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 125 毫秒
1.
林鹏  郑洪  朱庆枝  许金钩 《中国化学》2002,20(9):872-877
IntroductionThechemicalpropertiesofrare earthelementsareverysimilar ,consequently ,itisdifficulttodetectanin dividualionintheirmixturesowingtotheinterferenceofotherrare earthelements .Duetoahighersensitivityandselectivity ,fluorimetricmethodsofrare earthel…  相似文献   

2.
《Analytical letters》2012,45(10):2203-2215
Abstract

A new system of polarographic adsorptive wave for determining trace scandium was proposed. In 0.2 mol/L NH4OAc, the Sc(III)- ACBK [1,8- dihydroxy- 2- (2- hydroxy- 5- sulfo- 1- phenylazo)- 3,6- disulfo- naphthalene, called acid chrome blue K] complex emerged a sensitive adsorptive complex wave(Ep′ = -0.67V). The molar ratio of Sc(III) to ACBK in the complex was established as 1: 2 and the apparent stability constant β2 = 2.7 × 1015. But for Y(III), the molar ratio was 1: 1 and β = 1.5 × 105. Because of the particularity of Sc complex, the sensitivity and the selectivity of determination Sc are much better than that of other rare earth ions. The detection limit is 1.1 × 10?7 mol/L for oscillopolarography and 2.0 × 10?8 mol/L for adsorptive stripping voltammetry.  相似文献   

3.
A highly sensitive and selective catalytic adsorptive cathodic striping procedure for the determination of trace germanium is presented. The method is based on adsorptive accumulation of the Ge(IV)‐gallic acid (GA) complex onto a hanging mercury drop electrode, followed by reduction of the adsorbed species. The reduction current is enhanced catalytically by addition of vanadium(IV)‐EDTA. The optimal experimental conditions include the use of 0.03 mol/L HClO4 (pH1.6), 6.0×10?3 mol/L GA, 3.0×10?3 mol/L V(IV), 4.0×10?3 mol/L EDTA, an accumulation potential of ?0.10 V(vs. Ag/AgCl), an accumulation time of 120 s and a differential pulse potential scan mode. The peak current is proportional to the concentration of Ge(IV) over the range of 3.0×10?11 to 1.0×10?8 mol/L and the detection limit is 2×10?11 mol/L for a 120 s adsorption time. The relative standard deviation at 5.0×10?10 mol/L level is 3.1%. No serious interferences were found. The method was applied to the determination of germanium in ore, mineral water and vegetable samples with satisfactory results.  相似文献   

4.
《Analytical letters》2012,45(9):1741-1753
Abstract

An electroanalytical methood, based on derivative chronopotentiometry of the complex of beryllium(II) with 4-[(4-diethylamino-2-hydroxy-phenyl)-azo]-5-hydroxy-naphthalene-2,7-disurphonic acid (Beryllon II) accumulated adsorptively on the surface of a hanging mercury drop electrode, has been developed for determining trace beryllium in food. The dependences of the peak height on the dt/dE vs. E curve on the pre-concentration time, preconcentration potential and the constant reducing current are discussed. In 0.15 mol/1 NHs+0.05 mol/1 NH4Cl, 4×10?7 mol/l Beryllon II, and at a preconcentration potential of -0.30 V (ve. SCE), the limit of detection and linear range are 1 × l0?10 mol/l and 3 × 10?10 -2 × l0?7 mol/l, Iwpectively. The relative standard error of the method is 2.3% for 6 × 10?8 mol/l Be(II). The method WBB applied to samples of food. The electrode procees hae been diecueeed.  相似文献   

5.
A novel analytical approach for quantitative measurement of butylated hydroxyanisole (BHA) is dis‐ cussed in this paper. Such a method depends on the inhibitory effect of BHA on a Briggs‐Rauscher (B‐R) oscillating reaction. Unlike the classical B‐R system which involves Mn2+ as the catalyst, such a B‐R sys‐ tem is catalyzed by a macrocyclic nickel (II) complex [NiL](ClO4)2, where L in the complex is an unsatu‐ rated ligand 5,7,7,12,14,14‐hexemethyl‐1,4,8,11‐tetraazacyclotetradeca‐4,11‐diene. By perturbation of BHA on the system, the oscillation was inhibited in the presence trace amounts of BHA and the inhibition time was found to be proportional to the concentration of BHA over the range 1.00×10?7–1.20×10?4 mol/L. Two calibration curves were obtained: the first linear regression is over the range of 1.00×10?7–2.00×10?6 mol/L, and the second linear regression is over the range between 2.00×10?6 and 1.20×10?4 mol/L, with a lowest limit of detection of 4.00×10?8 mol/L. UV spectra measurements were employed to clarify the possible perturbation mechanism caused by BHA on the B‐R oscillating reaction.  相似文献   

