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.
Fuhrman DL  Latimer GW  Bishop J 《Talanta》1966,13(1):103-108
Ethylenediaminetetra-acetic acid (EDTA) and nitrilotri-acetic acid (NTA) can be differentiated and determined by titration with metal ions to visual metallochromic dye end-points. EDTA can be determined without interference from NTA, either by titrating with copper(II) at pH 5 using PAN indicator, or by titrating with iron(III) at pH 6 and 70 degrees using Tiron indicator. The total chelating power (EDTA + NTA) can be determined either by titrating with lead(II) at pH 4.4 using dithizone indicator, or by titrating with iron(III) at pH 3.5 using Tiron indicator ; NTA is determined by difference. The lowest concentration at which NTA can be determined in EDTA by titration to the iron(III)-Tiron end-point is about 1 wt.%. The apparent stability constants of the iron(III)-Tiron complexes under the conditions of the titration at pH 3.5 and pH 6 have been determined using the method of continuous variations.  相似文献   

2.
The use of a commercially available chelating resin with NTA-type functional groups for concentration of trace metals from seawater is described. Trace metal recoveries from this NTA Superflow chelating resin are pH dependent. At a pH of ≤2 only iron(III) and copper are quantitatively recovered from the resin. Iron(II) cannot be quantitatively recovered from this resin below a pH of 5. However, oxidation of acidified seawater samples (pH 1.7) with H2O2 prior to loading onto the resin has been demonstrated to allow quantitative recovery of total dissolved iron. Deferrioxamine and Rhodoturlic Acid, two commercially available siderophores were used to investigate the effect of strong Fe(III)-binding organic ligands on the ability to retain iron at different pH values. Acidification of seawater samples to pH 1.7 dissociates the iron complexed to these organic ligands, thereby allowing total dissolved iron and copper to be determined. Acidified samples from Monterey Bay were analyzed by a flow injection method coupled to ICP-SFMS detection using the NTA Superflow resin in the pre-concentration step. Results from this study show that when seawater samples are stored acidified (pH 1.7) over time, a portion of iron(III) is reduced to iron(II), thus necessitating the use of H2O2 to reoxidize the Fe(II) to Fe(III) prior to analysis. Total dissolved concentrations of iron and copper can be directly obtained on seawater samples at pH 1.7 with this method, eliminating the need to buffer the sample to a higher pH prior to column loading. This resin has the potential to be used in shipboard or in situ flow injection methods.  相似文献   

3.
Matsuda T  Nagai T 《Talanta》1986,33(7):614-616
Degradation of nitrilotriacetic acid (NTA) by oxidation with lead dioxide suspension has been studied by differential pulse polarography. The NTA was degraded over the pH range from 4 to 9, with formation of glycine or a mixture of iminodiacetic acid and glycine. After shaking with lead dioxide for 1 hr at 30 degrees and pH ~7, the NTA was almost completely decomposed, the molar reacting ratio of Pb(IV) to NTA being ~17:1; down to 1 x 10(-5)M NTA was decomposed in a shaking time as short as 15 min and at a temperature as low as 5 degrees . The iron(III)-NTA complex was also degraded under the same conditions, and the iron released was adsorbed on the lead dioxide.  相似文献   

4.
A new and convenient spectrophotometric method for the estimation of vanadium(IV) with NTA is described. The minimum ratio of metal ion to ligand, working pH, wavelength for maximum absorbance of the complex ion, and the effect of various cations and anions are described. The complex ion obeys Beer's law in the concentration range 1–32 mmol/liter of the vanadium(IV) ion. It is observed that iron(II), cobalt(II), nickel(II), copper(II), and oxidizing anions such as chromate and nitrite interfere in this determination, whereas managanese(II), chromium(III), iron(III), and anions like nitrate, chloride, bromide, iodide, thiocyanate, sulfate, and sulfite do not have any effect. Excessive amounts of acetate, phosphate, oxalate, tartrate and thiosulfate must also be avoided in this determination. Anions and cations which interfere in the determination of vanadium(IV) by NTA should not be present in the system.  相似文献   

