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 46 毫秒
1.
Complexation of aluminium(III) with the fluorogenic ligand chromotropic acid (4,5-dihydroxynaphthalene-2,7-disulfonic acid) has been revisited with the aim of using enhancement of the fluorescence intensity as an analytical tool. Complexation at the optimum pH4 was shown to lead to a 1:1 complex with a stability constant log 110=18.4±0.7. The fluorogenic effect was thoroughly investigated. Nearly selective excitation of the chelate rather than the ligand could be achieved at wavelengths longer than 360 nm. For analytical purposes the main interfering ion was Ga3+. The strongest competing ligand was shown to be citric acid. Competitive complexation by acetate or formate ions can also make their use in a buffer at the usual concentration, 0.2 mol L–1, questionable, whereas a 10–2 mol L–1 formic acid buffer was shown to be a good alternative. The calibration plot showed that the dependence of response on Al(III) concentration was linear up to 500 g L–1; the detection limit was 0.65 g L–1 (3SD blank, n=10, SD=±1.4% at 10 g L–1 and ±0.8% at 100 g L–1). The analytical procedure was successfully applied to several samples of tap water and the results were in good agreement with those from AAS determination.  相似文献   

2.
The kinetic results of the oxidation of lactic acid by CrVI in the absence and presence of ethylenediaminetetraacetic acid (EDTA), 2,2-bipyridyl (bpy), MnII and CeIV are presented. EDTA, bpy and MnII catalyse the reaction, whereas CeIV acts as an inhibitor. Irrespective of the conditions, the order with respect to [lactic acid] was found to lie between one and two, and one in [CrVI]. In the EDTA, bpy and MnII-catalysed paths, CrVI–EDTA, CrVI–bpy and lactic acid–MnII complexes have respectively been suggested as the active oxidant and reductant. The kinetic and cerium(IV) effect studies are consistent with a one-step two-electron transfer mechanism in the absence/presence of EDTA and bpy where a chromate–ester mechanism experiences a redox decomposition (C—H cleavage) in the rate-determining step. On the other hand, the catalytic effect of MnII is described as a one-step three-electron redox decomposition (C—C cleavage) mechanism. Mechanisms in accordance with the experimental data have been proposed for the reactions. Activation parameters have also been evaluated and discussed.  相似文献   

3.
A simple, sensitive and specific flow-injection spectrofluorimetric method has been developed for the determination of folic acid in pharmaceuticals. The method is based on use of a lead dioxide solid-phase reactor for on-line oxidation of folic acid into a strongly fluorescent compound with a maximum excitation wavelength of 281 nm and an emission wavelength of 450 nm. Under optimum conditions the fluorescence intensity of oxidation product is proportional to the concentration of folic acid over the range 0.008–2.5 g mL–1. The detection limit is 0.0001 g mL–1, the relative standard deviation is 0.85% for 11 replicate determinations of 0.05 g mL–1 folic acid, and the sample throughput is 20 h–1. In combination with an on-line filter and dilution, an automated drug-dissolution system was established. The proposed method has been successfully applied to the determination of folic acid in pharmaceutical preparations and dissolution testing.  相似文献   

4.
Three types of copolymers of poly(L ‐lactic acid) (PLLA) were synthesized by direct polycondensation of L ‐lactic acid and phenyl‐substituted α‐hydroxy acids (L ‐phenyllactic acid and D ‐ and L ‐mandelic acids). It was found that the glass transition temperature of the copolymers comprising L ‐mandelic acid became significantly higher (from 58 to 69 °C) with increasing content of L ‐mandelic acid (from 0 to 50 mol‐%) although the M w decreased (from 87 000 to 4 000 Da). The cast films of the L ‐mandelic acid containing copolymers showed improved tensile properties compared with those of the PLLA film. This may be due to a pinning effect of the L ‐mandelic acid units on the helix formation of PLLA, although 30% of the units were racemized. The enzymatic degradability of the L ‐mandelic acid containing copolymers was much higher than that of PLLA, as analyzed with Proteinase K® originating from Tritirachium album.

