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.
This paper presents the application of the tubular detector based on silver solid amalgam (TD‐AgSA) for electrochemical determinations of reducible inorganic (Cd2+, Zn2+) and organic (4‐nitrophenol) compounds under flow injection analysis conditions. The newly developed TD‐AgSA is simple, robust and inexpensive. The limits of detections of Zn2+, Cd2+ and 4‐nitrophenol are 1.4×10?6, 7.0×10?7, and 5.0×10?7 mol dm?3, respectively (i.e. 0.09, 0.08 and 0.07 ppm). The obtained results proved the long‐term stability of the detector (RSD of the determination of Zn2+, Cd2+, and 4‐nitrophenol were 0.8, 0.9 and 0.8 % (n=10; cZn=7.7×10?5 mol dm?3, cCd=4.5×10?5 mol dm?3 and c4‐NPh=3.6×10?5 mol dm?3), respectively and its applicability for cathodic measurements in aqueous solutions at potentials up to ?2 V.  相似文献   

2.
The differential pulse polarographic behaviour in dimethyl sulphoxide (DMSO) of 14 organotin(IV) compounds having the general formula R3SnX (R = Me, Ph; X? = NCS?, N3?, N3?, NO3?, OH?, NCO? and OAc?) and nBu3SnCl and nBu2SnCl2 has been studied. The peak potential was found to depend markedly on the organic group and to a lesser extent on the nature of the anion X. The phenyltin compounds were reduced at lower potentials than the corresponding methyltin compounds. The data obtained could be used for trace determination of these compounds. Linear calibration curves were obtained over the concentration range of 2.8 × 10?4 to 1.9 × 10?6 mol dm?3.  相似文献   

3.
The dependence of the limiting catalytic current in the system of In(III)-acetylsalycilic acid on the concentration of indium ions and the ligand anions was described theoretically. Dissociation and protonation constants of acetylsalycilic acid molecular form and kinetic parameter connected with the rate constant of formation of polarographically active complex {K a = 3.59×10?3, mol dm?3; K H = 0.12 mol dm?3; k = 4.0×102, A (mol/dm?3)2} were calculated.  相似文献   

4.
5.
The kinetics of oxidation of ethanol by cerium(IV) in presence of ruthenium(III) (in the order of 10?7 mol dm?3) in aqueous sulfuric acid media have been followed at different temperatures (25–40°C). The rate of disappearance of cerium(IV) in the title reaction increases sharply with increasing [C2H5OH] to a value independent of [C2H5OH] over a large range (0.2–1.0 mol dm?3) in which the rate law conforms to: where [Ru]T gives the total ruthenium (III) concentration. The values of 10?3kc and 10?3kd are 3.6 ± 0.1 dm3 mol?1 s?1 and 3.9 ± 0.2 s?1, respectively, at 40°C, I = 3.0 mol dm?3. The proposed mechanism involves the formation of ruthenium(III)? substrate complex which undergoes oxidation at the rate determining step by cerium(IV) to form ruthenium(IV)? substrate complex followed by the rapid red-ox decomposition giving rise to the catalyst and ethoxide radical which is oxidized by cerium(IV) rapidly. The mechanism is consistent with the existence of the complexes RuIII · (C2H5OH) and RuIII · (C2H5O?) and both are kinetically active. The overall bisulphate dependence conforms to: kobsd = A[Ru]T/{1 + C[HSO4?]} where A = 2.2 × 104 dm3 mol?1 s?1, C = 1.3 at 40°C, [H+] = 0.5 mol dm?3, and I = 3.0 mol dm?3. The observations are consistent with the Ce(SO4)2 as the kinetically active species. © 1995 John Wiley & Sons, Inc.  相似文献   

6.
《Electroanalysis》2006,18(2):158-162
Optimum conditions have been found for voltammetric determination of mutagenic 5‐aminoquinoline, 6‐aminoquinoline and 3‐aminoquinoline by differential pulse voltammetry and adsorptive stripping differential pulse voltammetry on carbon paste electrode. The lowest limits of determination were found for adsorptive stripping differential pulse voltammetry in 0.1 mol dm?3 H3PO4 (5×10?7 mol dm?3 , 1×10?7 mol dm?3, and 1×10?7 mol dm?3 for 5‐aminoquinoline, 6‐aminoquinoline and 3‐aminoquinoline, respectively). The possibility to determine mixtures of 8‐aminoquinoline with 3‐aminoquinoline or 5‐aminoquinoline or 6‐aminoquinoline, and mixtures of 5‐aminoquinoline with 3‐aminoquinoline or 6‐aminoquinoline by differential pulse voltammetry was verified. Binary mixtures of 8‐aminoquinoline with 3‐aminoquinoline or 6‐aminoquinoline, and of 3‐aminoquinoline with 5‐aminoquinoline could be successfully analyzed.  相似文献   

