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 62 毫秒
1.
Karayannis MI 《Talanta》1976,23(1):27-30
The rate constant k of the reaction of ascorbic acid with 2,6-dichlorophenolindophenol (DCPI) in oxalic acid solutions is determined, by stopped-flow techniques. Four different methods are used to evaluate the results. The values and errors are compared statistically. The average of the rate constant is 56.5 x 10(3) l. mole(-1), sec(-1) and the overall standard deviation is 0.6 x 10(3) l. mole(-1), sec(-1) or 1.0% relative. The pH-dependence of the rate constant suggests that DCPI reacts with undissociated ascorbic acid.  相似文献   

2.
Nakagawa G  Wada H  Nakazawa O 《Talanta》1976,23(2):155-157
The rate of the ligand-substitution reaction of nickel(II)-TAC chelate (NiR(2)) with EDTA (Y) and 1,10-phenanthroline (X) has been determined spectrophotometrically in 20% v v dioxan over the pH range 5.7-6.3 at mu = 0.1 (KNO(3)) and 25 +/- 1 degrees . The substitution reaction with EDTA proceeds through the following two pathways: NiR(2) + H(+) right harpoon over left harpoon NiR(+) + HR, and NiR(2) + H(2)O right harpoon over left harpoon NiR(OH) + HR, The reaction of NiR(+) or NiR(OH) with EDTA is the rate-determining step, and k(1) = 2.1 x 10(3) l .mole(-1) .sec(-1) and k(2) = 7.9 x 10(6) l .mole(-1) .sec(-1).The substitution reaction with 1,10-phenanthroline proceeds as follows: NiR(+) + X right harpoon over left harpoon NiRX(+) At higher concentrations of 1,10-phenanthroline the release of TAC from NiR(2) by hydrogen ion is the rate-determining step, and k(3) = 2.4 x 10(5) l .mole(-1). sec(-1). At lower concentrations of 1,10-phenanthroline -d[NiR(2)]/dt is proportional both to [H(+)] and [X]. The value k(4) = 5.1 x 10(4) l. mole(-1). sec(-1) was calculated by the use of the steady-state approximation for [NiRX(+)]. The substitution with 1,10-phenanthroline proceeds much faster than that with EDTA. By the addition of a small amount of 1,10-phenanthroline, Ni can be titrated with EDTA at 50 degrees, with TAC as an indicator.  相似文献   

3.
Abstract— ESR studies have been made of the kinetics of semiquinone radical formation and disappearance resulting from the reversible photosensitization by chlorophyll of hydroquinone oxidation in a pyridine-water solvent. The rate of radical decay was found to be second order with respect to the radical concentration, with a rate constant of 6.7 × 105 l./mole sec at -30°C and an activation energy of 6900 cal/mole. The rate of radical formation was recombination-limited and, through the use of β-carotene as a quencher, the rate constant was determined to be 8.81 × 105 l./mole sec at -30°C. The effect of light intensity and hydroquinone concentration on the rate of semiquinone radical formation and on the steady state radical concentration was also investigated and possible mechanisms to explain the results are discussed.  相似文献   

4.
Qu YB 《Talanta》1991,38(9):1061-1066
In 6M hydrochloric acid solution, salicylic acid reacts with pentachloronitrosyliridate [Ir(NO)Cl(5)](-) to form a new 1:1 complex giving two absorption maxima at 369 and 446 nm with molar absorptivities both of 1.1 x 10(4) l.mole(-1) cm(-). The reaction is second order with apparent energy, E(a), of 16.8 kJ/mole. At constant temperature, the apparent rate constant k increases to a maximum when the concentration of hydrogen ion is decreased to 4-5M and the concentration of chloride ion is kept at approximately 6M. The mechanism of the reaction is also discussed. In alkaline solution, the absorption maxima shift to 381 and 506 nm with molar absorptivities of 6.0 x 10(3) and 2.4 x 10(4) l.mole(-1) cm(-1), respectively. Salicylic acid can be determined spectrophotometrically with high sensitivity and precision. Benzoic and boric acids do not interfere.  相似文献   