Synthesis of copolymers of L ‐lactic acid and phenyl‐substituted α‐hydroxy acids.  相似文献   


5.
Summary In low acid (0.02 M HClO4) media, Pb2+ ion strongly catalyses the aquation of Cr(ox) 3 3– to givecis-Cr(ox)2(OH2) 2 ion. The catalytic efficiency of Pb2+ as represented by the second order rate constant, kpb (3.76 × 10–4 M–1 s–1 at 25 °C; I, 1.0 M perchlorate), for the Pb2+ catalysed path is remarkably higher than might be expected on the basis of Kpb-ox, the first formation constant for the lead-oxalate complex. This catalytic superiority of Pb2+ has been found to result mainly from a comparatively much lower H (65.2 ±0.8 kJ mol–1) value which more than compensates for the relatively unfavourable S value (–93.2 ±2.4 JK–1mol–1) for this catalysed path. This low S value is, however, in line with the entropy of hydration of Pb2+ ion. These facts, together with the different LFER plots, have been utilised to propose a plausible mechanism for such catalysed reactions.  相似文献   

6.
A simple high performance liquid chromatographic method is described for the determination of seven organic acids usually found in wines. The acids were eluted in the isocratic mode, in less than 12 min, on a reversed-phase ODS-2 (250 × 4 mm I.D.), 5 m, column with 0.02 M potassium dihydrogen phosphate (pH 2.88), to which was added a small amount of methanol (2%) as organic modifier, and were detected by ultraviolet absorbance at 230 nm. Galacturonic, tartaric, malic (both enantiomers), lactic, acetic, citric and succinic acids were determined, using xanthine as a chromatographic internal standard. The method was applied to white and red wines of Greek origin, after sample clean-up with polyvinylpyrrolidone, followed by passage through SAX cartridges and yielded recoveries from 78.0 to 106.8%. The limits of detection ranged between 0.001–0.05 g.L–1 and the linear ranges between 0.003–2.0 g.L–1.  相似文献   

7.
A chiral nitrogen-containing calix[4]crown 2 bearing optically pure 1,2-diphenyl-1,2-oxyamino residue at lower rim showed excellent chiral recognition between enantiomers of mandelic acid. Using competitive 1H NMR titration the ratio of association constants of (S)- and (R)-mandelic acid with the chiral calix[4]crown was determined to be 102, that is 98% de, which is the best result obtained from artificial receptors for the chiral recognition of mandelic acid up to now.  相似文献   

8.
From conductometric and UV-VIS spectrophotometric studies of the reaction between 18-crown-6 (L) and dichloropicric acid (HA) in dry and water saturated 1,2-dichloroethane, it has been concluded that formation of a 1:1 homoconjugate HA 2 accompanies the simple protonation of L, viz, L+HALH+A and L+2HALH+HA 2 . The electrolytes LH+A and LH+HA 2 are extensively, or practically completely dissociated in both solvents under the experimental conditions. The specie LH+A appears to be a contact ion pair in DCE. The stability constant of HA 2 in the dry solvent, 5.7×103 mol–1-cm3, is some 102.4 times that in propylene carbonate reflecting the difference in H-bond accepting capacity of the two solvents. Hydration of HA, A and HA 2 in wet dichloroethane is negligible or slight. As expected, LH+ is rather strongly hydrated, the ratio of the hydration constants of LH+ and L being about 1×101.  相似文献   

9.
The kinetics of oxidation of tartaric acid by Ce(IV) in the absence and presence of acrylamide has been investigated spectrophotometrically in aqueous H2SO4–HClO4 media at a constant ionic strength 2.0M and 25°C. Oxidation of tartaric acid in both cases was first order with respect to Ce(IV). Kinetic data showed that the reaction involves the formation of an unstable complex and an intermediate free radical. The activation parameters were calculated to be E a =91.3±0.4 kJ-mol–1, S=20.2±1.0 J-mol–1-K–1, H=88.8±0.4 kJ-mol–1. A polymerization mechanism is discussed.  相似文献   

10.
The kinetics of oxidation of uranium(IV) monofluoride complex by nitrous acid in nitric acid solution have been studied. The experiments were carried out at constant ionic strength of 2M (HNO3 and NaNO3) and temperature in the range of 18–47 °C. The rate of reaction was determined spectrophotometrically at a wavelength of 621 nm, at which the molar extinction coefficients of UF3+ and UF 2 2+ are the same. It was shown that reaction orders for [HNO2] and [HNO3] are equal to 0.12 and 0.39, respectively. The values of activation parameters H and S are determined to be 83 kJ mol–1 and 75 J (mol·K)–1, respectively. The rate order of the reaction studied has a weak direct dependence on [H+] in contrary to the strong and reverse dependence in the absence of fluoride ions. In conclusion, fluoride ions may strongly stabilize the U(IV) in nitric acid solutions.  相似文献   