7.
Mononuclear copper(II) complex with 2,4-dioxo-4-(4-hydroxy-6-methyl-2-pyrone-3-yl) butyric acid ethyl ester is readily precipitated in ethanolic medium. The metal to ligand ratio in the crystalline species was found to be 1:2. On the basis of the spectroscopic data collected so far, the site of coordination could not be identified. The detection limit of the precipitation of the binuclear complex in aqueous buffer, pH 7.00, solution is at a 2 × 10?5 mol dm?3 copper(II) concentration. By radiometric measurements with 64Cu isotope, the time neccessary for a quantitative precipitation, the amount of copper(II) in the precipitate and in the solution, the amount of ligand needed for the optimal precipitation yield, and the solubility product of the complex were determined.The precipitate separated from the supernatant can be dissolved in ethanol and copper(II) determined by absorbance measurement at 374 nm. The sensitivity of this procedure lies at the detection limit of the complex precipitation. The calibration diagram, a straight line (b = 0.00677; sb = 0.00003; s2 = 0.00146), confirms the validity of Beer's law in the range of 2 × 10?5? 4 × 10?4 mol dm?3 copper(II) concentrations, with a systematic error of 7 × 10?6 mol dm?3 arising from solubility loss of the precipitate remaining constant.Concentrations exceeding 10?6 mol dm?3 of nickel(II) cause too high values and those exceeding 10?5 mol dm?3 of aluminium, zinc, iron(II), thorium(IV), or vanadium(V) too low values in copper determinations.  相似文献   

8.
The kinetics of acrylamide polymerization has been investigated by employing cericammoniumnitrate-2-chloroethanol redox pair under nitrogen atmosphere at 30 ± 1°C. The rate of monomer disappearance is directly proportional to the concentration of 2-chloroethanol (1.0 × 10?2 ? 10.0 × 10?2 mol. dm?3) and is inversely proportional to the ceric ion concentration (2.5 × 10?3 ? 10.0 × 10?3 mol. dm?3) but shows square dependence to the concentration of monomer (5.0 × 10?2 ? 25.0 × 10?2 mol. dm?3). The rate of ceric ion disappearance is directly proportional to the initial concentration of ceric ion and 2-chloroethanol but independent of acrylamide concentration. The viscometric average molecular weight (M v) decreases on increasing the concentration of ceric ion and increases on increasing the concentrations of acrylamide and 2-chloroethanol. A tentative mechanism has been proposed.  相似文献   

9.
In this paper, the flow amperometric enzymatic biosensor based on polished silver solid amalgam electrode for determination of sarcosine in model sample under flow injection analysis conditions is presented. The biosensor works on principle of electrochemical detection of oxygen decrease during enzymatic reaction which is directly proportional to the concentration of sarcosine in sample. The whole preparation process takes about 3 h. The RSD of repeatability of 10 consecutive measurements is 1.6 % (csarcosine=1.0×10?4 mol dm?3). Under optimal conditions the calibration dependence was linear in the range 7.5×10?6–5.0×10?4 mol dm?3 and limit of detection was 2.0×10?6 mol dm?3.  相似文献   

10.
Temperature-dependent NMR spectra indicate that the α-chamigren-3-ones (?) -11 , (+) -12 , (+) -14 (?) -15 , (+) -16, 18 , and 19 bearing equatorial halogen atoms at C(8) and C(9) undergo slow conformational flipping of the envelope-shaped enone ring, while the cyclohexane ring is maintained in the chair conformation. The α-chamigren-3-ols (+) -20 and (+) -21 , obtained by hydride reduction of (+) -12 , behave similarly, with slow half-chair inversion of the cyclohexenol ring. In each case, both conformers are about equally populated and detectable by NMR, except in the case of (+) -15 , where repulsive interactions between Br? C(2) and Heq?C(7) make the population of the conformer 15b with Me—C(5) faced to Hax?C(10) so low that it escapes direct 1H-NMR detection. The energy barriers to these conformational motions are viewed to arise mainly from repulsive interactions between Me—C(5) and the axial H-atoms at C(8) and C(10), while, contrary to previous beliefs, no twist-boat conformations of the cyclohexane ring intervene. Similar conclusions hold for the 4,5-epoxides of both (?) -6 and (+) -7 . Clean Jones oxidatio of (?) -2 to 17 , where the CH2?C(5) bond is maintained, and acid dehydration-isomerization of the α-chamigrene (+) -21 to the β-chamigrene (+) -24 , reflect the special stability of β-chamigrenes, providing a reason for their frequent occurrence in nature.  相似文献   