5.
Summary Spectropbotometric Determination of Trace Amounts of Iron(III) by Extraction of Mixed-Ligand Iron-Tartrate-Purpurin or Iron-NTA-Purpurin Complex A selective method is described for the determination of microgram amounts of iron(III) by means of its reaction at pH 9.0 with purpurin (1,2,4-trihydroxyanthraquinone) and tartrate or NTA and extraction into methyl isobutyl ketone. The molar absorptivity of the 112 iron(III)-auxiliary ligand-purpurin complex is 4.8×104 1·mole–1·cm–1 at 595 nm. Beer's law is obeyed from 0.05 to 0.25 ppm of iron in the aqueous phase. Procedures for determination of iron in tartrate or NTA medium, and fluoride-tartrate-NTA medium are given. The method is suitable for determining iron in Zn metal, W metal, NTA, drinking water, wines, urine and tartrates.  相似文献   

6.
The hydrogen peroxide decomposition kinetics were investigated for both “free” iron catalyst [Fe(II) and Fe(III)] and complexed iron catalyst [Fe(II) and Fe(III)] complexed with DTPA, EDTA, EGTA, and NTA as ligands (L). A kinetic model for free iron catalyst was derived assuming the formation of a reversible complex (Fe–HO2), followed by an irreversible decomposition and using the pseudo‐steady‐state hypothesis (PSSH). This resulted in a first‐order rate at low H2O2 concentrations and a zero order rate at high H2O2 concentrations. The rate constants were determined using the method of initial rates of hydrogen peroxide decomposition. Complexed iron catalysts extend the region of significant activity to pH 2–10 vs. 2–4 for Fenton's reagent (free iron catalyst). A rate expression for Fe(III) complexes was derived using a mechanism similar to that of free iron, except that a L–Fe–HO2 complex was reversibly formed, and subsequently decayed irreversibly into products. The pH plays a major role in the decomposition rate and was incorporated into the rate law by considering the metal complex specie, that is, EDTA–Fe–H, EDTA–Fe–(H2O), EDTA–Fe–(OH), or EDTA–Fe–(OH)2, as a separate complex with its unique kinetic coefficients. A model was then developed to describe the decomposition of H2O2 from pH 2–10 (initial rates = 1 × 10−4 to 1 × 10−7 M/s). In the neutral pH range (pH 6–9), the complexed iron catalyzed reactions still exhibited significant rates of reaction. At low pH, the Fe(II) was mostly uncomplexed and in the free form. The rate constants for the Fe(III)–L complexes are strongly dependent on the stability constant, KML, for the Fe(III)–L complex. The rates of reaction were in descending order NTA > EGTA > EDTA > DTPA, which are consistent with the respective log KMLs for the Fe(III) complexes. Because the method of initial rates was used, the mechanism does not include the subsequent reactions, which may occur. For the complexed iron systems, the peroxide also attacks the chelating agent and by‐product‐complexing reactions occur. Accordingly, the model is valid only in the initial stages of reaction for the complexed system. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 24–35, 2000  相似文献   

7.
A method is described for the potentiometric titration with iron(III) of EDTA and NTA in detergents, a platinum electrode being used as indicator. EDTA and NTA were extracted at pH 9 and 50–60 °C in the presence of magnesium(II). Interference from polyphosphates was minimized by hydrolysis to orthophosphate followed by remotion as magnesium ammonium phosphate. Complexometric titration was carried out at pH 4.7 in acetate medium, in the presence of ferroin. A preliminary cleaning of the platinum electrode with thiosemicarbazide in hydrochloric acid was found to improve significantly the potential measurements. A mean recovery of 93% for EDTA and 89% for NTA was observed.  相似文献   

8.
The effect of iron(III) ions on the potential of the chalcocite electrode was investigated. Linear graphs were obtained for pFe(III) = 2-4, and were suitable for analytical purposes. The effect of ligands complexing iron(III) was studied, and the potential shown to be due to the concentration of free iron(III) ions only. The pH effect is mainly connected with solution reactions. A mechanism of potential response, based on a redox reaction, has been postulated, but the response does not depend on the redox potential in the bulk of the solution.  相似文献   

