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.
《Analytical letters》2012,45(8):1203-1218
Abstract

Ultraviolet (UV)–derivative spectrophotometry, bivariate calibration algorithm, and Vierodt methods were applied to simultaneous determination of ranitidine (R) and amoxicillin (AMX) in binary mixtures. The first-order derivative allows determination of R in the concentration range 4.0 · 10?6 mol · dm?3 to 6.0 · 10?5 mol · dm?3. Vierordt method enables Ranitidine assaying in the presence of 2.5-fold excess of AMX and 3-fold excess of R. The bivariate calibration method obeys Beer's law in the concentration ranges 4.0 · 10?6 mol · dm?3 to 6.0 · 10?5 mol · dm?3 for R and 2.0 · 10?6 to 2.0 · 10?5 mol · dm?3 for AMX.  相似文献   

2.
《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   

3.
The fundamental properties and extraction capability of an ionic liquid (IL), trioctylammonium nitrate ([HTOA][NO3]), for PdII and PtIV, are investigated. At room temperature, [HTOA][NO3] is a solid (melting point: 30.7 °C), but it becomes a liquid (melting point: 16.7 °C) when saturated with water. Water-saturated [HTOA][NO3] exhibits a viscosity of 267.1 mPa·s and an aqueous solubility of 2.821?×?10?4 mol·dm?3 at 25 °C, and can be used as an extraction solvent without dilution. [HTOA][NO3] exhibits an extremely high extraction capability for PdII and PtIV in dilute hydrochloric acid (0.1–2 mol·dm?3 HCl); the distribution ratio reaches 3 × 104 for both the metals. From electrospray ionization mass spectrometry analysis, the species extracted in the IL phase are [PdCl3]? and [PdCl2(NO3)]? for PdII and [PtCl6]2? and [PtCl5]? for PtIV. A majority of the other transition metals are considerably less or marginally extracted into [HTOA][NO3] from a 0.1 mol·dm?3 hydrochloric acid solution. The extraction capacity of [HTOA][NO3] is greater than that of other hydrophobic ILs such as [HTOA]Cl and bis(trifluoromethanesulfonyl)imide-based ILs. The metals extracted into the IL phase are quantitatively back-extracted using an aqueous solution containing thiourea and nitric acid. By controlling the thiourea concentration and shaking time, PdII and PtIV are mutually separated to some extent in the back extraction process. The IL phase used for the back extraction can be reused for the forward extraction of these metals after scrubbing it with an aqueous nitric acid solution.  相似文献   

4.

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   

5.
《Analytical letters》2012,45(7-8):1367-1376
The voltammetric behavior of epinephrine was investigated by differential pulse voltammetry (DPV) at a carbon paste electrodes (CPE) made of different carbon powders – CR-2, glassy carbon (GC) microparticles, and single-wall carbon nanotubes (SWNT). In Briton-Robinson (BR) buffer solution pH 6, the linear dependence was found for the determination of epinephrine by the given method in the concentration ranges of 1·10?6–1·10?4 (CR-2), 1·10?6–1·10?4 (GC microparticles) and 4·10?6–1·10?4 (SWNT) mol·dm?3. Limits of detection were 8·10?7, 8·10?7, and 2·10?6 mol·dm?3, respectively. The best results were obtained employing CPE containing carbon paste with 50% (m/m) of SWNT, a linear dynamic range being 4·10?7–1·10?4 mol·dm?3 and a limit of detection 2·10?7 mol·dm?3. The attempt to increase the sensitivity by adsorptive accumulation of epinephrine was not successful.  相似文献   

6.