11.
ABSTRACT

The influence of octaethylene glycol mono-n-hexadecyl ether (C16EO8) and sodium dodecyl sulphate (SDS) on the crystallization of calcium oxalate from 0.3 mol dm?3 sodium chloride solutions has been investigated. The critical micellar concentration (CMC) of C16EO8 in water and 0.3 mol dm?3 NaCl was determined by surface tension measurements (CMCH2O=CMCNaCl = 7.2.10?6 mol dm?3). The kinetics of precipitation of calcium oxalate was followed by Coulter counter, and solid phases were characterized by polarized microscopy, thermal analysis and X-ray powder diffraction patterns. Under the precipitation conditions employed, the thermodynamically stable monohydrate, CaC2O4?H2O (COM) was the predominant crystalline form. In the presence of micellar solutions of C16EO8 precipitation of this phase was facilitated as evidenced by higher initial precipitation rates and higher precipitate volume and number of particles, as compared to the controls. Micellar solutions of 50S retarded precipitation but induced crystallization of calcium oxalate dihydrate, CaC2O4?(2+×)H2O (COD, x≤0.5). Thus at c(SDS>CMC the precipitates contained ≥85 mass % COD. The results are discussed in relation to previously reported data on the precipitation of calcium oxalate in the presence of dodecyl ammonium chloride  相似文献   

12.
The sorption of thioflavine T (TT) and malachite green (MG) cationic synthetic dyes on dried biomass of green microalga (Chlorella pyrenoidosa) immobilised in polyurethane foam under continuous column systems conditions using spectrophotometric methods of detection was investigated. Data characterising the sorption of TT and MG on microalgal biomass immobilised in polyurethane foam in a column system from single (C 0 = 25 μmol dm?3) or binary equimolar (C 0 = 25 μmol dm?3) dye solutions in the form of breakthrough curves were well described by the Thomas (R 2 = 0.994–0.912), Yoon-Nelson (R 2 = 0.994–0.911), and Clark (R 2 = 0.993–0.911) models. Useful parameters characterising the sorption column system were obtained from these mathematical models. The Thomas model, in particular, provided the Q max (maximal sorption capacity in μmol g?1) parameter for characterisation of biosorbent and also for evaluation of competitive effects in the TT and MG dyes sorption. For the purposes of biomass regeneration, a one-step desorption of the dyes studied from the microalgal biomass in batch and continuous column systems was performed. Efficiency of TT desorption from microalgal biomass increased in the order: deionised H2O (50.7 %), 99.5 vol. % 1,4-dioxane (67 %), 20 mmol dm?3 NiCl2 (83 %), 96 vol. % ethanol (85 %), 0.1 mol dm?3 HCl (89 %), 1 mol dm?3 acetic acid (89 %). In the case of MG, the desorption efficiency increased in the order: deionised H2O (13 %), 20 mmol dm?3 NiCl2 (50 %), 0.1 mol dm?3 HCl (91 %), 99.5 vol. % 1,4-dioxane (94 %), 1 mol dm?3 acetic acid (99 %), 96 vol. % ethanol (> 99 %). The presence of carboxyl, phosphoryl, amino, and hydroxyl groups, the important functional groups for sorption of cationic xenobiotics, was also confirmed on the algae biomass surface by potentiometric titration and ProtoFit modelling software. The data obtained showed that the dried immobilised algae biomass could be used as a sorbent for removing toxic xenobiotics from liquid wastewaters or contaminated waters and also presenting the possibilities of mathematical modelling of sorption processes in continuous column systems in order to obtain important parameters for use in practice.  相似文献   

13.
The system comprises l(+)- and d(?)-lactate dehydrogenase reactors in parallel and a diaphorase electrode. Separate peaks are obtained for l(+)- and d(?)-lactic acid. The peak current is linearly related to the concentration of both isomers in the range 1 × 10?5?2 × 10?3 M.  相似文献   

14.
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   

15.
《Analytical letters》2012,45(8):1491-1499
ABSTRACT

Glassy Carbon Electrodes coated with stearic acid provide an amperometric sensor for detection of paraquat, the active ingredient of the herbicide Gramoxone. The linear dynamic range of the sensor for Paraquat is 1.02 × 10?3 mol dm?3 to 1.02 × 10?2 mol dm?3 with the minimum detection limit 6.37 × 10?4 mol dm?3.  相似文献   

