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 15 毫秒
1.
Preferential adsorption of allopurinol (1H-pyrazolo-[3,4-d]-pyrimidin-4-ol) at graphite electrodes from pH 7.0 phosphate buffer solutions of 2,6-diaminopurine and 2,6-diamino-8-purinol allowed in situ electrode modification. Modified electrodes were applied to the simultaneous determination of 2,6-diaminopurine and 2,6-diamino-8-purinol in phosphate buffer at pH 7.0. Differential pulse polarography allowed both compounds to be quantified in the concentration range 1 × 10?6–5 × 10?4 M. The relative standard deviation of the peak current is about 10%. Calibration curve characteristics are C (μM) = 0.0094 ± 0.0002 I (μA) + 0.22 ± 0.07 μA, with r = 0.9985 and C (μM) = 0.0065 ± 0.0001 I (μA) + 0.03 ± 0.04 μA with r = 0.9990 for 2,6-diamino-8-purinol and 2,6-diaminopurine, respectively.  相似文献   

2.
Photoirradiation of dipyridamole (DIP) solution at λ ⪢ 390 nm leads to a reduction of the intensity of the absorption band without change in its appearance. This reduction is due to the breaking of the π-conjugation chain of the DIP molecule. The experimentally measured rate constant k is proportional to the concentration of oxygen and to the ratio of the rates of photoreaction, kr, and of radiation, ko. In homogeneous solutions the values of kr and ko are three-times greater at pH 5.0 than at pH 7.0. So, the photoreaction is more effective when the DIP molecules are completely protonated. In alcoholic and alkaline solutions (pH 13.5), the deprotonation of DIP molecules is responsible for a significant reduction of kr. In the presence of microheterogeneous systems, lysophosphatidylcholine (L-PC) micelles and bovine serum albumin (BSA), a significant reduction of kr is observed. The value of kr in this case depends on the pH of the solution as well as on the concentration of L-PC and BSA. In the presence of L-PC or BSA, different values of k could be associated to free and bound DIP molecules. A kinetic procedure is proposed which permits the evaluation of the binding constants as well as the kinetic constants. The binding constant of DIP to L-PC micelles is estimated as 1.28 × 104 M−1 and the value of the limiting effective rate constant k is similar to the value obtained in ethanol. This is evidence for the localization of DIP molecules in the polar region of L-PC micelles. The value of k in this case is three-times greater at pH 5.0 than at pH 7.0 as occurs in homogeneous solutions. The binding to BSA was also studied and binding constants of (1.8±0.2) × 104 M−1 and (7.8±0.9) × 104 M−1 were obtained at pH 7.0 and pH 5.0, respectively. In this case the ratio k(ph 5.0)/k(ph 7.0) is also equal to three, the same as in homogeneous solutions or in the presence of L-PC micelles. This implies that DIP molecules maintain their protonation form in the presence of microheterogeneous systems.  相似文献   

3.
In this study, an oxadiazole multi-wall carbon nanotube-modified glassy carbon electrode (OMWCNT?GCE) was used as a highly sensitive electrochemical sensor for hydrazine determination. The surface charge transfer rate constant, k s, and the charge transfer coefficient, ??, for electron transfer between GCE and electrodeposited oxadiazole were calculated as 19.4?±?0.5?s?1 and 0.51, respectively at pH?=?7.0. The obtained results indicate that hydrazine peak potential at OMWCNT?GCE shifted for 14, 109, and 136?mV to negative values as compared with oxadiazole-modified GCE, MWCNT?GCE, and activated GCE surface, respectively. The electron transfer coefficient, ??, and the heterogeneous rate constant, k??, for the oxidation of hydrazine at OMWCNT?GCE were also determined by cyclic voltammetry measurements. Two linear dynamic ranges of 0.6 to 10.0???M and 10.0 to 400.0???M and detection limit of 0.17???M for hydrazine determination were evaluated using differential pulse voltammetry. In addition, OMWCNT?GCE was shown to be successfully applied to determine hydrazine in various water samples.  相似文献   