Dynamic interfacial tension (DIT) and interface adsorption kinetics at the n‐decane/water interface of 3‐dodecyloxy‐2‐hydroxypropyl trimethyl ammonium chloride (R12TAC) were measured using spinning drop method. The effects of RnTAC concentration and temperature on DIT have been investigated, the reason of the change of DIT with time has been discussed. The effective diffusion coefficient, D a, and the adsorption barrier, ?a, have been obtained with extended Word‐Tordai equation. The results show that the higher the concentration of surfactants is, and the smaller will be the DIT and the lower will be the curve of the DIT, and the R12TAC solutions follow a mixed diffusion‐activation adsorption mechanism in this investigation. With increase of concentration in bulk solution of R12TAC from 8×10?4 mol · dm?3 to 4×10?3 mol · dm?3, D a decreases from 2.02×10?10 m?2 · s?1 to 1.4×10?11 m?2 · s?1 and ? a increases from 2.60 kJ · mol?1 to 9.32 kJ · mol?1, while with increase of temperature from 30°C to 50°C, D a increases from 2.02×10?10 m?2 · s?1 to 5.86×10?10 m?2 · s?1 and εa decreases from 2.60 kJ · mol?1 to 0.73 kJ · mol?1. This indicates that the diffusion tendency becomes weak with increase strength of the interaction between surfactant molecules and that the thermo‐motion of molecules favors interface adsorption.  相似文献   

7.
The static (relative) permittivity of aqueous electrolyte solutions at 25 °C, ε(ω = 0) where ω is the frequency of the external field, has been obtained from the literature. The limiting dielectric decrements δ = ? lim(c → 0)dε/dc, where c is the electrolyte concentration (≤ 1 mol·dm?3), are tabulated. The most appropriate additive individual ionic δ i values are derived. The effect of ion pairing in the solutions is briefly discussed.  相似文献   

8.
The kinetics and mechanism of the reduction of the μ-adi-di[N,N′-bis{salicylideneethylenediaminatoiron(III)}] complex, [Fe2adi], by dithionate ion, S2O6 2?, have been investigated in aqueous perchloric acid at 29 °C, I = 0.05 mol dm?3 (NaClO4) and [H+] = 5.0 × 10?3 mol dm?3. Spectrophotometric titrations indicated that one mole of the reductant was oxidized per mole of oxidant. Kinetic profiles indicated first-order rate with respect to [Fe2adi] but zeroth-order dependence on [S2O6 2?]. The rate of reaction increased with increase in [H+], decreased with increased dielectric constant, but was invariant to changes in ionic strength of the medium. Addition of small amounts of AcO? and Mg2+ ions did not catalyse the reaction. A least-squares fit of rate against [H+]2 was linear (r 2 = 0.984) without intercept. The reaction was analysed on the basis of a proton-coupled outer-sphere electron transfer mechanism.  相似文献   

9.
Density measurements of good precision are reported for aqueous and aqueous salt (KBr) solutions containing 2,2,2-cryptand (4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo[8.8.8]hexacosane) (~0.009 to ~0.24 mol·kg?1) for the binary systems and for the ternary system with ~0.1 mol·kg?1 2,2,2-cryptand and varying KBr concentrations (~0.06 to ~ 0.16 mol·kg?1) at 298.15 K. The density data have been used to study the variation of apparent molar volume (\( \varphi_{V} \)) of 2,2,2-cryptand and of KBr as a function of concentration. 2,2,2-Cryptand is a diamine and hence it is hydrolyzed in aqueous solutions and needs an appropriate methodology to obtain meaningful thermodynamic properties. We have adopted a method of hydrolysis correction developed initially by Cabani et al. and later by Kaulgud et al. to analyze our volumetric data for the aqueous solutions. The method is described and we were successful in obtaining the limiting partial molar volume of the bare (free) cryptand in water at 298.15 K. Volumes of ionization as well as volumes of complexation (with KBr) are calculated. Estimations of the apparent molar volume of 2,2,2-cryptand in CCl4 are also reported. There is a loss in volume for the cryptand on transferring it from CCl4 to water. The volume changes due to ionization for the cryptand in water are calculated to be –20.5 and –0.6 cm3·mol?1 for the mono- and di-protonation equilibria respectively, while the volume of complexation for K+ is +24.5 cm3·mol?1. The results are discussed in terms of conformation, protonation equilibria and selective encapsulation of K+ ions in cryptand cavities. The solution volume properties seem to depend upon water–solute interaction as well on the solute–solute association because of hydrophobic interactions caused by lowering of the charge density on formation of cryptand-K+ species in solution.  相似文献   

