.00) × 10?4 mol dm?3. The rate of reaction increases with increasing [H+]. The reaction was carried out under pseudo-first-order conditions by taking the [Fe(III)] (>10-fold) the [MG+]. A mechanism of the reaction between malachite green and Fe(III) is proposed, and the rate equation derived from the mechanism was consistent with the experimental rate law as follows: Rate = (k 4 + K 1 k 5[H+]) [MG+][Fe(III)]. The effect of surfactants, such as cetyltrimethylammonium bromide (CTAB, a cationic surfactant) and sodium dodecylsulfate (SDS, an anionic surfactant), on the reaction rate has been studied. CTAB has no effect on the rate of reaction while SDS inhibits it. Also, the effect of ligands on the reaction rate has been investigated. It is proposed that electron transfer proceeds through an outer-sphere mechanism. The enthalpy and the entropy of the activation were calculated using the transition state theory equation.  相似文献   

11.
The reactions of OH* and eaq ? adducts of uracil and thymine with Cu(II) ions in aqueous solutions were followed by pulse radiolysis. The transient absorption spectra of the OH* adducts of uracil when followed in the presence of Cu(II) ions show growth in absorption at wavelengths 420 and 350 nm at 15 μs and 65 μs after the pulse respectively. Similar transient absorption spectra of thymine showed growth in absorption at wavelengths 390 and 320 nm at 38 μs and 65 μs after the pulse respectively. The rates of electron transfer from the OH* adducts of uracil and thymine to various Cu(II) compounds when monitored at 360 nm lie between 106 and 108 mol?1 dm3 s?1 this implies that the electron transfer process is not an efficient process. Low rate constants coupled with the spectral changes suggest formation of a radical copper adduct which decays by water insertion to give cis-glycols as the major product. The electron transfer from the electron adducts of uracil and thymine to various copper(II) compounds takes place more efficiently (rate constants of the order of 108 and 109 mol?1 dm3 s?1) compared with that from the OH* adducts. The t-butanol radicals formed on scavenging the OH* radicals also produce adducts with Cu(I) ions which are formed on oxidation of the electron adducts by Cu(II) ions. This adduct has absorption around 400 nm both in the case of uracil and thymine.  相似文献   

12.
The interaction of plumbagin (PLU) with human serum albumin (HSA) in physiological buffer (pH=7.4) was studied by fluorescence spectroscopy. Results obtained from analysis of the fluorescence spectra indicated that PLU has a strong ability to quench the intrinsic fluorescence of HSA through a static quenching procedure. Fluorescence quenching data revealed that the quenching constants (K) are 4.43×104, 3.26×104 and 1.69×104 L?mol?1 at 293, 303 and 313 K, respectively. The thermodynamic parameters ΔH° and ΔS° were calculated to be ?36.63 kJ?mol?1, and ?35.702 J?mol?1?K?1 respectively, which suggested that van der Waals interactions and hydrogen bonds play a major role in the interaction of PLU with HSA. The distance between donor (HSA) and acceptor (PLU) was calculated to be 3.76 nm based on Förster’s non-radiative energy transfer theory. The results of synchronous fluorescence spectra showed that binding of PLU to HSA can induce conformational changes in HSA.  相似文献   

13.
The rate constants for oxygen quenching in benzene solution of the triplet states of several organic compounds with relatively high triplet energies have been measured in laser photolysis and pulse radiolysis experiments. The previously observed trend for aromatic hydrocarbons where the quenching rate constants decrease from a limiting value of about one ninth of that expected for a diffusion controlled reaction to lower values for triplet states with increasing triplet energy was not observed for the triplet states of certain aromatic ketones and amines. The higher rate constants observed, e.g. oxygen quenching of triplet N-methyl indole has kQ = 1.4 × 1010 dm3 mol?1 s?1, are interpreted as being due to the presence of low lying triplet charge-transfer states which enhance the efficiency of quenching.  相似文献   

