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 15 毫秒
1.
A large data set obtained by a one-year monthly determination of ions (F, Cl, Br, NO3, NO2, PO43−, SO42−, Na+, K+, Ca2+, Mg2+, NH4+) and trace metals of environmental concern (Ni, Co, Mn, Fe) from the tributaries of Lake Como (Lombardy, Northern Italy) was treated by three-way Principal Component Analysis. The results showed that the chemical features of the investigated rivers are mainly related to the lithology of the watershed. Some cases of contamination were evidenced and rationalized on the basis of anthropic pollution or on the basis of the geochemical features of the territory. The method here proposed allows an easy and quick interpretation of the chemical data by means of graphical devices. The information extracted by the three-way models would be very useful to regional agencies in developing a strategy to manage water resources in the whole basin of Lake Como.  相似文献   

2.
A chemiluminescence (CL) method for the determination of humic acid (HA) based on the oxidation of HA with hydrogen peroxide in the presence of formaldehyde in alkaline solution is described. This method is sensitive and selective for the determination of HA in natural water. HA produces strong CL in the oxidation of HA with MnO4, Br2, ClO, and Cr2O72−, and the H2O2. HA-H2O2-HCHO system is suitable for the determination of HA because of its high sensitivity and high selectivity. The detection limit was 50 ppb and relative standard deviation for five measurements of 0.5 ppm (w/w) HA was 1.8%. Cations such as Na+, K+, Mg2+, Cu2+, and Fe3+ and anions such as PO43−, NO3, CO32−, SO42−, Cl, and Y (EDTA-Na) did not interfere with the determination of HA. Addition of Mn(II) increased the CL intensity. The concentration of HA in natural water determined with this method is in good agreement with that determined by fluorometric analysis.  相似文献   

3.
A sensitive voltammetric method has been developed for the determination of total or single species of sulfur anions containing sulfide, sulfite and thiosulfate. The method is based on the catalytic effect of tris(2,2'-bipyridyl)Ruthenium(II) (Ru(bpy)2+ 2) as a homogeneous mediator on the oxidation of those anions at the surface of a glassy carbon electrode. A reversible redox couple of Ru(II)/Ru(III) were observed as a solute in aqueous solution. Cyclic voltammetry study showed that the catalytic current of the system depends on the concentration of the anions. Optimum pH values for voltammetric determination of sulfite, thiosulfate and sulfide has been found to be 5.6, 10.0 and 10.0, respectively. Under the optimized conditions the calibration curves have been obtained linear in the concentration ranges of 0.8–500.0, 0.4–1000.0 and 0.5–5000.0 µmol L− 1 of SO32−, S2O32− and S2−, respectively. The detection limits have been calculated to be 0.40, 0.17 and 0.33 µmol L− 1 for SO32−, S2O32− and S2−, respectively. The diffusion coefficients of sulfite and thiosulfate have been estimated using chronoamperometry. The chronoamperometric method also has been used to determine the catalytic rate constant for catalytic reaction of the Ru(bpy)2+ 2 with sulfite and thiosulfate. Finally the proposed method has been used for the determination of total sulfur contents in real samples of water and wastewater. Moreover the sulfite content in sugar and sulfur dioxide in air has been determined with satisfactory results.  相似文献   

4.
The MnIV complex of tetra-deprotonated 1,8-bis(2-hydroxybenzamide)-3,6-diazaoctane (MnIVL) engrossed in phenolate-amido-amine coordination is reduced by HSO3 and SO32− in the pH range 3.15–7.3 displaying biphasic kinetics, the MnIIIL being the reactive intermediate. The MnIIIL species has been characterized by u.v.–vis. spectra {λ max, (ε, dm3 mol−1 cm−1): 285(15 570), 330 sh (7570), 469(6472), 520 sh (5665), pH=5.42}. SO42− was the major oxidation product of SIV; dithionate is also formed (18 ± 2% of [MnIV]T) which suggests that dimerisation of SO3−• is competitive with its fast oxidation by MnIV/III. The rates and activation parameters for MnIVL + HSO3 (SO32−) → MnIIIL; MnIIIL + HSO3 (SO32−) → MnIIL2− are reported at 28.5–45.0 °C (I=0.3 mol dm−3, 10% (v/v) MeOH + H2O). Reduction by SO32− is ca. eight times faster than by HSO3 both for MnIVL and MnIIIL. There was no evidence of HSO3/SO32− coordination to the Mn centre indicating an outer sphere (ET) mechanism which is further supported by an isokinetic relationship. The self exchange rate constant (k22) for the redox couple, MnIIIL/MnIVL (1.5 × 106 dm3 mol−1 s−1 at 25 °C) is reported.  相似文献   