6.
Rate constants for the reactions of OH and NO3 radicals with CH2?CHF (k1 and k4), CH2?CF2 (k2 and k5), and CHF?CF2 (k3 and k6) were determined by means of a relative rate method. The rate constants for OH radical reactions at 253–328 K were k1 = (1.20 ± 0.37) × 10?12 exp[(410 ± 90)/T], k2 = (1.51 ± 0.37) × 10?12 exp[(190 ± 70)/T], and k3 = (2.53 ± 0.60) × 10?12 exp[(340 ± 70)/T] cm3 molecule?1 s?1. The rate constants for NO3 radical reactions at 298 K were k4 = (1.78 ± 0.12) × 10?16 (CH2?CHF), k5 = (1.23 ± 0.02) × 10?16 (CH2?CF2), and k6 = (1.86 ± 0.09) × 10?16 (CHF?CF2) cm3 molecule?1 s?1. The rate constants for O3 reactions with CH2?CHF (k7), CH2?CF2 (k8), and CHF?CF2 (k9) were determined by means of an absolute rate method: k7 = (1.52 ± 0.22) × 10?15 exp[?(2280 ± 40)/T], k8 = (4.91 ± 2.30) × 10?16 exp[?(3360 ± 130)/T], and k9 = (5.70 ± 4.04) × 10?16 exp[?(2580 ± 200)/T] cm3 molecule?1 s?1 at 236–308 K. The errors reported are ±2 standard deviations and represent precision only. The tropospheric lifetimes of CH2?CHF, CH2?CF2, and CHF?CF2 with respect to reaction with OH radicals, NO3 radicals, and O3 were calculated to be 2.3, 4.4, and 1.6 days, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 619–628, 2010  相似文献   

7.
This study describes a fast and simple methodology for the preparation of Cerium (III) Hexacyanoferrate (II) (CeHCF) nanoparticles (NPs). The NPs were characterized by fourier transform infrared (FTIR), x‐ray diffraction (XRD), scanning electron microscopy (SEM) and cyclic voltammetry (CV). The CeHCF cyclic voltammogram indicate a well‐defined redox pair assigned as Fe2+/Fe3+ in the presence of cerium (III), with a formal potential of Eθ′=0.29 V (v=100 mV s?1, KNO3; 1.0 mol/L, pH 7.0). The carbon paste electrode modified with CeHCF (CeHCF‐CPE) was applied to the catalytic electrooxidation of dopamine applying Differential Pulse Voltammetry (DPV). DPV showed linear response at two concentration ranges, from 9.0×10?7 to 8.0×10?6 and 9.0×10?6 to 1.0×10?4 mol/L, with an LOD of 1.9×10?7 and 1.0×10?5 mol/L, respectively. The CeHCF‐CPE exhibited selectivity against substances commonly found in biological samples, with redox potentials close to that of dopamine, such as urea and ascorbic acid (AA). Subsequently the CeHCF‐CPE was successfully applied to the detection of dopamine in simulated urine samples, with recovery percentages ranging between 99 and 103%.  相似文献   