9.
Determinations of the aqueous iron species Fe(II) and Fe(III) are essential for a fully-informed understanding of redox processes involving iron. Most previous methods for speciation of iron have been based on the calorimetric determination of Fe(II) followed by reduction of Fe(III) and analysis for total iron. The indirect determination of Fe(III) and the consumption of relatively large sample volumes have limited the accuracy and utility of such methods. A method based on ion-chromatography has been developed for simultaneous direct determination of Fe(II) and Fe(III). Sample pretreatment involves only conventional filtration and acidification. No interferences with the iron(II) determination were found; in determination of iron(III) the only interference observed was an artifact peak (of unknown origin) that occurred only when iron(II) was present, and had an area that was a function of the iron(II) concentration and could hence be corrected for. Solutions of iron(II) free from iron(III) can be prepared by treatment with a mixture of hydrogen and nitrogen in the presence of palladium black as catalyst, to reduce the iron(III). Photoreduction of iron(III) in acidified samples increases the Fe(II)/Fe(III) ratio; no means of circumventing this effect is known, other than storing the samples in the dark and analysing them as soon as possible.  相似文献   

10.
A new sample preparation procedure to determine aminopolycarboxylic acids (ethylenediaminetetraacetic acid, EDTA, nitrilotriacetic acid, NTA, diethylenetriaminepentaacetic acid, DTPA, and cyclohexanediaminetetraacetic acid, CDTA) in river water is described. The procedure consists of the solid-phase extraction of the aminopolycaroxyllic acids on activated charcoal cartridges after increasing the ionic strength and acidifying the sample. The extract was eluted with methanol and the analytes were methylated in presence of BF3/methanol to determine them by GC with mass spectrometric detection. Recoveries were higher than 90% with good repeatabilities and inter-day precision for concentrations close to quantification limits (about 10 μg L−1) and higher. It has been verified that the proposed method is robust according to the Youden and Steiner test and free of matrix effects arisen from the presence of organic matter and iron(III) as deduced from statistical tests. A bottom-up approach was followed to estimate the uncertainty of the measured concentration. At concentrations close to 10 μg L−1 the most relevant step of the method is the calculus of the interpolated concentration which has a high value of relative standard uncertainty.  相似文献   

11.
The kinetics and mechanism of the substitution of coordinated water in nitrilotriacetate complexes of iron(III) (Fe(NTA)(OH(2))(2) and Fe(NTA)(OH(2))(OH)(-)) by phosphate (H(2)PO(4)(-) and HPO(4)(2)(-)) and acetohydroxamic acid (CH(3)C(O)N(OH)H) were investigated. The phosphate reactions were found to be pH dependent in the range of 4-8. Phosphate substitution rates are independent of the degree of phosphate protonation, and pH dependence is due to the difference in reactivity of Fe(NTA)(OH(2))(2) (k = 3.6 x 10(5) M(-)(1) s(-)(1)) and Fe(NTA)(OH(2))(OH)(-) (k = 2.4 x 10(4) M(-)(1) s(-)(1)). Substitution by acetohydroxamic acid is insensitive to pH in the range of 4-5.2, and Fe(NTA)(OH(2))(2) and Fe(NTA)(OH(2))(OH)(-) react at equivalent rates (k = 4.2 x 10(4) and 3.8 x 10(4) M(-)(1) s(-)(1), respectively). Evidence for acid-dependent and acid-independent back-reactions was obtained for both the phosphate and acetohydroxamate complexes. Reactivity patterns were analyzed in the context of NTA labilization of coordinated water, and outer-sphere electrostatic and H-bonding influences were analyzed in the precursor complex (K(os)).  相似文献   

12.
Iron uptake and toxicity in Caco-2 cells   总被引:2,自引:0,他引:2  
The differences between the in vitro effects of iron attributed to valence, chelation, and complexation are known in terms of markers of oxidative stress. Few studies, however, describe the effects of iron on general markers of toxicity used in the testing of cell cultures. The aim of the present study was to determine the toxicity and uptake of different salts and iron complexes in the human intestinal cell line, Caco-2.Cells were incubated with 1.5 mM of different species of iron [FeCl3/nitrilotriacetic acid (NTA) (1:2), FeCl3/citric acid (1:2), FeCl3 and FeSO4] for 22–24 h. Thereafter, toxicological and uptake experiments were performed.The iron uptake, viability (via MTT assay), and membrane stability (via LDH release) of Caco-2 cells incubated with various iron forms differed significantly from untreated controls which showed no detrimental effects on cells and less iron uptake. The lowest signal for cell viability (MTT assay) was found after the incubation of the cells with FeCl3/citric acid, being significantly different to treatment with FeCl3, where the highest MTT signal was detected (p=0.002). No differences between the tested iron species could be found regarding cell proliferation (via serial cell counting) and viability using the trypan blue exclusion test. The lowest membrane damage (via LDH release) was registered in cells treated with FeCl3/citric acid (1:2), whereas the highest LDH release could be found in cells incubated with FeCl3/NTA (1:2). The highest intracellular iron concentration (measured via GFAAS) was detected after the treatment of Caco-2 cells with FeCl3 and FeCl3/NTA (1:2).This study substantiates the importance of the choice of complexes, as NTA seemed to enhance the toxicity of iron, while citric acid inhibited iron uptake and toxicity.  相似文献   