14.
《Analytical letters》2012,45(5-6):449-459
Based on the study of voltammetric behavior of carcinogenic 1-nitropyrene (1-NP), 1-aminopyrene (1-AP), and 1-hydroxypyrene (1-HP), optimum conditions have been found for the determination of these analytes by differential pulse voltammetry (DPV) at a boron-doped diamond film electrode. The optimum medium was methanol-Britton–Robinson buffer (BR buffer) pH 3.0 (70:30) for 1-NP and 1-AP, and methanol-BR buffer pH 5.0 (70:30) for 1-HP. Concentration dependences of the DPV response were measured in the range 1 · 10?6–1 · 10?4 mol dm?3 (R = ?0.9998) with the limit of detection (LOD) 3 · 10?7 mol dm?3 for 1-NP, 1 · 10?7–1 · 10?5 mol dm?3 (R = 0.9971) with LOD 6 · 10?8 mol dm?3 for 1-AP, and 1 · 10?7–1 · 10?5 mol dm?3 (R = 0.9934) with LOD 1 · 10?7 mol dm?3 for 1-HP. Simultaneous determination of 1-NP and 1-AP in a mixture was tested in the methanol-BR buffer pH 3.0 (70:30) medium as well. The content of 1-AP in the concentration range from 1 · 10?6 to 1 · 10?4 mol dm?3 had no effect on the sensitivity of the determination of 1-NP, and vice versa. Due to the close peak potentials of 1-AP and 1-HP, the direct determination of their mixture using voltammetric methods is impossible.  相似文献   

15.
A novel supramolecular sensor derived from calix[4]pyrrole system i.e. calix[4]pyrrole bearing aminoanthraquinone derivative (CAAQ) have been designed and synthesized. The complexation behavior of metal cations [Ag(I), Ba(II), Ca(II), Ni(II), Co(II), Fe(III), Hg(II), Cu(II), Cr(II), Pb(II), Zn(II), (1 × 10?4 M)] with CAAQ (1 × 10?6 M) was studied by spectrophotometry and spectrofluorometry. Metal ion like Pb(II) produces red shift in absorption spectra and quenching in emission spectra likelihood of strong complexation of Pb(II) ions with CAAQ. Fluorescence cell imaging also supports the complexation of Pb(II) ions with CAAQ. The binding constants, quantum yield, stoichiometry of complex, mechanism of quenching by Stern–Volmer equation and Density functional theory calculation have been determined.  相似文献   

16.
《Analytical letters》2012,45(13):1849-1861
The anti-cancer synthetic drug irinotecan (CPT-11) and its active metabolite SN-38 have been determined by micellar electrokinetic capillary chromatography (MEKC). The detection of the analytes was made at 368 nm and their separation took less than 7 min using a borate buffer (pH 8.8 at 25 mmol L?1) solution containing sodium dodecyl sulfate (45 mmol L?1) and acetonitrile (13.5% v/v). On-line analyte concentration (normal stacking mode) and the use of a highly sensitive cell (Z shaped cell) improved detection limits (at the 10?8 mol L?1 level). Recovery in fortified human saliva was 108 ± 5%, in agreement with the result achieved with the reference HPLC method. For the analysis of urine from rats submitted to a single dose of CPT-11 and SN-38, camptothecin was used as internal standard enabling recoveries close to 100% when compared to the results achieved using HPLC.  相似文献   

17.
Abstract— The equilibrium constants, Kc, for complexation between methyl viologen dication (MV2+) and Rose Bengal, or Eosin Y, decrease with increasing ionic strength. At zero ionic strength Kc is 6500 (± 500) mol?1 dm3 for Rose Bengal and 3200 (± 200) mol?1 dm3 for Eosin Y, and these values decrease to 1500 (± 100) and 680 (± 40) mol?1 dm3, respectively, at an ionic strength of 0.1 mol dm?3. Kc is independent of pH between 4.5 and 10. ΔH is -25 (± 1) kJ mol?1 for complexation with either dye, whereas ΔS is -15 (± 3) J K?1 mol?1 for Rose Bengal, and - 23 (± 3) J K?1 mol?1 for Eosin Y. The complexation constant for Rose Bengal and the neutral viologen, 4,4'-bipyridinium-N, N'-di(propylsulphonate), (4,4'-BPS), is 420 (± 35) mol?1 dm3, and independent of ionic strength. No complexation could be observed for either Rose Bengal or Eosin with another neutral viologen, 2,2'-bipyridinium-N,N'-di(propylsulphonate), (2,2'-BPS). MV2+ quenches the triplet state of Rose Bengal with a rate constant of 7 × 109 mol?1 dm3 s?1, and this rate constant decreases slightly as ionic strength increases. The cage escape yield following quenching, Φcc is very low (Φcc= 0.02 (± 0.005), and independent of ionic strength. 4,4'-BPS quenches the triplet state of Rose Bengal with a rate constant of 2.2 (± 0.1) × 109 mol?1 dm3 s?1, and gives a cage escape yield of 0.033 (± 0.006). 2,2'-BPS quenches the Rose Bengal triplet with a rate constant of 6 (± 1) × 108 mol?1 dm3 s?1 and gives a cage escape yield of 0.07 (± 0.01). Conductivity measurements indicate that MV2+(Cl?)2 is completely dissociated at concentrations below 2 × 10?2 mol dm?3.  相似文献   