8.
The redox reaction between tris(1,10-phenanthroline)iron(II), [Fe(phen)3]2+, and azido-pentacyanocobaltate(III), [Co(CN)5N3]3? was investigated in three cationic surfactants: dodecyltrimethylammonium bromide (DTAB), tetradecyltrimethylammonium bromide (TTAB) and cetyltrimethylammonium bromide (CTAB) in the presence of 0.1?M NaCl at 35°C. Second-order rate constant in the absence and presence of surfactant, kw and kψ, respectively, were obtained in the concentration ranges DTAB?=?0???4.667?×?10?4?mol?dm?3, TTAB?=?0–9.364?×?10?5?mol?dm?3, CTAB?=?0???6.220?×?10?5?mol?dm?3. Electron transfer rate was inhibited by the surfactants with premicelllar activity. Inhibition factors, kw/kψ followed the trend CTAB?>?TTAB?>?DTAB with respect to the surfactant concentrations used. The magnitudes of the binding constants obtained suggest significant electrostatic and hydrophobic interactions. Activation parameters ΔH, ΔS, and Ea have larger positive values in the presence of surfactants than in surfactant-free medium. The electron transfer is proposed to proceed via outer-sphere mechanism in the presence of the surfactants.  相似文献   

9.
《Analytical letters》2012,45(15):2743-2753
Abstract

The electrochemical behavior of 6-MP was studied by cyclic voltammetry at a hanging copper amalgam dropping electrode (HCADE). It was found that 6-MP could form a complex with the Cu(II) stripped from the HCADE, showing a new peak at ?0.19V in the medium of 0.1mol/L LiClO4-0.5mol/L HClO4 solution. The mechanism of the reaction was proposed. This new peak was sensitive and could be used for the determination of trace 6-MP by differential pulse adsorption cathodic stripping voltammetry (DPAdCSV). The linear range was from 3.6×10?10 to 5.3×10?6 mol/L, and the detection limit was about 1.2×10?10 mol/L (S/N=3). The method was also successfully applied to the determination of 6-MP in pharmaceutical tablets.  相似文献   

10.
A capillary electrophoresis coupled with tris(2,2′‐bipyridyl) ruthenium(II) (Ru(bpy)32+) electrochemiluminescence detection system was developed to determine salbutamol and clenbuterol in urine. Some factors that affected the performances of separation and detection were investigated. Under the optimized conditions, one single quantitative analysis of salbutamol and clenbuterol was achieved at a separation voltage of 15 kV within 9 min, and the LODs (S/N=3) and LOQs (S/N=10) of salbutamol and clenbuterol were 8.43×10?8 mol/L, 2.61×10?7 mol/L and 2.73×10?7 mol/L, 8.21×10?7 mol/L, respectively. The recovery obtained from the analysis of spiked urine samples was between 88.6 % and 104.7 % with RSDs lower than 6.70 %. The method was successfully applied to determine salbutamol and clenbuterol in urine samples.  相似文献   

11.
Emission quenching of [Ru(bpy)2(4, 4'-dcbpy)] (PF6)2 (1) by benzenamine,4-[2-[5-[4-[4-dimethylamino]phenyl]-4,5-di-hydro-1-phenyl-1H-pyrazol-3-yl]-ethenyl]-N,N-dimetyl (2) or 1, 5-diphenyl-3-(2-phenothiazine)-2-pyrazoline (3) was observed. Measurements of the emission decay of 1 before and after addition of 2 or 3 by single photon counting technique con-finned the observations. The emission quenching of 1 by 2 or 3 was submitted to Stern-Volmer equation. It was calculated that the quenching rate constants (kq) are 5.5 × 109(mol/L)-1s-1 for 2 and 4.0 × 109(mol/L)-1s-1 for 3, respectively. These results indicated a character of dynamic quenching process. The singlet-state of 2 or 3 was also quenched by 1. The quenching behaviors did not conform to the Stern- Volmer equation and involved both static and dynamic quenching processes. The apparent quenching rate constant (kapp) was calculated to be 3 × 109 (mol/L)-1 for the interaction of excited 2 with 1, and 1.2 × 109 (mol/L)-1 for that of excited 3 wit  相似文献   

12.
A new method for the determination of trace copper was described. A multiwalled carbon nanotube modified carbon paste electrode was prepared and the adsorptive voltammetric behavior of copper‐alizarin red S (ARS) complex at the modified electrode was investigated. By use of the second‐order derivative linear sweep voltammetry, it was found that in 0.04 mol/L acetate buffer solution (pH 4.2) containing 4×10?6 mol/L ARS, when accumulation potential is 0 mV, accumulation time is 60 s and scan rate is 100 mV/s, the complex can be adsorbed on the surface of the electrode, yielding one sensitive reduction peak at ?172 mV (vs. SCE). The peak current of the complex is proportional to the concentration of Cu(II) in the range of 2.0×10?11–4.0×10?7 mol L?1 with a detection limit (S/N=3) of 8.0×10?12 mol/L (4 min accumulation). The proposed method was successfully applied to the determination of copper in biological samples with satisfactory results, the recoveries were found to be 96%–102%.  相似文献   

