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.
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.
Matsuda T  Nagai T 《Talanta》1983,30(12):951-954
The stoichiometry of the oxidation of IDA or NTA with lead dioxide suspension was studied by polarographic measurement and by derivative polarographic titration. One mole and two moles of Pb(IV) are reduced per mole of IDA and NTA respectively, with moderate speed at room temperature in nitric acid solutions. One mole each of carbon dioxide, formaldehyde and glycine are produced from the oxidation of 1 mole of IDA, and two moles of carbon dioxide, two moles of formaldehyde and one mole of glycine from 1 mole of NTA. The overall reaction in each case may be written as follows: Pb(IV) + IDA + H2O → Pb(II) + CO2 + HCHO + H2NCH2COOH + 2H+ 2Pb(IV) + NTA + 2H2O → 2Pb(II) + 2CO2 + 2HCHO + H2NCH2COOH + 4H+  相似文献   

3.
The extraction rates of cadmium(II) and zinc(II) with dithizone (H2dz) in the presence of nitrilotriacetic acid (NTA) were measured, and the possible kinetic separation of cadmium(II) from zinc(II) was investigated. Upon the addition of NTA, the difference in the extraction rate between cadmium(II) and zinc(II) became large. Based on the observed rate constant under the condition [NTA] = 1 x 10(-2) mol dm-3, [H2dz]org = 1 x 10(-3) mol dm-3, and pH = 7.0, the shaking time required for the quantitative separation of cadmium(II) from zinc(II) was calculated to be between 326 and 995 s. The experimental results agreed with the prediction, and the quantitative separation of cadmium(II) from zinc(II) was performed within the above-mentioned range of shaking times.  相似文献   

4.
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   

5.
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   

6.
Horácek J  Pribil R 《Talanta》1969,16(11):1495-1499
Consecutive potentiometric determination of EDTA and NTA or DTPA and NTA is described. The method is based on the titration with iron(III) chloride at pH 4-5 in the presence of a small amount of ferroin and 1,10-phenanthroline which keeps the concentration of free iron(II) ions constant, thus making the potential dependent only on the free iron(III) concentration.  相似文献   

7.
建立了用氢溴酸消除锑、砷、锡干扰,用硫酸将铅形成硫酸铅沉淀,再用EDTA络合滴定法测定粗二氧化碲中铅量的方法。试样用硝酸、盐酸溶解,用硫酸沉淀铅,氢溴酸消除锑、砷、锡的干扰后,过滤分离其他共存元素,以乙酸-乙酸钠缓冲溶液溶解硫酸铅沉淀,在pH=5.0~6.0时,以二甲酚橙作指示剂,用Na_2EDTA溶液滴定溶液中铅含量。实验结果表明,氢溴酸加入量为15mL,酒石酸加入量为10mL,沉淀体积为50~60mL,沉淀时间1h以上时,方法相对标准偏差(RSD)在0.10%~1.1%,加标回收率为97.1%~102%,满足粗二氧化碲中铅量的生产控制检测要求。  相似文献   

8.
The stability of 5-aminosalicylic acid and its metabolites has been investigated when stored frozen. N-beta-D-Glucopyranosyl-5-aminosalicylic acid was formed in considerable amounts concomitant with a decrease in 5-aminosalicylic acid in plasma samples spiked with 5-aminosalicylic acid as well as in standard solutions of 5-aminosalicylic acid buffered with potassium phosphate between pH 5.5 and pH 8.0 with 4.0 mM glucose added and stored at -20 degrees C. Thus N-beta-D-glucopyranosyl-5-aminosalicylic acid might not, as previously described, be a metabolite of 5-aminosalicylic acid but an artifact formed during storage of plasma samples. The N-glucoside formed could be quantitatively degraded to 5-aminosalicylic acid and glucose by adding 0.2 M potassium phosphate buffer pH 3.0 to the sample prior to the analysis. The metabolites of 5-aminosalicylic acid (N-formyl-5-aminosalicylic acid, N-acetyl-5-aminosalicylic acid and N-butyryl-5-aminosalicylic acid) were found to be stable in plasma stored at -20 degrees C for at least eight months.  相似文献   

9.
称取0.5 g试样,用盐酸、硝酸、硫酸分解,铅以硫酸铅沉淀的形式分离。以过硫酸铵为氧化剂,氨性溶液中沉淀分离铁、锰等共存元素,加氟化物掩蔽铝、硫脲掩蔽铜,调节溶液pH=5.0~6.0,加入碘化钾消除镉的干扰,以二甲酚橙为指示剂,用EDTA标准溶液滴定。将沉淀物重新溶解,用原子吸收光谱法测定其中的锌量。在0.2%~10.0%范围内结果重现性良好,重复测定7次,相对标准偏差(RSD)在0.32%~0.76%,样品加标回收率为99.4%~100%。方法的准确度和精密度可以满足分析需求。  相似文献   