16.
The contact angles of aqueous solutions of a polymeric surfactant namely hydrophobically modified inulin (INUTEC®SP1) were measured on hydrophilic and hydrophobised quartz glass surfaces using the sessile drop technique. These measurements showed a large difference (>10°) between the advancing contact angle θ 1 (that is measured immediately after placing the drop on the surface) and the constant contact angle θ 2 (that is measured 30 minutes after placing the drop). In all the results only the contact angle θ 2 was subsequently measured. θ versus INUTEC®SP1 concentration C s curves were obtained at various NaCl concentrations both on hydrophilic and hydrophobic glass surfaces. On hydrophilic glass surface the θ versus C s curves showed a maximum at a concentration range of 10–6 to 2?×?10–5 mol dm-3 INUTEC®SP1. These curves were shifted to lower values as the NaCl concentration was increased. On such hydrophilic surface the INUTEC®SP1 molecule adsorbs with the polyfructose loops and tails oriented towards the surface leaving the alkyl chains in solution. Saturation adsorption with this orientation occurs at 2?×?10–5 mol dm-3 INUTEC®SP1. However, the contact angles remain quite small (<18°) indicating the presence of several hydrophilic glass patches uncovered by surfactant molecules. At C s?>?2?×?10–5 mol dm-3 θ decreases with further increase of the INUTEC®SP1 concentration reaching 5° at the Critical Association Concentration (CAC) of the polymer. This indicates the formation of a bilayer of INUTEC®SP1 molecules with the alkyl chains hydrophobically attached to those of the first layer. On a hydrophobic glass surface, adsorption of INUTEC®SP1 occurs by multi-point attachment with the alkyl chains on the surface leaving the hydrophilic polyfructose loops and tails dangling in solution. This results in a gradual decrease of the contact angle with increase in INUTEC®SP1 concentration, reaching a plateau value (>85°) between 2?×?10–5 and 2?×?10–4 mol dm-3. The large contact angles obtained on adsorption of the polymeric surfactant on a hydrophobic surface indicate the presence of several uncovered hydrophobic patches. These results give a reasonable picture of the adsorption and orientation of the INUTEC®SP1 molecules on both hydrophilic and hydrophobic solid surfaces.  相似文献   

17.
The extraction behaviour of ion-pairs formed by dialkylphosphorodithioates with tetraphenylarsonium cation was examined for the dichloromethane—water system. The mole ratio method, with spectrophotometric and conductimetric measurements of the organic phase, showed the mole ratio of the ion-pair to be 1:1. Dissociation of the ion-pairs occurred at concentrations in the aqueous phase lower than 10?1 mol dm?3 for the dimethyl compound and 6 × 10?4 mol dm?3 for the diethyl analogue. Under favourable experimental conditions, dialkylphosphorodithioates can be collected from aqueous solutions at concentrations not exceeding 1 μg cm?3 (ca. 5 × 10?6 mol dm?3), with extraction efficiencies of 86–95%. Dialkylphosphorothioates are extracted much less efficiently; diethylphosphate and inorganic phosphate are not extracted.  相似文献   

18.
The extraction of niobium(V) in the form of a chloro complex has been studied. Radiometrical and spectrophotometrical studies show that niobium(V) is extracted practically completely from a solution containing more than 9 mol dm?3 chloride in the range of 2–5 M hydrogen ion concentration by chloroform solutions of tetraphenylarsonium (TPA) and tetraphenylphosphonium (TPP) chloride and that niobium is not extracted at chloride concentrations less than 6 mol dm?3. The mechanism of extraction is based on the formation of the ion-associated compounds that form between the onium cation and the oxotetra-chloroniobate(V) anion. The extracted complexes in chloroform have a maximum absorbance at 282 nm (TPA) and 285 nm (TPP); they obey Beer's law in the range of 1–10 μg Nb ml?1, and are stable for at least 24 hr. The molar absorptivity of the method is 1.33 × 104 dm3 mol?1cm?1. The composition of the extracted species [(C6H5)4X] [NbOCl4] where X = As or P was determined spectrophotometrically, radiometrically, and by characterization of the crystalline compounds isolated.  相似文献   

19.
《Analytical letters》2012,45(6):1183-1191
Abstract

A study of the enhancement effect on the fluorescence intensity of the Eu3+–-thenoyltrifluoroacetone (TTA)–-cetyltri–-methylammonium bromide (CTMAB) and the Dy3+ pyrocatechol–-3,5-disulphonic acid (Tiron) systems by Y3+has been carried out. In the presence of yttrium the fluorescence intensity of the systems was enhanced by a factor of about 100 and 15, respectively. The fluorescence intensity was a linear function of the concentration of europium or dysprosium in the range 1.0 × 10?10–-1.0 × 10?8mol dm?3 and 8.0 × 10?8–-9.0 × 10?6mol dm?3, respectively. The detection limit was 1.0 × 10?11mol dm?3 and 1.0 × 10?10mol dm?3, respectively. The standard addition method was used for the determination of europium or dysprosium in rare earth oxides and gave satisfactory results. The mechanism of enhanced fluorescence was proposed.  相似文献   

20.
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   

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