5.
Cedergren A 《Talanta》1974,21(4):265-271
Reaction rates between water and the Karl Fischer reagent have been determined by potentiometric measurement for various compositions of the Karl Fischer reagent. The study has been made with an iodine complex concentration of 0.3-1.2 mM and sulphur dioxide complex at 0.01-0.5M. The concentration of excess of pyridine had no measurable effect on the rate of the main reaction. The reaction was found to be first-order with respect to iodine complex, to sulphur dioxide complex, and to water. The rate constant was (1.2+/-0.2) x 10(3) 1(2). mole(-2). sec(-1). In an ordinary titration it is therefore essential to keep the sulphur dioxide concentration high for the reaction to go to completion within a reasonable time. The extent of side-reactions was found to be independent of the iodine concentration at low concentrations. The side-reactions increased somewhat with increasing sulphur dioxide pyridine concentrations and decreased to about 60% when the temperature was lowered from 24 degrees to 7 degrees.  相似文献   

6.
Konidari CN  Karayannis MI 《Talanta》1991,38(9):1019-1026
The reduction of 2,6-dichlorophenolindophenol (DCPI) by sulphides and sulphites has been studied kinetically by the stopped-flow technique. The reaction is first-order with respect to each of the reactants. From the distribution diagrams for the species DH(+)(2), DH and D(-) for DCPI and H(2)Q, HQ(-) and Q(2-) for sulphides or sulphites, a mechanism is proposed which suggests partial reactions of all possible combinations of the reacting species at any pH. An equation for calculation of the second-order reaction rate constants k at any pH is derived, which gives k as a function of [H(+)], the partial reaction rate constants and the dissociation constants of DCPI and H(2)S or H(2)SO(3). Values of the overall reaction rate constants over a wide pH-range have been determined, together with values of k for all possible partial reactions. For particular pH-values the second-order reaction rate constant was determined by four different methods. Mean values of k = 251 +/- 1 and 240 +/- 1 l.mole(-1).sec(-1) were obtained for pH 3.15 and 4.17, respectively, for the DCPI-Na(2)S reaction and k = 137 +/- 1, 127 +/- 1 and 136 +/- 1 l.mole(-1).sec(-1) for pH 2.02, 4.25 and 5.10, respectively, for the DCPI-Na(2)SO(3) reaction. From the slopes of the linear Arrhenius plots activation energies of 6.6 +/- 0.2 and 4.0 +/- 0.1 kcal/mole for the DCPI-Na(2)S and DCPI-Na(2)SO(3) reactions, respectively were calculated. The effect of ionic strength on the reactions supports the proposed mechanism.  相似文献   

7.
Echols RT  Tyson JF 《Talanta》1994,41(10):1775-1784
Equations have been derived for the concentration-time profiles of reactants and products in a first order reaction obtained on passage of a reactant plug through a single well-stirred tank reactor. When taken together with the equations for physical dispersion of such a reactor under plug flow conditions, an expression for the reaction rate constant was derived which allowed its experimental determination in a relatively simple fashion. The method was tested for reactions between cerium and oxalic acid and between dichromate and ascorbic acid, for which values of the rate constants of around 2 x 10(2) sec(-1) and 5.5 x 10(3) sec(-1) were obtained. Good agreement with other experimentally determined values was obtained. The scope and the limitations of the proposed method are critically discussed with the aid of some model calculations. The range of values for which the method might be suitable is approximately 10(-3)-10(-1) sec(-1). An equation analogous to a peak-width equation was derived as a further development of this approach. Good agreement with the previously determined values were obtained for both systems. The extension of the method to reactions other than first order is discussed.  相似文献   

8.
Safavi A  Ensafi AA  Massoumi A 《Talanta》1991,38(2):229-231
The reaction between nickel and ammonium 2-amino-1-cyclohexene-1-dithiocarboate in aqueous acetone medium at pH 3.0-8.0 results in a stable dark red complex. The ratio of reagent to nickel in the complex is 2:1 and the formation constant is 7.38 +/- 0.12 x 10(10). Beer's law is obeyed up to 4 microg/ml nickel at the absorption maximum at 535 nm. The apparent molar absorptivity is 2.8 x 10(4) l.mole(-1). cm(-1), and the detection limit is 10 ng/ml nickel. The method is applied to the determination of nickel in vegetable oil.  相似文献   