5.
Cu2+ binding on γ-Al2O3 is modulated by common electrolyte ions such as Mg2+, , and in a complex manner: (a) At high concentrations of electrolyte ions, Cu2+ uptake by γ-Al2O3 is inhibited. This is partially due to bulk ionic strength effects and, mostly, due to direct competition between Mg2+ and Cu2+ ions for the SO surface sites of γ-Al2O3. (b) At low concentrations of electrolyte ions, Cu2+ uptake by γ-Al2O3 can be enhanced. This is due to synergistic coadsorption of Cu2+ and electrolyte anions, and . This results in the formation of ternary surface species (SOH2SO4Cu)+, (SOH2PO4Cu), and (SOH2HPO4Cu)+ which enhance Cu2+ uptake at pH < 6. The effect of phosphate ions may be particularly strong resulting in a 100% Cu uptake by the oxide surface. (c) EPR spectroscopy shows that at pH  pHPZC, Cu2+ coordinates to one SO group. Phosphate anions form stronger, binary or ternary, surface species than sulfate anions. At pH  pHPZC Cu2+ may coordinate to two SO groups. At pH  pHPZC electrolyte ions and are bridging one O-atom from the γ-Al2O3 surface and one Cu2+ ion forming ternary [γ-Al2O3/elecrolyte/Cu2+] species.  相似文献   

6.
Surleva AR  Neshkova MT 《Talanta》2008,76(4):914-921
A new flow injection approach to total weak acid-dissociable (WAD) metal–cyanide complexes is proposed, which eliminates the need of a separation step (such as gas diffusion or pervaporation) prior to the detection. The cornerstone of the new methodology is based on the highly selective flow-injection potentiometric detection (FIPD) system that makes use of thin-layer electroplated silver chalcogenide ion-selective membranes of non-trivial composition and surface morphology: Ag2 + δSe1 − xTex and Ag2 + δSe. An inherent feature of the FIP-detectors is their specific response to the sum of simple CN + Zn(CN)42− + Cd(CN)42−. For total WAD cyanide determination, ligand exchange (LE) and a newly developed electrochemical pre-treatment procedure for release of the bound cyanide were used. The LE pre-treatment ensures complete recovery only when the sample does not contain Hg(CN)42−. This limitation is overcome by implementing electrochemical pre-treatment which liberates completely the bound WAD cyanide through cathodic reduction of the complexed metal ions. A complete recovery of toxic WAD cyanide is achieved in the concentration range from 156 μg L−1 up to 13 mg L−1. A three-step protocol for individual and group WAD cyanide speciation is proposed for the first time. The speciation protocol comprises three successive measurements: (i) of non-treated, (ii) LE-exchange pre-treated; (iii) electrochemically pre-treated sample. In the presence of all WAD complexes this procedure provides complete recovery of the total bound cyanide along with its quantitative differentiation into the following groups: (1) Hg(CN)42−; (2) CN + Cd(CN)42− + Zn(CN)42−; (3) Cu(CN)43− + Ni(CN)42− + Ag(CN)2. The presence of a 100-fold excess in total of the following ions: CO32−, SCN, NH4+, SO42− and Cl does not interferes. Thus the proposed approach offers a step ahead to meeting the ever increasing demand for cyanide-species-specific methods. The equipment simplicity makes the procedure a good candidate for implementing in portable devices for in-field cyanide monitoring.  相似文献   

