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.
Composite diazotization-coupling reagents containing sulfanilamide (SAM), sulfapyridine (SP) or sulfathiazole (ST) as the diazotizable aromatic amines and sodium 1-naphthol-4-sulfonate (NS) as the coupling agent using column preconcentration on naphthalene-tetradecyldimethylbenzylammonium(TDBA)-iodide adsorbent have been used for the spectrometric determination of trace nitrate and nitrite in soil and water samples. Nitrite ion reacts with SAM in the pH range 2.0–5.0, SP in the pH range 2.0–2.5 and ST in the pH range 2.0–3.3 in HCl medium to form water-soluble colourless diazonium cations. These cations were coupled with NS in the pH range 9.0–12.0 for the SAM system, 9.6–12.0 for the SP system and 8.5–12.0 for the ST system to be retained on naphthalene-TDBA-I material packed in a column. The solid mass is dissolved from the column with 5 ml of dimethylformamide and the absorbance is measured spectrometerically at 543 nm for SAM-NS, 533 nm for SP-NS and 535 nm for ST-NS. Nitrate is reduced to nitrite by a copper-coated cadmium reductor column and the nitrite is then treated with the diazotization-coupling reagent by column preconcentration. The absorbance due to the sum of nitrate and nitrite is measured and nitrate is determined by difference. The calibration graph was linear over the range 2–40 ng NO2-N ml−1 and 1.5–30 ng NO3-N ml−1 in aqueous samples for the SAM and ST systems and 2–48 ng NO2-N ml−1 and 1.5–36 ng NO3-N ml−1 in aqueous samples for the SP system, respectively. The sensitivity, accuracy and precision of the systems decreased in the order STSAMSP. The detection limits were 1.4 ng NO2-N ml−1 and 1.1 ng NO3-N ml−1 for SAM, 1.6 ng NO2-N ml−1 and 1.2 ng NO3-N ml−1 for SP, and 1.0 ng NO2-N ml−1 and 0.75 ng NO3-N ml−1 for ST, respectively. The preconcentration factors are 8, 5 and 6 for SAM-NS, SP-NS and ST-NS, respectively. Interferences from various foreign ions have been studied and the methods have been applied to the determination of ng ml−1 levels of nitrite and nitrate in soil and water samples. The mean recovery was 95–102% for all three systems.  相似文献   

2.
Ohura H  Imato T  Yamasaki S 《Talanta》1999,49(5):1383-1015
A rapid potentiometric flow injection technique for the simultaneous determination of oxychlorine species such as ClO3–ClO2 and ClO3–HClO has been developed, using both a redox electrode detector and a Fe(III)–Fe(II) potential buffer solution containing chloride. The analytical method is based on the detection of a large transient potential change of the redox electrode due to chlorine generated via the reaction of the oxychlorine species with chloride in the potential buffer solution. The sensitivities to HClO and ClO2 obtained by the transient potential change were enhanced 700–800-fold over that using an equilibrium potential. The detection limit of the present method for HClO and ClO2 is as low as 5×10−8 M with use of a 5×10−4 M Fe(III)–1×10−3 M Fe(II) buffer containing 0.3 M KCl and 0.5 M H2SO4. On the other hand, sensitivity to ClO3 was low when a potential buffer solution containing 0.5 M H2SO4 was used, but could be increased largely by increasing the acidity of the potential buffer. The detection limit for ClO3 was 2×10−6 M with the use of a 5×10−4 M Fe(III)–1×10−3 M Fe(II) buffer containing 0.3 M KCl and 9 M H2SO4. By utilizing the difference in reactivity of oxychlorine species with chloride in the potential buffer, a simultaneous determination method for a mixed solution of ClO3–ClO2 or ClO3–HClO was designed to detect, in a timely manner, a transient potential change with the use of two streams of potential buffers which contain different concentrations of sulfuric acid. Analytical concentration ranges of oxychlorine species were 2×10−5–2×10−4 M for ClO3, and 1×10−6–1×10−5 M for HClO and ClO2. The reproducibility of the present method was in the range 1.5–2.3%. The reaction mechanism for the transient potential change used in the present method is also discussed, based on the results of batchwise experiments. The simultaneous determination method was applied to the determination of oxychlorine species in a tap water sample, and was found to provide an analytical result for HClO, which was in good agreement with that obtained by the o-tolidine method and to provide a good recovery for ClO3 added to the sample.  相似文献   