11.
Two simple, selective and sensitive spectrophotometric methods are described for the determination of 6-aminopenicillanic acid (6-APA). The methods are based on the reaction of 6-APA with either bromophenol blue (BPB) or bromothymol blue (BTB), to give orange-red and green species, respectively. The coloured products are quantified spectrophotometrically at 625 and 616 nm for BPB and BTB, respectively. The optimization of the different experimental conditions is described. No interferences from different -lactams and common degradation products were observed in the determination of 6-APA using BTB, while flucloxacillin, dicloxacillin, adrenaline, vitamin C, urea and common degradation products in any percentage interfere on using BPB only. The BTB method was better than the BPB method, because of its wider range of determination (0.4–20 g ml–1 vs. 0.4–7.2 g ml–1 on using BPB), higher molar absorptivity and Sandell sensitivity (3.27 × 103l mol–1 cm–1 and 0.099 g cm–2 vs. 2.82 × 103lmol–1 cm–1 and 0.115 g cm–2), greater stability (3 and 10 days on using BTB and BPB, respectively) and better selectivity. The results were compared with those given by the Official United States Pharmacopeial XXI method.  相似文献   

12.
The ionization constant of orthophosphoric acid, determined by conductivity measurements, decreased from 7.11×10–3 at 25°C to 6.2×10–4 mol-kg–1 at 200°C. The pressure effect to 2000 bar was also measured and the ratio K2000/K1 is 2.7 at 25°C and 3.7 at 200°C. The standard partial molar volume change for the ionization at 1 bar, , changes from –16.1 at 25°C to –33.3 cm3-mol–1 at 200°C. The partial molar compressibility change for the ionization, , varies from –3.8×10–3 to –8.3×10–3 cm3-mol–1 bar–1 over the same temperature range.  相似文献   

13.
Polyaniline was deposited potentiodynamically on a stainless steel substrate in the presence of an inorganic acids (sulfuric acid). The electrochemical characterization of the electrode was carried out by means of cyclic voltammetry and electrochemical impedance spectroscopy in the organic acids (p-toluene sulfonic acid) solution. The results show that polyaniline has a high specific capacitance of 431.8 F g−1 at 1 mV s−1, high coulombic efficiency of 95.6% at 20 mV s−1, and exhibits a high reversibility. This indicates the promising feasibility of the polyaniline used as an electrochemical capacitor material in the electrolyte of p-toluene sulfonic acid solution especially at high charge–discharge process.  相似文献   

14.
Summary The kinetics of chromic acid oxidation of phenylphosphinic acid to phenylphosphonic acid has indicated the formation of an anhydride between HCrO 4 and phenylphosphinic acid in its active PhP(OH)2 and inactive PhPH(O)OH forms. The ambiguity about the reactive form of phenylphosphinic acid arises from the fact that protonation of the anhydride leads to the same transition state which disproportionates in the rate-determining step to phosphonium ion and chromium(IV). These, through different reactions in the fast step, yield phenylphosphonic acid and chromium(III) as the final products. That HCrO 4 is the reactive species of chromium(VI) is confirmed by the fact that k0 is independent of the inital [CrVI] where k0 is defined by the Equation k0=kobs[CrVI]/[HCrO 4 ]; kobs is the pseudo first-order rate constant with respect to chromium(VI) ([Phenyl-phosphinic acid][CrVI]).The plot between k0 and [H+] passes through the origin indicating that the reaction does not occur in the absence of H+-ions. Furthermore, the plot between log k0 and –H0, the Hammett acidity function, is linear with a slope value of 1.02±0.02 confirming the protonation of the anhydride prior to its rate-limiting disproportionation.The equilibrium constant for the anhydride formation and the composite rate constant kK, K is the protonation constant of anhydride, are reported. The equilibrium constant is almost independent of temperature.Sen Gupta and Chakaladar(3) reported values of 8.5, 9.2, 11, 12 and 13, respectively, at 26°, 30.4°, 36°, 39.4° and 46° C. The uncertainty limits were not reported. Nevertheless it is apparent from the data that the values are not greatly influenced by temperature. The statistical mean is 11±2 dm3 mol–1, in fair agreement with our values.  相似文献   

15.
Uranium(VI) reacts withN-phenylcinnamohydroxamic acid to form an orange-yellow complex in the pH range 5.5–8.5. The orange-yellow complex, having the composition of 12 (metal:ligand), is quantitatively extractable into ethyl acetate. The spectrum of the complex exhibits a maximum absorption at 400 nm with a molar absorptivity of 6500 M–1·cm–1. The coloured system obeys Beer's law in the concentration range 2–40g·ml–1 of uranium(VI). The photometric sensitivity of the colour reaction is 0.037 g·cm–2 of uranium(VI). Most of the common ions do not interfere and the method has been found to be simple, precise, and free from the rigid control of experimental conditions. The method has been applied to the determination of uranium in synthetic matrices and potable water.  相似文献   