7.
TiO2 photocatalytic mineralization of β-naphthol: influence of some inorganic ions, ethanol, and hydrogen peroxide. In this work, the photocatalytic oxidation of β-naphthol in aqueous suspensions of TiO2 was investigated at room temperature, by following the formation of CO2. The disappearance of β-naphthol fits a Langmuir-Hinshelwood kinetic model. The activation energy for the degradation reaction of β-naphthol is estimated at 10.2 kJ/mol. The effects of some additives such as ethanol, H2O2, and inorganic ions (Cl, SO42−, HCO3, NO3, Fe3+, Cu2+, and Cr3+) on the photomineralization of β-naphthol were examined. The inhibition of the anions for this reaction was in the order : NO3 < HCO3 < SO42− < Cl. This can be due to a partial blockage of catalyst active sites by these ions or their reaction with an oxidizing radical such as OH. The most photoactive systems for β-naphthol degradation were found in the presence of ferric ions, while the addition of Cr3+ strongly inhibited the photocatalytic decomposition of β-naphthol.  相似文献   

8.
The scanning reference electrode technique (SRET) was used to characterize the corrosion behavior of carbon steel in the NaNO2-containing chloride solution and tap water. In 10−3 MNaNO2+10−3 MNaCl+10−3 MNa2SO4 solution, the passivated carbon steel surface suffered the pitting. In this case, the size of anodic and cathodic areas on the corroding surface varied with the exposed time, but anodic and cathodic sites did not change. On the contrary, the carbon steel was corroded generally in the tap water. The localized anodes and cathodes on the corroding surface were not fixed but movable with the exposed time during the whole corrosion process, and the “movable anodes” can be in situ monitored and momentarily identified by SRET measurements.  相似文献   

9.
When the sodium ion (Na+) concentration is increased above 0.5 mol-dm−3 (M), the concentrations of dissolved silica in aqueous sodium chloride (NaCl) and sodium nitrate (NaNO3) solutions decrease because of the salting out effect. On the other hand, the concentration of the dissolved silica in aqueous sodium sulfate (Na2SO4) solutions increases monotonously as the concentration of Na+ is increased above 0.5 M. The purpose of this study is to determine the reasons why the salting-out effect is not observed in Na2SO4 solutions. FAB-MS (Fast Atom Bombardment Mass Spectrometry) was used to sample directly the silica species dissolved in aqueous Na2SO4, NaCl, and NaNO3 solutions. In the FAB-MS spectra of these solutions, the peak intensity ratios of the linear tetramer to the cyclic tetramer largely increased for Na+ concentrations between (0.1 and 1) M. This shows that some characteristics of the Na2SO4 solutions are similar to those of the NaCl and NaNO3 solutions. In Na2SO4 solutions, however, when the concentration of Na+ is higher than 1 M, the peak intensity of the dimer is much higher than those of the other silicate complexes. In Na2SO4 solutions, the SO42− ion undergoes partial hydrolysis to form HSO4 and OH is produced. In particular, in the range where the concentration of SO42− is high, the pH of the solution increases slightly. This higher pH yields more dimers from the hydrolysis of silicate complexes. This increase in dimer production agrees with the observation that silica dissolves in sodium hydroxide (NaOH) solutions mainly as a dimer when the concentration of NaOH is less than 0.1 M. In Na2SO4 solutions at high concentrations, a salting-out effect is not observed for silica. This is due to the increase in the concentration of OH, which accelerates the hydrolysis of silica and results in dimer formation.  相似文献   