10.
The kinetics of the oxidation of tris(2,2′-bipyridyl)iron(II) and tris(1,10-phenanthroline)iron(II) complexes ([Fe(LL)3]2+, LL = bipy, phen) by nitropentacyanocobaltate(III) complex [Co(CN)5NO2]3? was investigated in acidic aqueous solutions at ionic strength of I = 0.1 mol dm?3 (HCl/NaCl). The reactions were carried out at fixed acid concentration ([H+] = 0.01 mol dm?3) and the temperature maintained at 35.0 ± 0.1 °C. Spectroscopic evidence is presented for the protonated oxidant. Protonation constants of 360.43 and 563.82 dm3 mol?1 were obtained for the monoprotonated and diprotonated Co(III) complexes respectively. Electron transfer rates were generally faster for [Fe(bipy)3]2+ than [Fe(phen)3]2+. The redox complexes formed ion-pairs with the oxidant with increasing concentration of the oxidant over that of the reductant. Ion-pair constants for these reaction were 160.31 and 131.9 dm3 mol?1 for [Fe(bipy)3]2+ and [Fe(phen)3]2+, respectively. The activation parameters measured for these systems have values as follows: ?H (kJ K?1 mol?1) = +113.4 ± 0.4 and +119 ± 0.3; ?S (J K?1) = +107.6 ± 1.3 and 125.0 ± 1.6; ?G (kJ K?1) = +81 ± 0.4 and +82.4 ± 0.4; and E a (kJ mol?1) = 115.9 ± 0.5 and 122.3 ± 0.6 for LL = bipy and phen, respectively. Effect of added anions (Cl?, $ {\text{SO}}_{4}^{2 - } $ and $ {\text{ClO}}_{4}^{ - } $ ) on the systems showed decrease in the electron transfer rate constant. An outer-sphere mechanism is proposed for the reaction.  相似文献   

11.
The kinetics of the oxidation of ketorolac by hexacyanoferrate(III) (HCF) in aqueous alkaline medium at a constant ionic strength of 0.75 mol·dm?3 was studied spectrophotometrically at 300 K. A plausible mechanism was proposed and the rate law was derived. The mechanism of oxidation of ketorolac (KET) in alkaline medium has been shown to proceed via a KET-HCF complex, which decomposes in a slow step followed by other fast steps to give the products. The main oxidative product was identified as (2,3-dihydro-1-hydroxy-1H-pyrrolizin-5-yl-)(phenyl)methanone and is characterized by its LC–ESI–MS spectrum. Thermodynamic parameters of various equilibria of the mechanism were calculated and activation parameters ΔH , ΔS , ΔG and log10 A were found to be 29.9 kJ·mol?1, ?220 J·K?1·mol?1, 96 kJ·mol?1 and 2.70 respectively.  相似文献   

12.
The properties and extraction for [Ni(NH3)6]2+ of anionic aqueous two-phase systems (ATPS-a) that formed in mixtures of cetyltrimethylammonium bromide (CTAB) and excess sodium dodecyl sulfate (SDS) aqueous solutions were investigated. The results showed that the properties and extraction effects were strongly affected by the surfactant concentration, the temperature of system, and the mole fraction of surfactants. The increase of temperature induces narrower phase region and larger phase volume ratio. In addition, [Ni(NH3)6]2+ was extracted into the surfactant-rich phase with higher distribution coefficient when the liquid crystal had the birefringent properties. Moreover, the distribution coefficient can be improved through reducing the concentration of surfactant from 0.15 to 0.05 mol · L?1 or increasing mole fraction of CTAB from 21.9% to 23.1%. The results showed that ATPS of cationic–anionic surfactants was efficient for [Ni(NH3)6]2+ extraction with distribution coefficients of 13.5 when the total surfactant concentration was 0.05 mol · L?1, mole fraction of CTAB was 21%, and temperature was 34°C.  相似文献   