9.
Abstract— ESR studies have been made of the kinetics of semiquinone radical formation and disappearance resulting from the reversible photosensitization of reduction or oxidation, by chlorophyll, pheophytin or hematoporphyrin, of several quinone-hydroquinone pairs in various solvents. The rate of radical decay was found to be second order with respect to the radical concentration in all systems. Radical formation rates were determined by the initial production rate minus the decay rate. The kinetic constants for single electron transfer between triplet porphyrins and quinones or hydroquinones were determined usingβ-carotene as a quencher in aqueous pyridine, and by measuring the initial rate of radical formation at various concentrations of quinones and hydroquinones in methanol and ethanol. These constants were found to be approximately the same in a given solvent for benzoquinone and hydroquinone with all porphyrins, though the rates differed in different solvents: pyridine-water ~ 106I./mole sec, and methanol and ethanol ~ 5X 104l./mole sec. Trimethylquinone and its hydroquinone also give similar rate constants for radical formation in pyridine-water, ~ 106 l./mole sec. The second order radical decay constants for both benzoquinone and hydroquinone in pyridine-water were the same, ~ 105 I./mole sec, with either chlorophyll, pheophytin or hematoporphyrin as sensitizer. The same activation energy, 6900 cal/mole, was found for chlorophyll-benzoquinone and hydroquinone in aqueous pyridine; 5500 cal/mole was obtained for these systems in ethanol. In methanol and ethanol solutions of chlorophyll, the same radical decay rate constants, ~106 I./mole sec, were observed for both benzoquinone and hydroquinone. Also, the same decay constants, ~ 106 I./mole sec, were found for trimethylquinone and its hydroquinine in pyridine-water. These latter two compounds gave extremely small steady-state ESR signals in ethanol compared with aqueous pyridine. We have also observed that the steady-state signal obtained with chlorophyll-menadione in ethanol-water was much enhanced by the presence of NADH. In contrast, NAD+ was found to decrease radical production, by increasing the decay rate, in the chlorophyll-hydroquinone system in aqueous pyridine. These results are discussed in terms of possible mechanisms for radical formation and disappearance. The most likely possibility is considered to be a one-electron oxidation or reduction of the porphyrin triplet, followed by radical disproportionation and redox reactions between the disproportionation products.  相似文献   

10.
Hung SC  Qu CL  Wu SS 《Talanta》1982,29(7):629-631
A highly sensitive method for spectrophotometric determination of uranium has been devised. The method is based on formation of a red-violet 1:2 (metal:ligand) complex from the reaction of uranium(VI) with 2-(3,5-dibromo-2-pyridylazo)-5-diethylaminophenol (3,5-diBr-PADAP) in the presence of an anionic surfactant, sodium lauryl sulphate. Its molar absorptivity is found to be 9.1 x 10(4)l.mole(-1).cm(-1). The absorbance is constant in the range pH 8.4-9.9 Beer's law is obeyed for 0-1.4 mug/ml concentrations of uranium. In the presence of DCTA the method is selective for uranium, and can be used for the determination of trace amounts of uranium in water samples.  相似文献   

11.
Trimm HH  Ushio H  Patel RC 《Talanta》1981,28(10):753-757
A combined stopped-flow temperature-jump apparatus interfaced with a dedicated microcomputer has been used to study the complexation reaction of iron(III) with thiocyanate in aqueous solution. Kinetic rate-constants (k(f) = 143 l.mole(-1) .sec(-1) from T-jump, k(f) = 150 l.mole(-1) .sec(-1) from stopped flow), equilibrium constants (K = 143 from T-jump, K = 150 from stopped flow) and the thermodynamic enthalpy change (DeltaH(c) = -6.7 kJ/mole) could be independently determined from the simultaneous application of the two techniques.  相似文献   

12.
Uesugi K  Shigematsu T 《Talanta》1977,24(6):391-393
A new spectrophotometric method for the determination of gallium with Pontachrome Azure Blue B and cetyltrimethylammonium chloride is described. The sensitivity of the colour reaction between gallium and Pontachrome Azure Blue B is greatly increased in the presence of cetyltrimethylammonium chloride. The gallium complex has maximal absorbance at 680 nm and pH 6.0-6.6. Beer's law is obeyed over the range 0.08-0.6 ppm of gallium; the molar absorptivity is 1.39 x 10(5) 1.mole(-1). cm(-1) and sensitivity 4.9 x 10(-4) mu/cm(2). The mole ratio of the complex, the formation constant and effect of interfering ions are described.  相似文献   