4.
《Electroanalysis》2017,29(11):2572-2578
We report in this work, a new method for the determination of captopril by differential pulse voltammetry using a glassy carbon electrode modified with a copper metal‐organic framework (H‐Kust‐1 or Cu3(BTC)2 or Cu‐BTC), immobilized on the surface by a copolymer of acrylamide and sodium acrylate. This compound is detected by the formation of a copper(II)‐captopril complex that is observed in an oxidation potential at ca. +0.28 V vs . Ag/AgCl. A linear dynamic range is obtained for a captopril concentration of 0.5 μM to 7.0 μM and the voltammetric response is highly reproducible within 3.52 % error. The sensitivity of 9.71±0.37 nA μM−1 and the limit of detection of 0.20±0.01 μM make this methodology highly applicable for practical applications. The determination of captopril in a commercial pharmaceutical sample showed a recovery of 93.3 %.  相似文献   

5.
The dark reduction kinetics of micromolar concentrations of Fe(III) in aqueous solution were studied in the presence of millimolar concentrations of ferrozine (FZ) over the pH range 4.0–7.0. A pseudo-first-order kinetics model was used to describe Fe(III) reduction at pH 4.0 and 5.0, and the reduction rate decreased with increasing pH or initial Fe(III) concentration. A more molecular-based kinetics model was developed to describe Fe(III) reduction at pH 6.0 and 7.0. From this model, the intrinsic rate constants (k1) of Fe(III) reduction by FZ in the dark were obtained as 0.133 ± 0.004 M?1 s?1 at pH 6.0 and 0.101 ± 0.009 M?1 s?1 at pH 7.0. It was also found in this model that a higher pH, a higher concentration of Fe(III), a lower concentration of FZ and less incubation time led to a lower fraction of Fe(III) reduction by FZ in the dark.  相似文献   

6.
The decay of Br-2 in Ar-purged or N2O-saturated aqueous solutions of KBr (0.01-1.0 M) in the pH range 1–7 has been re-examined using the techniques of pulse radiolysis and computer simulation. The dependence of the rate constant for the intrinsic decay of Br-2 on ionic strength (controlled by KBr) has been established; the values of k (Br-2 + Br-2) are (1.9 ± 0.1) × 109, (2.2 ± 0.3) × 109 and (2.4 ± 0.3) × 109 M-1 s-1 in the presence of 0.01, 0.1 and 1.0 M KBr, respectively, independent of pH between 2 and 7. The computer simulation of the decay of Br-2 has also generated, for the latter species, ϵ = 10,000 ± 700 M-1 cm-1 at λmax = 360 nm; this value has been calculated without making any assumption concerning G(Br-2). For the reduction of Br-2 by H atoms, a value of k (H + Br-2) = (1.4 ± 0.3) × 1010 M-1 s-1 has been obtained in the presence of 0.01-1.0 M KBr, independent of pH between 1–4. For the reduction of Br-2 by e-aq at pH 7 (10-3 M phosphates) and μ = 0.1, a value of k (Br-2 + e-aq = (1.1 ± 0.2) × 1010 M-1 s-1 has been obtained.  相似文献   

7.
《Analytical letters》2012,45(7):1213-1233
Abstract

Anodic waves of methimazole (I) (1-methylimidazole-2-thiol) and carbimazole (II) (1-ethoxycarbonyl-3-methyl-2-thio-4-imidazoline) on mercury electrodes correspond to mercury salt formation. Both compounds form in the thiono form a soluble complex at pH < 6, compound (I) at higher pH-values a slightly soluble salt of the thiol form. Electrode processes involving the thiol form are complicated by adsorption. Oxidation at solid electrodes occurs only at potentials more than 0.5 V more positive. For compound (I) spectrophotometry indicated pKa=12.0 ± 0.2. By d.c. polarography in 0.1 M H2SO4 containing 10% ethanol the determination of both compounds is possible between 4 × 10? and 1 × 10?3 M, by differential pulse polarography between 1 × 10? and 1 × 10?4 M, by differential pulse voltammetry at HMDE between 5 × l0?7 and 6 × 10? M.  相似文献   