13.
The thermodynamic properties of 2-mercaptopyridine-N-oxide (pyrithione, PT) were studied potentiometrically in NaCl aqueous solutions at different ionic strengths and temperatures. A set of protonation constants is provided, together with distribution (water/2-methyl-1-propanol) and solubility data. The total and the specific solubility (solubility of neutral species) values of pyrithione were determined and, for example, are 0.0561 and 0.0518 mol·dm?3 at c NaCl = 0.244 mol·dm?3 and T = 298.15 K. By fitting the distribution and solubility results against the ionic strength, the Setschenow coefficient and the activity coefficients of the neutral species were determined. In pure water, the specific solubility is log10 \( S_{m 0}^{0} = \, {-} 1. 20 \, \pm \, 0.0 4 \) . To determine the activity coefficient of the charged species and the protonation constant at infinite dilution, the data were analyzed by different models, namely the Debye–Hückel type equation, the SIT (Specific ion Interaction Theory) and the Pitzer approach. The interaction coefficient of the deprotonated pyrithione species was determined [ε(Na+, PT?) = ?0.105 ± 0.002]. The protonation enthalpy was also determined, is slightly positive, and the protonation process is entropic in nature. At infinite dilution and T = 298.15 K, log10 K H0 = 4.620 ± 0.002, ΔG 0 = –26.4 ± 0.1 kJ·mol?1, ΔH 0 = 2.1 ± 0.5 kJ·mol?1 and TΔS 0 = 28.5 ± 0.5 kJ·mol?1. The electrochemical behavior of pyrithione was studied in NaCl solutions at T = 298.15 K. It was found that voltammetry can be used to study the binding ability of pyrithione towards metal cations. The results of this work are in agreement with literature findings and improve the knowledge of the chemistry of pyrithione in aqueous solutions.  相似文献   

14.

The graft copolymerization of acrylonitrile (AN) and ethyl acrylate (EA) comonomers onto cellulose has been carried out using ceric ammonium nitrate (CAN) as an initiator in the presence of nitric acid at 35±0.1°C. The addition of ethyl acrylate as comonomer has shown a significant effect on overall and individual graft copolymerization of acrylonitrile on cellulose. The graft yield (%GY) and other grafting parameters viz. true grafting (%GT), graft conversion (%CG), cellulose number (Ng) and frequency of grafting (GF) were evaluated on varying the concentration of comonomers from 6.0–30.0×10?1 mol dm?3 and ceric (IV) ions concentration from 2.5–25×10?3 mol dm?3 at constant feed composition (fAN 0.6) and constant concentration of nitric acid (7.5×10?2 mol dm?3) in the reaction mixture. The graft yield (%GY) and other grafting parameters were optimal at 15×10?1 mol dm?3 concentration of comonomers and at 10×10?3 mol dm?3 concentration of ceric ammonium nitrate. The graft yield (%GY) and composition of grafted chains (FAN) was optimal at a feed composition (fAN) of 0.6. The energy of activation (Ea) for graft copolymerization has been found to be 16 kJ mol?1. The molecular weight (Mw) and molecular weight distribution (Mw/Mn) of grafted chains was determined by GPC and found to be optimum at 15×10?1 mol dm?3 concentration of comonomer in the reaction mixture. The composition of grafted chains (FAN) determined by IR method was used to calculate the reactivity ratios of monomers, which has been found to be 0.62 (r1) and 1.52 (r2), respectively for acrylonitrile (AN) and ethyl acrylate (EA) monomers used for graft copolymerization. The energy of activation for decomposition of cellulose and grafted cellulose was determining by using different models based on constant and different rate (β) of heating. Considering experimental observations, the reaction steps for graft copolymerization were proposed.  相似文献   

15.
The Pd(DAP)Cl2 complex, where DAP is 2,6-diaminopyridine, was synthesized and characterized. The stoichiometries and stability constants of the complexes formed between various biologically relevant ligands (amino acids, amides, DNA constituents, and dicarboxylic acids) and [Pd(DAP)(H2O)2]2+ were investigated at 25 °C and at constant 0.1 mol·dm?3 ionic strength. The concentration distribution diagrams of the various species formed were evaluated. A further investigation of the binding properties of the diaqua complex [Pd(DAP)(H2O)2]2+ with calf thymus DNA (CT-DNA) was investigated by UV–Vis spectroscopy. The intrinsic binding constants (K b) calculated from UV–Vis absorption studies is 1.04 × 103 mol·dm?3. The calculated (K b) value was found to be of lower magnitude than that of the classical intercalator EB (ethidium bromide) (K b = 1.23 (±0.07) × 105 mol·dm?3), suggesting an electrostatic and/or groove binding mode for the interaction with CT-DNA.  相似文献   