13.
Singh I  Mrs Poonam  Kadyan PS 《Talanta》1985,32(5):387-390
1-(2-Quinolylazo)-2,4,5-trihydroxybenzene has been proposed as a sensitive chromogenic reagent for the simultaneous determination of iron and copper in presence of thiosemicarbazide. The molar absorptivities are 1.86 x 10(4) l.mole(-1).cm(-1) at 510 nm for iron and 2.54 x 10(4) l.mole(-1).cm(-1) at 550 nm for copper. Trace levels of these two metals have been determined in some foodstuff's, body tissues and milk samples.  相似文献   

14.
Cadmium (II) ion-catalyzed degradation of ampicillin in methanol at 20 degrees C has been studied. It has been observed that the rate values tend to saturate when the concentration of ampicillin or the metal ion is increased. The results obtained in the present study suggest that ampicillin degradation occurs through the formation of a 1:1 (SM) and 2:1 (S(2)M) ampicillin-metal complexes. These complexes decompose giving a single product (absorption maximum at 285 nm; ((p)=1.82x10(4) l mol(-1) cm(-1)) that has been isolated and identified (Cd(II) (L(2-))(2) (H(2)O)(4) Na(2)). The appearance of this product reflects a first order reaction with respect to the 1:1 complex, with a rate constant of 3.87x10(-2) min(-1) and the existence of an equilibrium between the 1:1 and 2:1 initial complexes. The equilibrium constant value, calculated from kinetic data, is 1.7x10(3) l mol(-1).  相似文献   

15.
Reported here is a detailed study of the kinetics and mechanism of formation of a 1,4 GG interstrand cross-link by [(trans-PtCl(NH(3))(2))(2)(mu-NH(2)(CH(2))(n)NH(2))](2+) (1,1/t,t (n = 6), 1), the prototype of a novel class of platinum antitumor complexes. The reaction of the self-complementary 12-mer duplex 5'-[d(ATATGTACATAT)(2)] with (15)N-1 has been studied at 298 K, pH 5.4, by [(1)H,(15)N] HSQC 2D NMR spectroscopy. Initial electrostatic interactions with the duplex are observed for 1 and the monoaqua monochloro species (2). Aquation of 1 to yield 2 occurs with a pseudo-first-order rate constant of (4.15 +/- 0.04) x 10(-5) s(-1). 2 then undergoes monofunctional binding to the guanine N7 of the duplex to form 3 (G/Cl) with a rate constant of 0.47 +/- 0.06 M(-(1) s(-1). There is an electrostatic interaction between the unbound [PtN(3)Cl] group of 3 and the duplex, which is consistent with H-bonding interactions observed in the molecular model of the monofunctional (G/Cl) adduct. Closure of 3 to form the 1,4 GG interstrand cross-link (5) most likely proceeds via the aquated (G/H(2)O) intermediate (4) (pseudo-first-order rate constant = (3.62 +/- 0.04) x 10(-5) s(-1)) followed by closure of 4 to form 5 (rate constant = (2.7 +/- 1.5) x 10(-3) s(-1)). When closure is treated as direct from 3 (G/Cl) the rate constant is (3.39 +/- 0.04) x 10(-5) s(-1). Closure is ca. 10-55-fold faster than that found for 1,2 GG intrastrand cross-link formation by the diaqua form of cisplatin. Changes in the (1)H and (15)N shifts of the interstrand cross-link 5 indicate that the initially formed conformer (5(i)) converts irreversibly into other product conformer(s) 5(f). The NMR data for 5(i) are consistent with a molecular model of the 1,4 GG interstrand cross-link on B-form DNA, which shows that the NH(2) protons have no contacts except with solvent. The NMR data for 5(f) show several distinct NH(2) environments indicative of interactions between the NH(2) protons and the DNA. HPLC characterization of the final product showed only one major product peak that was confirmed by ESI-FTICR mass spectroscopy to be a cross-linked adduct of (15)N-1 and the duplex. The potential significance of these findings to the antitumor activity of dinuclear platinum complexes is discussed.  相似文献   

16.
Ying-Quan Z  Lin Z  Jun-Yi L 《Talanta》1983,30(4):291-293
Aluminium reacts with Chlorophosphonazo I at pH 4-6 to form a water-soluble 1:1 red-violet complex. The absorption maximum is at 610 nm and the apparent molar absorptivity is 1.96 x 10(4) l.mole(-1).cm(-1). The apparent instability constant has been calculated to be 1.8 x 10(-4). The colour of the complex is stable for 24 hr. Beer's law is obeyed over the concentration range up to 0.8 mug ml . Titanium and zirconium interfere seriously.  相似文献   