10.
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   

11.
By applying an internal standardization, we could use a rapid coprecipitation technique using lanthanum phosphate as a coprecipitant for preconcentration of iron(III) and lead in their flame atomic absorption spectrometric determination. Indium as an internal standard was added to the initial sample solution together with lanthanum and phosphoric acid; the coprecipitation of iron(III) and lead was then carried out at pH about 3. After measuring the atomic absorbances of iron, lead, and indium in the final sample solution, we determined the contents of iron(III) and lead in the original sample solution by using the internal standardization with indium. In this method, complete collection of the precipitate was not required after the coprecipitation of iron(III), lead, and indium, because the ratio of the recovery of iron(III) or lead to that of indium was almost constant regardless of the recovery of the precipitate. This method was simple and rapid, and was available for the determination of 2-300 micrograms L-1 of iron(III) and 5-400 micrograms L-1 of lead in some water samples.  相似文献   

12.
The equilibria and kinetics for the process of In(3+) exchange between nitrilotriacetic acid (NTA) and bovine serum transferrin (T) have been investigated in aqueous solution containing sodium bicarbonate. The metal exchange equilibria have been measured by difference ultraviolet spectroscopy at 25 degrees C, pH=7.4, and I=0.2 M (NaClO4). The acid dissociation constants of NTA and the binding constants of In(III) to NTA have also been measured. Kinetic experiments revealed that the process of In(3+) uptake by transferrin from [In(NTA)2](3-) is biphasic, the fast phase being completed in a few seconds, the slow phase lasting for hours. The fast phase has been investigated by the stopped-flow method and results in monoexponential kinetics. It involves rapid interaction of the 1:1 complex ML (M=In, L=NTA) with TB (T=transferrin, B=CO3(2-)) to give a quaternary intermediate MLTB which then evolves to an "open" MTB* ternary complex complex with expulsion of L. In turn, this complex interconverts to a "closed", more stable, form MTB. Neither the prevailing complex M2L nor the TB2 form of transferrin are directly involved in the exchange process but act as metal and protein reservoirs. The pH dependence of the reaction has been also investigated. The slow phase has not been investigated in detail; it takes several hours to go to the completeness, its slowness being ascribed to metal redistribution between the C-site and N-site of the protein, and/or metal release from polynuclear In(III) species.  相似文献   

13.
乙醇—水混合溶剂中次氮基三乙酸离解平衡的研究   总被引:4,自引:0,他引:4  
次氮基三乙酸(NTA)在非水或混合溶剂中离解常数的测定尚未见报道,本文用电位法测定了乙醇-水混合溶剂中溶解度较小的NTA的离解常数,用标准盐酸溶液返滴定其钾盐可增加NTA浓度,提高测定准确度,测量系统防止CO_2装置和数据处理方法也比文献简单。  相似文献   

14.
Mizoguchi T  Ishii H 《Talanta》1979,26(1):33-39
Bauxites can be decomposed by condensed phosphoric acid (CPA) very rapidly without the need for subsequent manipulations such as elimination of silica, digestion of fused products and filtration. It is best to heat the samples at about 700 degrees prior to the decomposition, to prevent them from floating on the surface of the CPA. Under the proposed conditions (100 mg of sample, 10 g of CPA, heating at 300 degrees for 30 min), aluminium, iron and titanium are dissolved quantitatively. Iron is determined by photometry with 1,10-phenanthroline after solvent extraction with MIBK, while titanium is determined with N-benzoyl-N-phenylhydroxylamine (BPHA). The effect of phosphate on the determination of titanium is reduced to a minimum at a BPHA concentration of 0.3% and a hydrochloric acid concentration of 7.2M. Aluminium and iron are precipitated quantitatively as the oxinates at pH 5.5 in the presence of orthophosphoric acid or hydrolysed CPA, while the precipitation of titanium oxinate is completely suppressed by the addition of hydrogen peroxide. The total amount of aluminium and iron is obtained by determining the amount of oxine by bromination method. The amount of aluminium is obtained by subtracting the amount of iron from the sum of the two.  相似文献   

15.
Flow field–flow fractionation–inductively coupled plasma optical emission spectrometry (FlFFF–ICP–OES) was applied to achieve the size-based fractionation of iron in a food suspension in order to gain insights into iron availability. The binding of iron with phytic and tannic acids, employed as model inhibitors of iron availability in foods, was investigated at pH 2.0 (representing stomach fluid), pH 5.0 (the transition stage in the upper part of the duodenum), and pH 7.0 (the small intestine). In the presence of phytic acid, iron was found as a free ion or it was associated with molecules smaller than 1 kDa at pH 2.0. Iron associated with molecules larger than 1 kDa when the pH of the mixture was raised to 5.0 and 7.0. In the presence of tannic acid, iron was again mostly associated with molecules smaller than 1 kDa at pH 2.0. However, at pH 5.0, iron and tannic acid associated in large molecules (∼25 kDa), while at pH 7.0, most of the iron was associated with macromolecules larger than 500 kDa. Iron size-based distributions of kale extract and tea infusion containing phytic and tannic acids, respectively, were also examined at the three pH values, with and without enzymatic digestion. Without enzymatic digestion of the kale extract and the tea infusion at pH 2.0, most of the iron was released as free ions or associated with molecules smaller than 1 kDa. At other pH values, most of the iron in the kale extract and the tea infusion was found to bind with ~2 kDa and >500 kDa macromolecules, respectively. Upon enzymatic gastrointestinal digestion, the iron was not observed to bind to macromolecules >1 kDa but <500 kDa, due to the enzymatic breakdown of large molecules to smaller ones (<1 kDa). Figure Flow field–flow fractionation was exploited in order to achieve size-based iron fractionation and thus investigate iron-binding behavior under gastrointestinal conditions  相似文献   