10.
The effects of CO complexation on highly exothermic vanadium oxidation reactions is evaluated. We study the chemiluminescent (CL) reaction products formed when vanadium vapor entrained in Ar or CO is oxidized by O3 or NO2. The multiple collision V+Ar+O3→VO*(C 4Σ, 4Φ, 2X)+Ar+O2 reactive encounter yields two previously unreported VO excited states, whereas the V+Ar+NO2→VO*+Ar+NO reactive encounter populates states up to and including VO* C 4Σ. The multiple collision V+nCO+O3 reactive encounter would appear to form a VOCO excited state complex, emitting in the region 420–560 nm, via the formation and oxidation of V(CO)2 viz. V(CO)2+O3→VOCO*+CO+O2 and a relaxed VO excited state emitter via V+nCO+O3→VO*+nCO+O2 where the VO excited state excitation is mediated by V–CO complexation. In complement, the much less exothermic V–NO2 encounter displays an emission which, in concert with previous studies of CO complexation, suggests the formation of a VO(CO)2 excited state complex viz. V(CO)2+NO2→VO(CO)2*+NO. The experiments characterizing CL are complemented by comparative laser-induced fluorescence studies of the VO X 4Σ–CO and Ar interactions and their influence on the VO C 4Σ–X 4Σ laser-induced excitation spectrum. These studies, in conjunction with further attempts to excite LIF in the 420–560 nm region, suggest that the observed complex emissions result primarily from VO excited state interactions. Complementary time-of-flight mass spectroscopy of vanadium and vanadium-oxide–carbonyl complex formation demonstrates the formation of V(CO), V(CO)2, V2(CO), and VOCO, the latter three of which demonstrate clear metastable-ion dissociation peaks for the processes VOCO+→V++CO2, V(CO)2+→V++2CO, and V2(CO)+→V2++CO, suggesting that these vanadium complexes when formed in a reaction-based environment may be photodissociated with light in the visible and ultraviolet regions.  相似文献   

11.
A new ion chromatography method is described for the simultaneous determination of Cl, NO3 and SO42−, using a selected eluent 1.3-mM sodium gluconate/1.3-mM borax (pH 8.5). The extraction methods of Cl, NO3, SO42− in vegetables are studied. The determination limits of Cl, NO3, SO42− are 0.17 μg/ml, 0.63 μg/ml and 0.81 μg/ml. The linear ranges are 060 μg/ml, 090 μg/ml and 090 μg/ml. The relative S.D. are <2.5%. The mean recoveries of Cl, NO3, SO42− in vegetables range from 97.0 to 104%.  相似文献   

12.
An electrochemical quartz crystal microbalance was employed to monitor directly the growth of vanadium hexacyanoferrate (VHF) films on platinum substrates during electrodeposition and interfacial coagulation in the solution containing sulfuric acid electrolyte, vanadium(IV) and hexacyanoferrate(III). Mass changes of the gold/crystal working electrode were correlated with cyclic voltammetry data. Effects of cations (NH4+, Li+, Na+ and K+), anions (SO42− and NO3) and solvent during redox reactions of the films were studied. The results show that cations were incorporated into the film during reduction and expelled from the film during oxidation. Solvent also participates in VHF electrochemistry, and its role cannot be neglected. Anions play no role in VHF electrochemistry.  相似文献   

13.
Cyclic voltammetry of antimony was studied in aqueous media (HCl-LiCl) and in nonaqueous media after extraction with 20% tri-n-butylphosphate in toluene, with a rotating glassy carbon disc electrode. Reduction of antimony to the element in aqueous media is nearly reversible, but irreversible in nonaqueous media. Anodic stripping voltammetric and chronopotentiometric determinations were also studied in nonaqueous media; methanol and LiCI, NH4SCN or NH4NO3, were used as base electrolytes. In nonaqueous media, antimony can be determined down to concentrations of 1O−8 M by stripping voltammetry, and lO−7 M by stripping chronopotentiometry. Electrochemical stripping determinations of 10−6 M antimony(III) were not affected by Co2+, Ni2+, Cd2+, Zn2+ or As3+ (5 · 10−3 M), ag+ (4 · 10+4 M in stripping voltammetry or 10−3 M in stripping chronopotentiometry), Hg2+ (5 · 10−4M), Pb2+ (3 · 10−4 M), Cu2+ (1.5 · 10−4 M)Sn2+ and Sn4+ (7 · 10−4 M), Fe3+ (4 · 10−4 M), Au3+ (5 · 10−5 M) and Bi3+ (1.5 · 10−5 M). Thestripping chronopotentiometric determination showed better selectivity.  相似文献   