3.
The three cyanocuprate(I) complexes, Cu(CN)2, Cu(CN)32−, and Cu(CN)43−, photoeject electrons with high efficiency when excited in aqueous solution by 266 nm laser pulses of 7 ns duration with quantum yields of 0.37±0.06, 0.224±0.021, and 0.240±0.005, for Cu(CN)2 (at 2 M ionic strength), Cu(CN)32−, and Cu(CN)43− (both measured at 1 M ionic strength). Along with hydrated electrons, two transient intermediates, absorbing at 460 and 340 nm, respectively, form consecutively after excitation through bimolecular reactions with ground-state Cu(I) in solutions of Cu(CN)2, and Cu(CN)32−, but not in Cu(CN)43−. All photoprocesses are essentially monophotonic. A mechanism is proposed that suggests the formation of a dinuclear excited-state complex such as an excimer.  相似文献   

4.
A simple and rapid flow injection (FI) method is reported for the determination of phosphate (as molybdate reactive P) in freshwaters based on luminol chemiluminescence (CL) detection. The molybdophosphoric heteropoly acid formed by phosphate and ammonium molybdate in acidic conditions generated chemiluminescence emission via the oxidation of luminol. The detection limit (3× standard deviation of blank) was 0.03 μg P l−1 (1.0 nM), with a sample throughput of 180 h−1. The calibration graph was linear over the range 0.032–3.26 μg P l−1 (r2=0.9880) with relative standard deviations (n=4) in the range 1.2–4.7%. Interfering cations (Ca(II), Mg(II), Ni(II), Zn(II), Cu(II), Co(II), Fe(II) and Fe(III)) were removed by passing the sample through an in-line iminodiacetate chelating column. Silicate interference (at 5 mg Si l−1) was effectively masked by the addition of tartaric acid and other common anions (Cl, SO42−, HCO3, NO3 and NO2) did not interfere at their maximum admissible concentrations in freshwaters. The method was applied to freshwater samples and the results (26.1±1.1–62.0±0.4 μg P l−1) were not significantly different (P=0.05) from results obtained using a segmented flow analyser method with spectrophotometric detection (24.4±4.45–84.0±16.0 μg P l−1).  相似文献   

5.
A calorimetric study was performed for adducts of general formula CdBr2·nL (n=1 and 2; L=ethyleneurea (eu) and propyleneurea (pu)). The standard molar reaction enthalpy in condensed phase: CdBr2(c)+nL(c)=CdBr2·nL(c); ΔrHmθ, were obtained by reaction–solution calorimetry, to give the following values for mono- and bis-adducts: −19.54 and −34.59; −7.77 and −19.05 kJ mol−1 for eu and pu adducts, respectively. Decomposition (ΔDHmθ) and lattice (ΔMHmθ) enthalpies, as well as the mean cadmium---oxygen bond dissociation enthalpy, DCd---O, were calculated for all adducts.  相似文献   

6.
Formation constants for recrystallized thymol blue were determined in water, using the SQUAD and SUPERQUAD programs. The best model correlating spectrophotometric, potentiometric and conductimetric data was fitted with the dissociation of HL=L2−+H+−log K=8.918±0.070 and H3L2=2L2−+3H+−log K=29.806±0.133 with the SUPERQUAD program at variable low ionic strength (1.5×10−4–3.0×10−4 M); and HL=L2−+H+−log K=8.9±0.000, H3L2 =2L2−+3H+−log K=30.730±0.032, H4L2=2L2−+4H+−log K=32.106±0.033 with SQUAD at 1.1 M ionic strength.  相似文献   

7.
N. Miralles  A. Sastre  M. Aguilar 《Polyhedron》1987,6(12):2145-2149
The complex equilibria between HCrO4 and Cl ions has been studied spectrophotometrically at a constant ionic strength of 3.0 mol dm−3 and the data have been analyzed both graphically and numerically by means of the program LETAGROP-SPEFO (L. G. Sillen and B. Warnquist, Arkiv. kemi. 1968, 31, 377). The experimental results can be explained on the basis of the following reaction: HCrO4+H++Cl = CrO3Cl+H2O (log β11 = 1.37±0.08). Molar absorptivities of HCrO4 and CrO3Cl were also reported.  相似文献   

8.
The paper reports results of a study on the specific adsorption of F, Cl, Br, I, ClO3, BrO3, IO3 and IO4 on hydrous γ-Al2O3. The isotherms of the anion adsorption and the adsorption dependencies on pH and the ionic strength of the solution have been determined under the equilibrium conditions. According to the degree of affinity to γ-Al2O3, the anions can be ordered as: I3334−. It has been established that the sorption of IO4 and F involves the formation of surface complexes in the inner co-ordination sphere, whereas that of Cl, Br, I, ClO3, BrO3 and IO3 takes place through formation of ion pair complexes in the outer co-ordination sphere. In the dynamic system, the exchange isoplanes and elution curves have been determined for selected anions on columns filled with Al2O3. It has been shown that γ-Al2O3 can be used for isolation and concentration of IO3 from natural waters in order to decrease the limit of the ions determination to 2 μg l−1. Using differential pulse voltammetry (DPV), after isolation and concentration on γ-Al2O3, the content of iodates has been determined in mineral, marine and tap water doped with these ions.  相似文献   