16.
复杂高铋物料中,铋、砷、锑、锡四元素含量高且共存时会影响铅的测定。特别是铋含量高时对铅的测定影响大。实验用EDTA—酒石酸联合掩蔽铋、砷、锑、锡,在稀硫酸介质中以硫酸钾为沉淀剂,使铅生成硫酸铅钾复盐沉淀而与铋、砷、锑、锡、铁、铜、锌、铝、钴、镍等干扰离子分离,沉淀以乙酸-乙酸钠浸取,二甲酚橙为指示剂,Na2EDTA滴定法测定铅。试验进一步优化了测定条件,确定最佳条件:硫酸(1 1)加入量为7mL、硫酸钾用量为5g、煮沸时为5min、沉淀陈化时间为2h、EDTA 50g/L 加入量为10mL、酒石酸用量为0.5g,铅的回收率99.70% ~100.65%。将实验方法应用于测定复杂高铋物料中铅,标样BY0111-1与给定值一致,相对标准偏差(n=11)RSD 0.20%~0.23%,满足生产测试要求。  相似文献   

17.
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   

18.
5‐Hydroxymethylfurfural (5‐HMF) is a natural occurring substance taken up by everyday food. In former studies it was shown that 5‐HMF is completely decomposed in the body after oral or intravenous application resulting in three main metabolites named 5‐hydroxymethylfuroic acid, 2,5‐furandicarboxylic acid, and N‐(hydroxymethyl)furoyl glycine, and possibly a forth metabolic substance, termed 5‐sulphoxymethylfurfural, is formed. Determination is possible via HPLC using a hydrophilic interaction chromatography (HILIC) column with an appropriate gradient system (ACN/ammonium formate 100 mM, pH 2.35). Urine samples were purified by use of an SPE method beforehand working with ScreenA cartridges. This cleaning procedure was validated based on ICH guidelines in terms of linearity, quantification, and detection limit, as well as precision, repeatability, and accuracy. Analysis of real‐life samples coming from two healthy probands and one cancer patient, who all received 240 mg 5‐hydroxymethylfurfural orally once a day, showed dicarboxylic acid and the glycine conjugate in their urine samples. Recovery of the initial compound in form of transformed metabolites was up to 90% within 48 h. Potentially toxic 5‐sulphoxymethylfurfural could not be found.  相似文献   

19.
Summary By using an electrodeposited lead dioxide electrode as an indicator electrode, the electrode reaction of ethylenediaminetetraacetic acid (EDTA) and other complexanes was investigated in unbuffered solution covering a pH range of about 2–12. Well-defined anodic and cathodic peaks, the heights of which increase with EDTA concentration, were observed. In an anodic single-sweep process, one anodic pre-peak appeared only at a pH higher than 4. It was found that in the cathodic process lead dioxide on the electrode is reduced to lead ion which reacts with EDTA to form the complex and in the anodic process the lead complex is oxidized again and deposited onto the electrode as lead dioxide. The height of the anodic and cathodic peaks is directly proportional to the EDTA concentration. The lead dioxide electrode seems to be useful for determining a chelating agent such as the EDTA-type complexanes.
Untersuchung der Elektrodenreaktion von EDTA und anderen Komplexanen in wäßriger Lösung an der Bleidioxidelektrode
Zusammenfassung Mit Hilfe der elektrolytisch belegten PbO2-Elektrode wurde die Elektrodenreaktion von EDTA (sowie von IDA, NTA und DTPA) in ungepufferter Lösung im pH-Bereich 2–12 untersucht. Gut ausgebildete anodische und kathodische Peaks wurden beobachtet, deren Höhe mit der EDTA-Konzentration zunimmt. Im anodischen single-sweep-Prozeß erschien nur ein anodischer Vorpeak bei pH>4. Es wurde festgestellt, daß bei dem kathodischen Vorgang PbO2 an der Elektrode zu Pb-Ion reduziert wird, das mit EDTA unter Komplexbildung reagiert, während an der Anode der Pb-Komplex wieder oxidiert und PbO2 abgeschieden wird. Die Peakhöhen sind der EDTA-Konzentration direkt proportional. Die PbO2-Elektrode erscheint daher von Nutzen für die Bestimmung von Komplexanen vom Typ des EDTA.
  相似文献   

20.
近年来,杂环偶氮苯甲酸类显色剂取得了较快的发展。由于氨基取代结构对试剂性能有较大的影响[1],因此本文合成了新显色剂2-(2-噻唑偶氮)-5-二乙氨基苯甲酸,(TAEB)并研究了它与钴的显色反应。  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