14.
Room temperature rate coefficients and product distributions are reported for the reactions initiated in D2O with dications of the alkaline-earth metals Mg, Ca, Sr and Ba. The measurements were performed with a selected-ion flow tube (SIFT) tandem mass spectrometer and electrospray ionization (ESI). Mg2+ reacts with water by a fast electron transfer leading to charge separation with a rate coefficient of 1.4 × 10−9 cm3 molecule−1 s−1. Ca2+ reacts with D2O in a first step to form the adduct Ca2+(D2O), with an effective bimolecular rate coefficient of 2.3 × 10−11 cm3 molecule−1 s−1, which then undergoes rapid charge separation by deuteron transfer to form CaOD+ and D3O+ in a second step with k = 7.9 × 10−10 cm3 molecule−1 s−1. The CaOD+ ion reacts further by clustering up to five more D2O molecules. Sr2+ clusters up to eight D2O molecules and Ba2+ up to seven D2O molecules, with the first addition of D2O being rate determining in each case and the last addition being distinctly slower, as might be expected from a transition in the occupation of the added water molecules from an inner to an outer hydration shell.  相似文献   

15.
The synthesis of goethite by oxidation of Fe2+in presence of metallic iron was undertaken in an aqueous medium containing indifferent salts such as Na2SO4, (NH4)2SO4, NaCl, and NH4Cl. Temperature and bubbling air rate were maintained, respectively, at 70°C and 1 L/min. The influence of anions and cations on the kinetics of each step of the process has been followed distinctly, the iron dissolution rate has been determined by the variation of the medium acidity, and the precipitation of goethite has been determined by gravimetric measurements. With respect to Cl, the SO42−anion decreases the rate of the two reactions. NH4+acts as an inhibitor when it is present at low concentrations and as an accelerator for higher concentrations; the limit corresponding to the change of NH4+behavior depends on the nature of the counter ion. The reaction product is composed of pure goethite in the presence of sulfate salts, whereas a mixture of goethite and lepidocrocite, respectively, 60–70 and 40–30%, was observed in the presence of chloride salts.  相似文献   

16.
Recombination rate coefficients of protonated and deuterated ions KrH+, KrD+, XeH+ and XeD+ were measured using Flowing Afterglow with Langmuir Probe (FALP). Helium at 1600 Pa and at temperature 250 K was used as a buffer gas in the experiments. Kr, Xe, H2 and D2 were introduced to a flow tube to form the desired ions. Because of small differences in proton affinities of Kr, D2 and H2 mixtures of ions, KrD+/D3+ and KrH+/H3+ are formed in the afterglow plasma, influencing the plasma decay. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The obtained rate coefficients, αKrD+(250 K) = (0.9 ± 0.3) × 10−8 cm3 s−1 and αXeD+(250 K) = (8 ± 2) × 10−8 cm3 s−1 are compared with αKrH+(250 K) = (2.0 ± 0.6) × 10−8 cm3 s−1 and αXeH+(250 K) = (8 ± 2) × 10−8 cm3 s−1.  相似文献   

17.
The rate constants and product ion branching ratios were measured for the reactions of various small negative ions with O2(X 3Σg) and O2(a 1Δg) in a selected ion flow tube (SIFT). Only NH2 and CH3O were found to react with O2(X) and both reactions were slow. CH3O reacted by hydride transfer, both with and without electron detachment. NH2 formed both OH, as observed previously, and O2, the latter via endothermic charge transfer. A temperature study revealed a negative temperature dependence for the former channel and Arrhenius behavior for the endothermic channel, resulting in an overall rate constant with a minimum at 500 K. SF6, SF4, SO3 and CO3 were found to react with O2(a 1Δg) with rate constants less than 10−11 cm3 s−1. NH2 reacted rapidly with O2(a 1Δg) by charge transfer. The reactions of HO2 and SO2 proceeded moderately with competition between Penning detachment and charge transfer. SO2 produced a SO4 cluster product in 2% of reactions and HO2 produced O3 in 13% of the reactions. CH3O proceeded essentially at the collision rate by hydride transfer, again both with and without electron detachment. These results show that charge transfer to O2(a 1Δg) occurs readily if the there are no restrictions on the ion beyond the reaction thermodynamics. The SO2 and HO2 reactions with O2(a) are the only known reactions involving Penning detachment besides the reaction with O2 studied previously [R.S. Berry, Phys. Chem. Chem. Phys., 7 (2005) 289–290].  相似文献   