13.
Complexation in the Co(II)-phosphonomethylaminosuccinic acid (H4L) system in aqueous solutions at component ratios of 1 : 1 and 1 : 2 and c Co(II) = 1 × 10?2 mol/L was studied by electronic absorption spectroscopy. The formation of various protonated complexes of the general formula Co(H n L) m (OH) q (n = 3?0; m = 1?C2; q = 0?C2) was found, and their stability constants and distribution diagrams were calculated. It was demonstrated that the bis complexes have the structure of a distorted octahedron, and the octahedron ?? tetrahedron rearrangement of the coordination polyhedron occurs in the equimolar complexes at pH > 8.  相似文献   

14.
This paper reports the synthesis, characterization, and dehydration kinetics of a rare earth hydroxide, Gd(OH)3. Uniform rod‐like Gd(OH)3 powder was prepared by a colloidal hydrothermal method. The powder thus obtained dehydrated into its oxide form in a two‐step process, where crystalline GdOOH was obtained as the intermediate phase. Crystal structure study revealed a monoclinic structure for GdOOH, with space group P2/1m and lattice parameters a = 6.0633, b = 3.7107, c = 4.3266, and β = 108.669. The first‐step dehydration follows the F2 mechanism, while the second step follows the F1 model, indicating that both the steps are controlled by nucleation/growth mechanism. The activation energy Ea and frequency factor A are 231±12 kJ/mol and 2.08 × 1018 s?1 for the first step and 496 ± 32 kJ/mol and 7.88 × 1033 s?1 for the second step, respectively. Such high activation energy calculated from the experimental data can be ascribed to the high bonding energy of Gd? O bond, and the difference in activation energy for the two steps is due to the change in the bond length of hexagonal Gd(OH)3 and monoclinic GdOOH. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 39: 75–81, 2007  相似文献   

15.
The mechanism by which an excess of iron(II) ion reacts with aqueous chlorine dioxide to produce iron(III) ion and chloride ion has been determined. The reaction proceeds via the formation of chlorite ion, which in turn reacts with additional iron(II) to produce the observed products. The first step of the process, the reduction of chlorine dioxide to chlorite ion, is fast compared to the subsequent reduction of chlorite by iron(II). The overall stoichiometry is The rate is independent of pH over the range from 3.5 to 7.5, but the reaction is assisted by the presence of acetate ion. Thus the rate law is given by At an ionic strength of 2.0 M and at 25°C, ku = (3.9 ± 0.1) × 103 L mol?1 s?1 and kc = (6 ± 1) × 104 L mol?1 s?1. The formation constant for the acetatoiron(II) complex, Kf, at an ionic strength of 2.0 M and 25°C was found to be (4.8 ± 0.8) × 10?2 L mol?1. The activation parameters for the reaction were determined and compared to those for iron(II) ion reacting directly with chlorite ion. At 0.1 M ionic strength, the activation parameters for the two reactions were found to be identical within experimental error. The values of ΔH? and ΔS? are 64 ± 3 kJ mol?1 and + 40 ± 10 J K?1 mol?1 respectively. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 554–565, 2004  相似文献   