13.
Using a fast reversible reaction of aminopolycarboxylic acids (APCAs) into Fe(III)-APCA complexes in the presence of Fe(III) ions, seven kinds of APCAs [nitrilotriacetate (NTA), N-(2-hydroxyethyl)ethylenediamine-triacetate (HEDTA), ethylenediamine-tetraacetate (EDTA), 1,3-propanediamine-tetraacetate (PDTA), diethylenetriamine-pentaacetate (DTPA), 1,2-diaminopropane-tetraacetate (MeEDTA), and O,O'-bis(2-aminoethyl)ethyleneglycol-tetraacetate (GEDTA)] in cosmetics and synthetic detergents were separated on two reversed-phase C30 columns connected in series and detected with ultraviolet detection. Simple pretreatment, consisted of thousand times dilution of samples and addition of 100 microl of the Fe(III) solution containing 10 mM Fe(III) chloride and 0.5 M sulfuric acid to 10 ml of diluted samples, permitted the determination of APCAs in cosmetics and synthetic detergents at concentration level of 0.1 mM, except 0.3 mM for GEDTA. APCAs except GEDTA could be detected at concentration level of 0.03 mM and GEDTA could be detected at concentration level of 0.09 mM. Good recoveries (95-110%) were obtained for each APCA by the standard addition method on two diluted samples with high accuracy (RSD 0.2-9.1%). Three APCAs (EDTA, HEDTA and NTA) were detected in various concentrations in cosmetics and synthetic detergents and the other APCAs were not detected in any of the samples. This method requires no tedious pretreatment and takes only 15 min for one analysis, so it is useful for determination of APCAs.  相似文献   

14.
This study describes a fast and accurate method for the sample preparation, identification, and quantitation of nitrilotriacetic (NTA) acid in environmental aqueous samples at a concentration of ppb level. The method is sensitive, specific, and free from the interferences of fatty and amino acids. The tri-n-propyl- and tri-n-butyl-NTA acid esters were prepared by the reaction of n-propyl-HCl and n-butyl-HCl solutions and NTA acid, respectively. The derivatives were analyzed by a gas chromatograph equipped with a mass spectrometric detector. The method detection limit, 0.006 mg/L of each NTA ester, was determined and validated by an analysis of a fortified water sample. The overall recoveries were 103-115%, n = 8. The method was applied to a real sample and a 0.90 mg/L concentration of NTA acid was found. Mass fragmentation patterns of the derivatives are also reported.  相似文献   

15.
Summary The phototransformation of iron(III) nitrilotriacetate, Fe(NTA), was studied at 20 °C under monochromatic excitation at different pHs. The conjugation of excitation wavelength and pH gives rise to different photochemical behaviour. In acidic medium, it always results in a redox process giving rise to FeII, HCHO and CO2 but the stoichiometry of the photoproducts depends on the excitation wavelength. At long wavelength (365 nm), the FeII/HCHO ratio of unity implies a redox reaction between FeIII and the carboxylic group whereas at short wavelength (254 nm) the Fe/HCHO ratio is equal to 2 and implies a redox process between FeIII and a water ligand. In neutral solution and at 365 nm, a photosolvation is observed with NTA release; at 254 nm a subsequent redox process between OH and the hydrous ferric oxide is involved. In terms of the fate of Fe(NTA) in the environment at pH 5–6 and under sunlight, all of the above photochemical reactions can occur.  相似文献   