17.
The reaction between cadmium and 2-(5-chloro-2-pyridylazo)-5-dimethylaminophenol (5-Cl DMPAP) in aqueous alcohol media at pH 8.8-10.7 results in an intense violet colour which is stable for at least 8 hr. The composition is 2:1 reagent:metal and the formation constant (5.29 +/- 0.01) x 10(18). Beer's law is obeyed up to 1.34 ppm of cadmium at 550 nm. The optimal concentration range (Ringbom) is between 0.16 and 0.72 ppm. The apparent molar absorptivity at 550 nm is (1.20 +/- 0.01) x 10(5) l.mole(-1). cm(-1), making the sensitivity one of the highest known. The interference due to copper(III), iron(III), cobalt(II), nickel(II), gold(III), zinc(II) and manganese(II) can be suppressed.  相似文献   

18.
The uptake of NH3 and the heterogeneous reaction of NH3 + HOBr --> products on ice surfaces at 190 K have been investigated in a flow reactor coupled with a differentially pumped quadrupole mass spectrometer. The uptake coefficient gammat for NH3 was determined to be (3.8 +/- 1.4) x 10(-4) on ice films at 189.8 K, for a partial pressure of NH3 in the range of 7.0 x 10(-7) to 3.8 x 10(-6) torr. The amount of NH3 uptake on the ice film was determined to be >2.9 x 10(15) molecules/cm(2), based on the total ice surface area at 189.2 K. The heterogeneous reaction of NH3 + HOBr on ice surfaces has been studied at 190 K. The reaction probability gammat was determined to be (5.3 +/- 2.2) x 10(-4) and was found to vary insignificantly as HOBr surface coverage changes from 2.1 x 10(13) to 2.1 x 10(14) molecules/cm(2). A reaction pathway is proposed on the basis of experimental observations.  相似文献   

19.
Qian-Feng W  Peng-Fei L 《Talanta》1983,30(4):275-276
A highly sensitive spectrophotometric method for determination of arsenic, based on the formation of an ion-association complex between arsenoantimonomolybdenum blue and Malachite Green, has been developed. The ion-association complex is soluble in the presence of Triton X-305, so arsenic can be determined directly in aqueous solution. The apparent molar absorptivity for arsenic is 1.13 x 10(5) l.mole(-1).cm(-1) at 640 nm. Beer's law is obeyed for 0-5 mug of arsenic. The lower limit of determination (absorbance = 0.01) is 4 ng ml in the final solution.  相似文献   

20.
The reactivity of -SC(NH)NH2 (1), MeCOS- (2), and PhCOS- (3) toward 1-naphthyl radicals was studied in DMSO. The photostimulated reaction of anions 1, 2, and 3 with 1-bromonaphthalene (4) after quenching with MeI renders 1-(methylthio)naphthalene (6) as a main product together with bis(1-naphthyl) sulfide (7) and naphthalene (5). The thioacetate ion (2) and thiobenzoate ion (3) were unreactive toward 4 as electron-donor under photostimulation; however, in the presence of potassium tert-butoxide anion (entrainment conditions), they gave the mentioned products 5, 6, and 7, after the addition of MeI. Quenching of the triplet state of 4 was assigned as the photoinduced initiation step, with a rate constant value of (4.6+/-0.5)x10(8) M-1 s-1 for tert-butoxide anion and a rough estimated value of (8+/-7)x10(7) M-1 s-1 for anion 1. By using hydrogen abstraction from DMSO as the competitive reaction, the absolute rate constants for the addition of anions 1, 2, and 3 to 1-naphthyl radicals have been determined to be 1.0x10(9), 1.2x10(9), and 3.5x10(9) M-1 s-1, respectively. This reactivity order is in agreement with the stability of the resulting radical anions (ArNu)*- (10-12)*-. The inhibition experiments of the photoinduced substitution reaction in the presence of radical scavengers and the global quantum yield higher than the unity are evidence of a radical chain mechanism for these substitution reactions by anions 1 and 2. Anion 3 adds to the 1-naphthyl radical, but is neither able to initiate nor to keep the propagation cycle. Evaluation of the electron-transfer driving forces for the reaction between (ArNu)*- and 4 together with the absence of a chain reaction for the anion 3 indicate that the propagation in the proposed mechanism is given by an acid-base reaction between the radical .C(O)Me or .C(NH)NH2 (13) and a base.  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