16.
The initiation of the polymerization of acrylamide by 4-4′-dicyano-4-4′-azopentanoic acid in aqueous solution has been studied kinetically at 25°C. Ferric chloride and ferric sulfate were used to terminate polymerization so that rates of initiation could be calculated from the rates of production of ferrous iron. Velocity coefficients at 25°C. for the initiation reaction were found to be (25.7 ± 2.8) × 10?7 sec.?1 for the ferric chloride terminated reaction and (73.6 ± 0.6) × 10?7 sec.?1 for the ferric sulfate-terminated polymerization. The value reported for the initiation reaction when acrylamide is polymerized in the absence of metal salts is 1.29 × 10?7 sec.?1. Velocity coefficients for the termination reaction have been calculated from the overall rates of polymerization obtained with ferric salts present. In the case of the ferric chloride-terminated reaction, it has been shown that the rate of polymerization is reduced by increasing the total concentration of chloride ions. Termination velocity coefficients at 25°C. for the inner sphere complexes FeCl2+·5H2O and FeSO4+·4H2O have been calculated to be 18.9 × 104 and 7.98 × 104 l./mole-sec., respectively. The dependence on the concentration of ferric chloride of the molecular weights of the polymers produced has also been considered.  相似文献   

17.
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   

18.
The properties of diflunisal, a widely used analgesic, were studied in physiologic solutions, 0.15 mol·dm?3 NaCl. Solubility and protonation constants were determined and its behavior as ligand towards Ca(II) and Mg(II) was investigated. Solubility and protonation constants of diflunisal at 25 °C and 0.15 mol·dm?3 were obtained from electromotive force measurements of galvanic cells using coulometric titrations. The experimental data yielded the solubility, s, of –log10 s = 3.86 ± 0.02 and the protonation constants log10 K 1 = 11.98 ± 0.10 and log10 K 2 = 3.86 ± 0.03. Equilibria between diflunisal and Ca(II) and Mg(II) were investigated by means of electromotive force measurements and by comparing solubilities of diflunisal in the presence and absence of Ca(II) or Mg(II), respectively. Experimental data were explained by assuming the formation of 1:1 complexes for Ca(II) and Mg(II) along with evaluating the relative stability constants.  相似文献   

19.
Taylor dispersion is widely used to measure binary mutual diffusion. Studies of three- and four-component solutions show that the dispersion method is also well suited for multicomponent diffusion measurements, including cross-coefficients for coupled diffusion. Numerical procedures are reported here to calculate mutual diffusion coefficients from dispersion profiles measured for solutions of any number of components. The proposed analysis is used to measure the sixteen quinary mutual diffusion coefficients of five-component aqueous mannitol + glycine + urea + KCl solutions and aqueous NBu4Cl + LiCl + KCl + HCl solutions. Mannitol, glycine, urea and KCl interact weakly at the low solute concentrations used (0.010 mol·dm?3). The diffusion coefficients of this system are compared with pseudo-binary predictions. Strong coupling of the NBu4Cl, LiCl, KCl and HCl fluxes is interpreted by using ionic conductivities and Nernst equations to calculate limiting quinary diffusion coefficients for mixed electrolytes that interact by the electric field generated by ion concentration gradients.  相似文献   

20.
A new approach has been proposed to study the structure of aqueous electrolyte solutions. NIR, Raman and attenuated total reflectance, Fourier transform infra-red (ATR FTIR) spectra have been measured for aqueous calcium perchlorate solutions in the 0.22–4.3 mol·L?1 (0.22–7.46 mol·kg?1) concentration range at 25 °C. By the methods of principal component analysis (PCA) and multivariate curve resolution—alternating least squares (MCR-ALS) the number, spectra and concentration profiles have been determined for spectroscopically distinguishable forms of water and ClO 4 ? ion in solutions. The results have been analyzed using a phenomenological model, establishing thereby: concentration ranges for structural rearrangements of the solution, the nature of structural microirregularities and different states of the ClO 4 ? ion in the areas of domination of the natural water structure, and of cybotactic groups of calcium perchlorate hexa and tetra hydrates.  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