18.
The title compound, cobalt 4′,7-diethoxylisoflavone-3′-sulfonate([Co(H2O)6](X)2⋅8H2O, X = C19H17O4SO3) was synthesized and its structure was determined by single-crystal X-ray diffraction analysis. It crystallizes in the triclinic space group P-1 with cell parameters a = 9.026(3) Å, b = 16.431(5) Å, c = 18.195(6) Å, α = 72.289(4), β = 87.498(4), γ = 82.775(5), V = 2550.1(13) Å−3, Dc = 1.419 Mg m−3, and Z = 2. The results show that the title compound consists of one cobalt cation, six coordinated water molecules, eight lattice water molecules, and two 4′,7-diethoxylisoflavone-3′-sulfonate anions, C19H17O4SO3. Two anions have different conformations. Twelve H atoms of six coordinated water molecules, as donors, form hydrogen bonds with four oxygen atoms of sulfo-groups of two anions and eight oxygen atoms of eight lattice water molecules. In addition, π < eqid1 > ⋅ < eqid2 > π stacking interactions exist in the crystal structure, which together with hydrogen bonds lead to supramolecular formation with a three-dimensional network.  相似文献   

19.
A new heterometallic Bi(III) complex with diethylenetriaminepentaacetic acid anion (Dtpa)5− of the composition [Co(Tsc)3]2[Bi(Dtpa)]2SO4 ⋅ 6H2O (I) (Tsc is thiosemicarbazide) is synthesized and its crystal structure is determined. The complex consists of the [Co(Tsc)3]3+ cations, [Bi(Dtpa)]2− and SO 4 2− anions, and crystallization water molecules. The SO 4 2− ion and two water molecules are randomly disordered over two positions. In the complex cation [Co(Tsc)3]3+, the metal polyhedron has fac-form. The carboxyl groups of octadentate (Dtpa)5− ligand in the [Bi(Dtpa)]2− anion are fully deprotonated. The coordination polyhedron of the Bi atom is a distorted bicapped trigonal prism. Thermogravimetric analysis of complex I indicates that its decomposition occurs through several stages, i.e., dehydration, burning of organic ligands, and the formation of inorganic residue.__________Translated from Koordinatsionnaya Khimiya, Vol. 31, No. 6, 2005, pp. 446–454.Original Russian Text Copyright © 2005 by Bulimestru, Petrenko, Gulea, Gdaniec, Simonov.  相似文献   

20.
The complex species formed in aqueous solution (25 C, I = 3.0 mol-dm−3 KCl ionic medium) between V3+ cation and the ligands: picolinic acid (Hpic, HL) and dipicolinic acid (H2dipic, H2L), have been studied potentiometrically and by spectrophotometric measurements. The application of the least-squares computer program LETAGROP to the experimental emf (H) data, taking into account the hydrolytic species of V3+ ion, indicates that under the employed experimental conditions, the formation of the complexes [VL]2+, [V(OH)L]+, [VL2]+, [VL3], [V2OL4] with picolinic acid and the complexes [VL]+, [V(OH)L], [V(OH)2L], [V(HL)(L)], and [VL2] with dipicolinic acid were observed. The stability constants of the complexes formed were determined by potentiometric measurements, and spectrophotometric measurements were done in order to perform a qualitative characterization of the complexes formed in aqueous solution.  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