9.
The anodization of mercury microelectrodes was investigated in synthetic samples containing several strong and weak electrolytes at different concentrations. In particular, the effects on mercury anodization due to the presence of NaOH, HClO4, NaCl, NaI, NaF, Na2SO4, NaHCO3, Na2CO3, tartaric and citric acids, were studied in solutions containing either each species or mixtures of them, and without addition of supporting electrolyte. Some of the electrode processes studied led to linear calibration plots e.g. 1 × 10−5 − 1 × 10−4M Cl, 1 × 10−6 − 1 × 10−5M I, 5 × 10−4 − 3 × 10−3M SO42−, 5 × 10−4 − 2 × 10−2M HCO3, with typical correlation coefficients of 0.998–0.999. The anodization of mercury microelectrodes was also investigated directly in wine, rain, tap and mineral water, without pretreatment and without addition of supporting electrolyte. In the real samples only the ions Cl and HCO3 could be quantified, and the values found were in agreement, within 3–5%, with the reference values obtained by using Italian standard methods for food.  相似文献   

10.
The new iodoammonium salts o-C6H4(NH2)2I+I (1) and o-C6H4(NH2)2I+ AsF6 (2) were prepared by reaction of o-phenylene diamine with I2 or I3+AsF6, respectively. Compound 1 reacts with AlI3 yielding quantitatively the corresponding tetraiodoaluminate o-C6H4(NH2)2I+AlI4 (3). The species were characterized by chemical analysis, vibrational (IR and Raman) and temperature-dependent 1H NMR spectropscopy. Direct evidence for a N---I bond was found in the Raman spectra of 1, 2 and 3 (ν(NI) = 599–600 cm−1).  相似文献   

11.
Burakham R  Oshima M  Grudpan K  Motomizu S 《Talanta》2004,64(5):1259-1265
A novel spectrophotometric reaction system was developed for the determination of nitrite as well as nitrate in water samples, and was applied to a flow-injection analysis (FIA). The spectrophotometric flow-injection system coupled with a copperised cadmium reductor column was proposed. The detection was based on the nitrosation reaction between nitrite ion and phloroglucinol (1,3,5-trihydroxybenzene), a commercially available phenolic compound. Sample injected into a carrier stream was split into two streams at the Y-shaped connector. One of the streams merged directly and reacted with the reagent stream: nitrite ion in the samples was detected. The other stream was passed through the copperised cadmium reductor column, where the reduction of nitrate to nitrite occurred, and the sample zone was then mixed with the reagent stream and passed through the detector: the sum of nitrate and nitrite was detected. The optimised conditions allow a linear calibration range of 0.03–0.30 μg NO2-N ml−1 and 0.10–1.00 μg NO3-N ml−1. The detection limits for nitrite and nitrate, defined as three times the standard deviation of measured blanks are 2.9 ng NO2-N ml−1 and 2.3 ng NO3-N ml−1, respectively. Up to 20 samples can be analyzed per hour with a relative standard deviation of less than 1.5%. The proposed method could be applied successfully to the simultaneous determination of nitrite and nitrate in water samples.  相似文献   

12.
A new method has been developed for ion-interaction chromatography with suppressed conductivity detection and a new graphitized carbon packing, which is sintered from carbonic material at a high temperature. Combinations of various eluting agents, tetrabutylammonium hydroxide (TBA) and acetonitrile have been investigated to optimize the separation of eight common anions (F, Cl, NO2, Br, NO3, SO42−, HPO42− and I). Calibration curves were linear from 0.5 to 10 μg/ml for F, from 1.0 to 20 μg/ml for Cl, NO2 and NO3, from 2.5 to 50 μg/ml for Br and SO42− and from 5.0 to 100 μg/ml for HPO42− and I with a correlation coefficient (r) of 0.999 or better. The relative standard deviations (R.S.D.s) of peak areas were between 0.2 and 0.9% for 10 repeated measurements. The application of this newly developed method was demonstrated by the determination of chloride, bromide and sulfate in pharmaceutical compounds using the direct injection method. The analytical results were within ±2% (relative) of the theoretical value, and thus in good agreement with the theoretical value for each sample.  相似文献   