16.
The hydrolysis of the phosphate ester bond in D-pantothenic acid 4-phosphate in strongly acid (9 and 6 N HCl), weakly acid, weakly alkaline, and strongly alkaline (5 N KOH) aqueous media has been studied. In strongly acid and strongly alkaline media the phosphate ester bond of (I) breaks down completely in 3 and 2.5 h, respectively. The rate of hydrolysis at pH values between 1.3 and 11 remains constantly low but almost doubles in the presence of lanthanum salts at pH 8–9 and in the presence of lithium salts at pH 4 and a molar ratio of (I):Me+=1:1.Institute of Biochemistry Academy of Sciences of the Belorussian SSR, Grodno. Translated from Khimiya Prirodnykh Soedinenii, No. 2, pp. 262–265, March–April, 1987.  相似文献   

17.
Summary The oxidation of MnII by S2O8 2– to MnVII in phosphoric acid medium proceeds via a stable MnIII and MnIV species. The reaction is catalysed by Ag+ and exhibits first order dependence on [S2O8 2–], [Ag+] and, is independent of [MnII]. The [H+] has no significant effect on the reaction. It is observed that the PO4 3– ion stabilises the transient manganese(III) and manganese(IV) species by forming a stable and soluble phosphato-complexes. The activation parameters for the two stages of oxidation, namely MnII MnIV and MnIVMnVII at 25° C are Ea=52 ±4 kJ mole–1, S*=–57±2 JK–1 mole–1 and Ea =56±4 kJ mole–1, S*=–44±2 JK–1 mole–1, respectively. A mechanism consistent with the experimental observations is proposed.Presented at the National Symposium on Reaction Kinetics and Mechanism, Department of Chemistry, University of Jodhpur, Jodhpur, India, Nov. 15–18, 1986.  相似文献   

18.
Summary The kinetics of the reaction betweencis-dichlorobisbipyridineruthenium(II) and nitric acid have been investigated spectrophotometrically in the 25°–40° range in the presence of 0.03 to 0.2 mol dm–3 HNO3. The reaction proceeds with the stepwise formation of monoaqua and diaqua products. Only the formation of the monoaqua intermediate was followed as this species could not be obtained in a pure state. Aquation proceeds through a dissociative process. The second order rate constants are 11.8 (25°), 17.5 (30°); 30.0 (35°) l mol–1 s–1. Activation parameters are H 52±3 kJ mol–1; S–108±8 JK–1 mol–1.  相似文献   

19.
Aspartic acid was covalently grafted on to a glassy carbon electrode (GCE) by amine cation radical formation in the electrooxidation of the amino-containing compound. X-ray photoelectron spectroscopic (XPS) measurement and cyclic voltammetric experiments proved the aspartic acid was immobilized as a monolayer on the GCE. Electron transfer to Fe(CN)64– in solution of different pH was studied by cyclic voltammetry. Changes in solution pH resulted in the variation of the charge state of the terminal group; surface pKa values were estimated on the basis of these results. Because of electrostatic interactions between the negatively charged groups on the electrode surface and dopamine (DA) and ascorbic acid (AA), the modified electrode was used for electrochemical differentiation between DA and AA. The peak current for DA at the modified electrode was greatly enhanced and that for AA was significantly reduced, which enabled determination of DA in the presence of AA. The differential pulse voltammetric (DPV) peak current was linearly dependent on DA concentration over the range 1.8×10–6–4.6×10–4 mol L–1 with slope (nA mol–1 L) and intercept (nA) of 47.6 and 49.2, respectively. The detection limit (3) was 1.2×10–6 mol L–1. The high selectivity and sensitivity for dopamine was attributed to charge discrimination and analyte accumulation. The modified electrode has been used for determination of DA in samples, in the presence of AA, with satisfactory results.  相似文献   

20.
A sensitive spectrophotometric method has been developed for the determination of microamounts of thorium using 0.05% thorin in a 3M perchloric acid solution as a chromogenic reagent and measuring the absorbance at 544 nm. The complex of thorium thus formed, is stable for more than two months with a constant absorbance of ±0.55%. Beer's law is obeyed from 0 to 25 g g–1 of thorium in a solution with a molar absorptivity (544 nm) = 1.69×104 M–1 cm–1 at 26±1 °C. Among the anions tested, only phosphate, acetate and cyanide at >200-fold excess of thorium interfere in the determination, whereas cations like Zn(II), Al(III), Na(I), Mg(II), and Ca(II) do not effect the absorbance. Thorium can be determined in the presence of oxalate, nitrate, tartrate, sulfate, thiosulfate, citrate, and ascorbate. The accuracy of the method has been checked by measuring the known concentration of thorium in the range of 100 g-5 mg g–1 and found to be in the range of 7.7–0.9%. The method has been applied successfully to determine thorium at g g–1 level in local ore samples with a precision of ±0.3%. The sensitivity of the method on Sandell's scale is 0.082±0.002 g g–1 cm–1.  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