16.
《Electroanalysis》2005,17(23):2129-2136
The investigation of the dissolved iron(III)–nitrilotriacetate–hydroxide system in the water solution (I=0.1 mol L?1 in NaClO4; pH 8.0±0.1) using differential pulse cathodic voltammetry, cyclic voltammetry, and sampled direct current (DC) polarography, was carried out on a static mercury drop electrode (SMDE). The dissolved iron(III) ion concentrations varied from 2.68×10?6 to 6×10?4 mol L?1 and nitrilotriacetate concentrations were 1×10?4 and 5×10?4 mol L?1. By deconvoluting of the overlapped reduction voltammetric peaks using Fourier transformation, four relatively stable, dissolved iron(III) complex species were characterized, as follows: [Fe(NTA)2]3?, mixed ligand complexes [FeOHNTA]? and [Fe(OH)2NTA]2?, showing a one‐electron quasireversible reduction, and binuclear diiron(III) complex [NTAFeOFeNTA]2?, detected above 4×10?4 mol L?1 of the added iron(III) ions, showing a one‐electron irreversible reduction character. The calculations with the constants from the literature were done and compared with the potential shifts of the voltammetric peaks. Fitting was obtained by changing the following literature constants: log β2([Fe(NTA)2]3?) from 24 to 27.2, log β1([FeNTA]?) from 8.9 to 9.2, log β2([Fe(NTA)2]4?) from 11.89 to 15.7 and log β2([Fe(OH)2NTA]3?) from 15.63 to 19. The determination of the electrochemical parameters of the mixed ligand complex [FeOHNTA]?, such as: transfer coefficient (α), rate constant (ks) and formal potential (E°') was done using a sampled DC polarography, and found to be 0.46±0.05, 1.0±0.3×10?3 cm s?1, and ?0.154±0.010 V, respectively. Although known previously in the literature, these four species have now for the first time been recorded simultaneously, i.e. proved to exist simultaneously under the given conditions.  相似文献   

17.
Selective spectrophotometric determinations of milligrain amounts of chromium(III) with complexans are described, based on the fact that the chromium(IIl) complexes are formed rapidly at boiling temperatures, but very slowly at room temperature, while the formation of some interfering complexes takes place instantaneously. Determinations with EDTA are more sensitive, but the combined presence of cobalt and other metals still interferes; there is no interference with the less sensitive NTA. The combined presence of a l00-fold amount of copper, nickel, cobalt and iron generally has no effect on the results. The use of DCTA, DTPA and HEDTA is discussed.  相似文献   

18.
Karadakov BP  Venkova DI 《Talanta》1970,17(9):878-883
The reaction between bismuth(III) and nitrilotriacetic acid (NTA or H(3)X) has been investigated by ultraviolet spectrophotometry. It has been established that bismuth(III) and NTA form two complexes with compositions bismuth(III): NTA = 1:1 and 1:2. The absorption maxima are at 243 nm (1:1) and 271 nm (1:2), the molar absorptivities being 8.00 x 10(3) and 8.20 x 10(3) l.mole(-1).cm(-1) respectively. The stability constants (at mu = 1.0) are: log beta(BiX) = 17.53 +/- 0.06 and log beta(B)(2)(3-) = 26.56 +/- 0.07. The possibility of the analytical application of BiX is briefly discussed.  相似文献   

19.
The extraction of iron(III) from aqueous phosphoric acid was studied using di-(2-ethylhexyl)phosphoric acid and trioctylphosphine oxide in nonaromatic hydrocarbon diluent. Distribution ratios have been investigated as a function of concentration of iron(III), phosphoric acid concentration, extractant concentration and extraction temperature. The apparent enthalpy change for the extraction reaction has been calculated from the temperature dependence data. It was found that the extractant dependency for iron(III) is first power indicating hydrolysis of iron(III) in the aqueous phase.  相似文献   

20.
As a continuation of recent mechanistic studies into the influence of seawater ligands on the surface chemistry of the iron chalcogenide glass membrane ion‐selective electrode (ISE), the present study has investigated the response of the iron(III) ISE in a seawater ligand mimetic system to examine its suitability as a calibration medium for the electroanalysis of raw or natural seawater. Significantly, dip method calibrations of the ISE in a mixture of salicylate, ethylene diamine tetraacetic acid (EDTA), ethylene diamine and minor amounts of dissolved iron(III) and copper(II) yielded the expected Nernstian response of 30 mV/decade according to the known ion‐exchange/electron transfer response mechanism of this ISE. Furthermore, ideal Nernstian response of the electrode is also obtained in a continuous flow analysis (CFA) mode, noting that this provides scope for using a hydrodynamic flow regime to minimize the electrode release of iron and the concomitant detection limit of the ISE. Ultimately, repetitive CFA analyses of free iron(III) in raw or natural seawater yielded a free iron(III) level commensurate with the expected inorganic and organic speciation of iron(III) in seawater.  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