13.
Pulse radiolysis technique has been employed to study the reactions of oxidizing (OH, N3) and reducing radicals (eaq, CO2√−, acetone ketyl radical) with 2-hydroxy-3-methoxybenzaldehyde (o-vanillin) at different pH. Hydroxyl radicals react mostly by addition reaction forming radical adducts (λmax=420 nm) and the oxidation is only a minor process even in the alkaline region. The reaction with azide radicals produced phenoxyl radicals (λmax=340 nm), which are formed on fast deprotonation of solute radical cation. Using PMZ√+/PMZ and ABTS√−/ABTS2− as the reference couple, different methods are employed to determine the one-electron reduction potential of o-vanillin and the average value is estimated to be 1.076±0.004 V vs. NHE at pH 6. The phenoxyl radicals of o-vanillin were able to oxidize ABTS2− quantitatively. The eaq is observed to react with o-vanillin with rate constant value of 2×1010 dm3 mol−1 s−1. CO2√− and acetone ketyl radical are also observed to react with o-vanillin by electron transfer mechanism and showed the formation of transient absorption bands with λmax at 350 and 390 nm at pH 4.5 and 9.7, respectively. The pKa of the one-electron reduced species was determined to be 8.1. The results indicate that the aldehydic group is the most preferred site for electron addition.  相似文献   

14.
De Marco R  Phan C 《Talanta》2003,60(6):1215-1221
The direct flow injection potentiometric (FIP) analysis of phosphate in hydroponic nutrient solution has been carried out using a cobalt-wire ion-selective electrode (ISE). Synthetic hydroponic nutrient solution, commercial hydroponic nutrient solution and working hydroponic farm nutrient solution were analysed for phosphate using the FIP technique. It is shown that FIP results compare favourably to standard methods of analysis such as spectrophotometry and indirect photometric ion-pair chromatography. Reproducible FIP response curves with a slope of −(47.57±0.03) mV per decade and intercept of −(169.7±0.1) mV were obtained for four separate calibrations in the concentration range 5.0×10−4–1.0×10−2 M H2PO4. Anion corrections for interferences by Cl, NO3 and SO42− were applied to all samples using the selectivity coefficients determined independently using a fixed interference method. Nevertheless, it was found that anion corrections were not necessary, as the deviations fell within the bounds of experimental error for the cobalt-wire ISE technique (i.e.±2–5% R.S.D.). The proposed FIP method enables the direct determination of phosphate in hydroponic nutrient solutions.  相似文献   

15.
The self-termination rates of the benzyl radical (C6H5---CH2) and para-substituted benzyl radicals (X---C6H4---CH2) were studied in aqueous solutions. The Arrhenius parameters and activation energies were determined in the temperature range 275.5–328 K. The kinetic activation energies of these radicals were close to the dynamic activation energy of the solvent, indicating that the termination rate is controlled by diffusion. The values for the rate constants (2kt (109 dm3 mol−1 s−1)) and the activation energies (E (kJ mol−1)) were 5.94±0.52 and 14.69±0.61 for CH3O---C6H4---CH2, 4.52±0.2 and 17.65±1.16 for CH37z.sbnd;C6H4---CH2, 3.07±0.45 and 17.58±0.97 for H---C6H4---CH2, 4.13±0.81 and 19.10±1.20 for Cl---C6H4---CH2 and 4.17±0.44 and 14.62±0.52 for NO2---C6H4---CH2.  相似文献   

16.
Reactions of OH radicals and some one-electron oxidants with 2-aminopyridine (2-AmPy) and 3-aminopyridine (3-AmPy) were studied in aqueous solutions using pulse radiolysis technique. The OH adduct of 2-AmPy at pH 9 has an absorption maximum at 360 nm along with a weak absorption band in the visible region and was found to be reactive with oxygen. The rate constant for its reaction with O2 was determined to be 1.0×108 dm3 mol−1 s−1. At pH 4 also, the OH adduct of 2-AmPy has an absorption band at 360 nm. However, there are differences in the absorption at other wavelengths. From the plot of ΔOD vs. pH at 340 nm, the pKa of the OH adduct was determined to be 6.5. Among the specific oxidants, only SO4−√ radicals were able to oxidize 2-AmPy. In the case of 3-aminopyridine (3-AmPy), the transient species formed by OH radical reaction at pH 9 has an absorption maximum at 410 nm with shoulder bands on both the sides. Its absorption spectrum at pH 4 was different indicating the existence of a pK value for the OH adduct. pKa of 3-AmPy-OH radical adduct species was evaluated to be 5.7. This adduct species was also found to be reactive with oxygen (k=7.6×106 dm3 mol−1 s−1). Specific one-electron oxidants like N3, Br2−√ C2−√ and SO4−√ were able to oxidize 3-AmPy indicating that it is easier to oxidize 3-AmPy as compared to 2-AmPy.  相似文献   