16.
The kinetics of oxidation of tartaric acid (TAR) by peroxomonosulfate (PMS) in the presence of Cu(II) and Ni(II) ions was studied in the pH range 4.05–5.20 and also in alkaline medium (pH ~12.7). The rate was calculated by measuring the [PMS] at various time intervals. The metal ions concentration range used in the kinetic studies was 2.50 × 10?5 to 1.00 × 10?4 M [Cu(II)], 2.50 × 10?4 to 2.00 × 10?3M [Ni(II)], 0.05 to 0.10 M [TAR], and µ = 0.15 M. The metal(II) tartarates, not TAR/tartarate, are oxidized by PMS. The oxidation of copper(II) tartarate at the acidic pH shows an appreciable induction period, usually 30–60 min, as in classical autocatalysis reaction. The induction period in nickel(II) tartarate is small. Analysis of the [PMS]–time profile shows that the reactions proceed through autocatalysis. In alkaline medium, the Cu(II) tartarate–PMS reaction involves autocatalysis whereas Ni(II) tartarate obeys simple first‐order kinetics with respect to [PMS]. The calculated rate constants for the initial oxidation (k1) and catalyzed oxidation (k2) at [TAR] = 0.05 M, pH 4.05, and 31°C are Cu(II) (1.00 × 10?4 M): k1 = 4.12 × 10?6 s?1, k2 = 7.76 × 10?1 M?1s?1 and Ni(II) (1.00 × 10?3 M): k1 = 5.80 × 10?5 s?1, k2 = 8.11 × 10?2 M?1 s?1. The results suggest that the initial reaction is the oxidative decarboxylation of the tartarate to an aldehyde. The aldehyde intermediate may react with the alpha hydroxyl group of the tartarate to give a hemi acetal, which may be responsible for the autocatalysis. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 620–630, 2011  相似文献   

17.
Introduction  Ionchromatography (IC)hasbeenrecognizedasausefulmethodfortheseparationofinorganicanionsandcationssinceitsintroductionbySmalletal .in 1975 .1AsignificanttrendinthedevelopmentofICmethodissearchforsensitiveanduniversaldetectionmethods .Themaindet…  相似文献   

18.
In this paper a carbon ionic liquid electrode (CILE) was fabricated by using ionic liquid 1‐ethyl‐3‐methylimidazolium ethylsulphate ([EMIM]EtOSO3) as the modifier and further used as the working electrode for the sensitive anodic stripping voltammetric detection of Pb2+. The characteristics of the CILE were investigated by cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS). In pH 4.5 NaAc‐HAc buffer Pb2+ was accumulated on the surface of CILE due to the extraction effect of IL and reduced at a negative potential (‐1.20 V). Then the reduced Pb was oxidized by differential pulse anodic stripping voltammetry with an obvious stripping peak appeared at ?0.67 V. Under the optimal conditions Pb2+ could be detected in the concentration range from 1.0 × 10?8 mol/L to 1.0 × 10?6 mol/L with the linear regression equation as Ip(μA) = ?0.103 C (μmol/L) + 0.0376 (γ = 0.999) and the detection limit as 3.0 × l0?9 mol/L (3σ). Interferences from other metal ions were investigated and Cd2+ could be simultaneously detected in the mixture solution. The proposed method was further applied to the trace levels of Pb2+ detection in water samples with satisfactory results.  相似文献   

19.
IntroductionChlordiazepoxide (7 chloro 2 methylamino 5 phenyl 3H 1,4 benzodiazepine 4 oxide)showingpowerfulan tianxietyeffecthasbeenwidelyusedasapsychotherapeu ticdrug .Consequently ,theneedaroseforsensitiveandrapiddeterminationofchlordiazepoxideinblood ,urinean…  相似文献   

20.
The rate coefficients for the removal of Cl atoms by reaction with three HCFCs, CF3CHCl2 (HCFC-123), CF3CHFCl (HCFC-124), and CH3CFCl2 (HCFC 141b), were measured as a function of temperature between 276 and 397 K. CH3CF2Cl (HCFC-142b) was studied only at 298 K. The Arrhenius expressions obtained are: k1 = (3.94 ± 0.84)× 10?12 exp[?(1740 ± 100)/T] cm3 molecule?1 s?1 for CF3CHCl2 (HCFC 123); k2 = (1.16 ± 0.41) × 10?12 exp[?(1800 ± 150)/T] cm3 molecule?1 s?1 for CF3CHFCl (HCFC 124); and k3 = (1.6 ± 1.1) × 10?12 exp[?(1800 ± 500)/T] cm3 molecule?1 s?1 for CH3CFCl2 (HCFC 141b). In case of HCFC 141b, non-Arrhenius behavior was observed at temperatures above ca. 350 K and is attributed to the thermal decomposition of CH2CFCl2 product into Cl + CH2CFCl. In case of HCFC-142b, only an upper limit for the 298 K value of the rate coefficient was obtained. The atmospheric significance of these results are discussed. © 1993 John Wiley & Sons, Inc.  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