18.
Complexation of some o-, m- and p-carborane derivatives with ??-cyclodextrin was investigated using phenolphthalein in pH 10.5 (0.05 mol dm?3) borate buffer. Some carborane derivatives indicated large inclusion constants Kass > 1 × 10dm3 mol?1.  相似文献   

19.
The kinetics of interaction between di-μ-hydroxobis(1,10-phenanthroline)dipalladium(II) perchlorate and thioglycolic acid and with glutathione has been studied spectrophotometrically in aqueous medium as a function of the complex concentration as well as the ligand concentrations, pH, and temperature at constant ionic strength. The observed pseudo-first-order rate constants k obs (s?1) obeyed the equation k obs = k 1[Nu] (Nu = nucleophile). At pH = 6.5, the interaction with thioglycolic acid shows two distinct consecutive steps and both steps are dependent on the concentration of thioglycolic acid. The rate constants for the process are: k 1 ≈ 10?5 s?1 and k 2 ≈ 10?3 dm3 · mol?1 · s?1. The association equilibrium constant (K E) for the outer sphere complex formation has been evaluated together with the rate constants for the two subsequent steps. The other bio-active ligand, glutathione, showed a single step reaction depending on [ligand] with a second-order anation rate constant: the 102 (k 2) values are (61.72, 79.20, 109.24 and 154.33) dm3 · mol?1 · s?1 at 20, 25, 30 and 35 °C, respectively. On the basis of the kinetic observations and evaluated activation parameters, plausible associative mechanisms are proposed for both interaction processes.  相似文献   

20.
Pulse radiolysis involving reactions of solvated electrons and benzophenone ketyl radicals in 1-octanol with selected compounds from bis-triazinyl pyridines and bis-triazinyl bipyridines, BT(B)P family, developed for extraction of trivalent actinides have been studied. The designated ligands were: 2,6-bis(5,5,8,8-tetramethyl-5,6,7,8-tetrahydro-benzo-[1,2,4]triazin-3-yl)pyridine, 6,6′-bis(5,5,8,8-tetramethyl-5,6,7,8-tetrahydro-benzo-[1,2,4-]triazin-3-yl)-[2,2′]bipyridine, 6,6′-bis(5,6-diethyl-[1,2,4]triazin-3-yl)-[2,2′]bipyridine and 6,6′-bis(5,6-dipentyl-[1,2,4]triazin-3-yl)-[2,2′]bipyridine. Reactions of the ligands with solvated electrons in 1-octanol are fast. The rate constants were determined as equal to: $ k_{{{\text{CyMe}}_{4} {\text{BTP}}}} . $  = (2.4 ± 0.2) × 109 dm3 mol?1 s?1, $ k_{{{\text{CyMe}}_{ 4} {\text{BTBP}}}} $  = (1.7 ± 0.3) × 109 dm3 mol?1 s?1, $ k_{{{\text{C}}_{ 2} {\text{BTBP}}}} $  = (1.3 ± 0.3) × 109 dm3 mol?1 s?1 and $ k_{{{\text{C}}_{ 5} {\text{BTBP}}}} $  = (1.7 ± 0.3) × 109 dm3 mol?1 s?1. Reactions of the ligands with benzophenone ketyl radicals are much slower and the measured rate constants were as follows: $ k_{{{\text{CyMe}}_{ 4} {\text{BTP}}}} $  = 6.7 × 107 dm3 mol?1 s?1 and $ k_{{{\text{CyMe}}_{ 4} {\text{BTBP}}}} $  = 3.2 × 107 dm3 mol?1 s?1.  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