17.
The rate constant for the reaction between the sulphate radical (SO4√−) and the ruthenium (II) tris-bipyridyl dication (Ru(bipy)32+) is (3.3±0.2)×109 mol−1 dm3 s−1 in 1 mol dm−3 H2SO4 and (4.9±0.5)×109 mol−1 dm3 s−1 in 0.1 mol dm−3, pH 4.7 acetate buffer. The SO4√−radical produced by the electron transfer quenching of Ru(bipy)32+* by S2O82− reacts rapidly with both acetate buffer and chloride ions. These side reactions result in a reduction in the overall quantum yield of Ru(bipy)33+ production and reduced reaction selectivity when Ru(bipy)32+* is quenched by persulphate.  相似文献   

18.
Bismuth as BiCl4 and BH4 ware successively retained in a column (150 mm × 4 mm, length × i.d.) packed with Amberlite IRA-410 (strong anion-exchange resin). This was followed by passage of an injected slug of hydrochloric acid resulting in bismuthine generation (BiH3). BiH3 was stripped from the eluent solution by the addition of a nitrogen flow and the bulk phases were separated in a gas–liquid separator. Finally, bismutine was atomized in a quartz tube for the subsequent detection of bismuth by atomic absorption spectrometry. Different halide complexes of bismuth (namely, BiBr4, BiI4 and BiCl4) were tested for its pre-concentration, being the chloride complexes which produced the best results. Therefore, a concentration of 0.3 mol l−1 of HCl was added to the samples and calibration solutions. A linear response was obtained between the detection limit (3σ) of 0.225 and 80 μg l−1. The R.S.D.% (n = 10) for a solution containing 50 μg l−1 of Bi was 0.85%. The tolerance of the system to interferences was evaluated by investigating the effect of the following ions: Cu2+, Co2+, Ni2+, Fe3+, Cd2+, Pb2+, Hg2+, Zn2+, and Mg2+. The most severe depression was caused by Hg2+, which at 60 mg l−1 caused a 5% depression on the signal. For the other cations, concentrations between 1000 and 10,000 mg l−1 could be tolerated. The system was applied to the determination of Bi in urine of patients under therapy with bismuth subcitrate. The recovery of spikes of 5 and 50 μg l−1 of Bi added to the samples prior to digestion with HNO3 and H2O2 was in satisfactory ranges from 95.0 to 101.0%. The concentrations of bismuth found in six selected samples using this procedure were in good agreement with those obtained by an alternative technique (ETAAS). Finally, the concentration of Bi determined in urine before and after 3 days of treatment were 1.94 ± 1.26 and 9.02 ± 5.82 μg l−1, respectively.  相似文献   

19.
Leachates from different areas of a modern hazardous-waste landfill were investigated. In addition, samples of the waste to be buried were taken, and leachability tests in accordance with German standards were performed. The composition of the leachates from the landfill and the leachates produced by the leachability tests varied over a wide range, depending on the kind and volume of hazardous waste buried and the weather conditions. High concentrations of some anions were often found in combination with low concentrations of other anions. In the leachates the following concentration ranges were found: Cl, 30–6000 mg/l; NO2, 0–150 mg/l; NO2, 0–150 mg/l; SO2 2−, 100–6000 mg/l. Therefore, several dilutions of one sample often had to be measured. The complex matrix often also requires several sample preparation steps for the elimination of interfering effects. Experience to date has shown that ion chromatography in this application field is efficient.  相似文献   

20.
Tirumalesh K 《Talanta》2008,74(5):1428-1434
This study describes a new ion chromatography method using a low-capacity anion exchange column with amperometric and absorbance detection for rapid and simultaneous determination of Br and NO3 in contaminated waters where one of these ions is present in excess compared to other. The use of two detectors overcomes the problem of baseline separation for Br and NO3 for accurate quantification, which was commonly encountered when using a low-capacity anion exchange column and suppressed conductivity detection mode. The method achieved accurate quantification of these two ions without requirement of baseline separation. The accuracy of 2.8% for NO3 was determined using a quality control sample obtained from UN GEMS/Water PE Study No. 6. The detection limits for Br and NO3 were 20 and 6 μg l−1 (25 μl sample), respectively. Linearity of these two ions was over three orders of magnitude with a correlation coefficient >0.998. The influence of potential interfering ions was also studied followed by the determination of Br and NO3 in seawater, unsaturated zone water, soil extract and groundwater.  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