8.
A new approach to obtaining thermoset organotin polymers, which permits control of crosslinking site distribution and, through it, a better control of properties of organotin antifouling polymers, is reported. Tri-n-butyltin acrylate and tri-n-butyltin methacrylate monomers were prepared and copolymerized, by the solution polymerization method with the use of free-radical initiators, with several vinyl monomers containing either an epoxy or a hydroxyl functional group. The reactivity ratios were determined for six pairs of monomers by using the analytical YBR method to solve the differential form of the copolymer equation. For copolymerization of tri-n-butyltin acrylate (M1) with glycidyl acrylate (M2), these reactivity ratios were n = 0.295 ± 0.053, r2 = 1.409 ± 0.103; with glycidyl methacrylate (M2) they were r1 = 0.344 ± 0.201, r2 = 4.290 ± 0.273; and with N-methylolacrylamide (M2) they were r1 = 0.977 ± 0.087, r2 = 1.258 ± 0.038. Similarly, for the copolymerization of tri-n-butyltin methacrylate (Mi) with glycidyl aery late (M2) these reactivity ratios were r1 = 1.356 ± 0.157, r2 = 0.367 ± 0.086; with glycidyl methacrylate (M2) they were r1 = 0.754 ± 0.128, r2 = 0.794 ± 0.135; and with N-methylolacrylamide (M2) they were r1 ?4.230 ± 0.658, r2 = 0.381 ± 0.074. Even though the magnitude of error in determination of reactivity ratios was small, it was not found possible to assign consistent Q,e values to either of the organotin monomers for all of its copolymerizations. Therefore, Q,e values were obtained by averaging all Q,e values found for the particular monomer, and these were Q = 0.852, e = 0.197 for the tri-n-butyltin methacrylate monomer; and Q = 0.235, e = 0.401 for the tri-n-butyltin acrylate monomer. Since the reactivity ratios indicate the distribution of the units of a particular monomer in the polymer chain, the measured values are discussed in relation to the selection of a suitable copolymer which, when cross-linked with appropriate crosslinking agents through functional groups, would give thermoset organotin coatings with an optimal balance of mechanical and antifouling properties.  相似文献   

9.
Acrolein was copolymerized by radical initiation in aqueous solutions with sodium p-styrenesulfonate and acrylic acid, respectively, in the pH range of 3–7. The reactivities were shown to be pH-dependent. For the acrolein (M1)–sodium p-styrenesulfonate (M2) pair, r1 = 0.33 ± 0.15 and r2 = 0.32 ± 0.05 at pH 3; r1 = 0.23 ± 0.12 and r2 = 0.05 ± 0.03 at pH 5; r1 = 0.26 ± 0.03 and r2 = 0.025 ± 0.025 at pH 7. For the acrolein (M1)–acrylic acid (M2) pair, r1 = 0.50 ± 0.30 and r2 = 1.15 ± 0.2 at pH 3; r1 = 2.40 ± 0.50 and r2 = 0.05 ± 0.05 at pH 5; r1 = 6.70 ± 3.00 and r2 = 0.00 at pH 7. For acrolein, the new values of Q = 1.6 and e = 1.2 have been calculated. For sodium p-styrenesulfonate, the values Q = 0.76 and e = ?0.26 at pH 3, Q = 0.51 and e = ?0.87 at pH 5, Q = 0.39 and e = ?1.00 at pH 7 were obtained; and for acrylic acid, the values Q = 1.27 and e = 0.50 at pH 3, Q = 0.11 and e = ?0.22 at pH 5 were derived. The changes in reactivity are explained on the basis of inductive and resonance effects.  相似文献   

10.
The differential pulse polarographic behavior of p-nitrosodimethylaniline has been investigated in Britton-Robinson buffer and phosphate buffer. The peak obtained at pH 8.0 is recommended for the trace determination of this compound, with an experimental detection limit of 18 ppb (1.2 × 10?7M) in simple aqueous solution. The method is also applied for the indirect microdetermination of NADH. The experimental detection limit of NADH is shown to be 1.05 × 10?6M.  相似文献   

11.
New copolymers of the vinyl saccharide 2-deoxy-2-methacrylamido-D-glucose (M1) with acrylic and methacrylic (M2) acids differing in composition and molecular mass have been synthesized by free-radical copolymerization. The relative activities of the comonomers are determined. It is found that, for acrylic acid, r 1 = 3.03 ± 0.15 and r 2 = 0.5 ± 0.08 and, for methacrylic acid, r 1 = 1.070 ± 0.1 and r 2 = 1.18 ± 0.13. As is evidenced by potentiometric and viscometric measurements, the vinyl saccharide and acid units are capable of interacting, a circumstance that affects the conformational states of macromolecules.  相似文献   

12.
The electrochemical properties of violuric acid (VA) have been investigated at pH 4.0–10.0 by using cyclic voltammetry on a glassy carbon electrode. The peak current was proportional to the square root of the potential scan rate. The calculated diffusion coefficient was 2.0±0.7×10−6 cm2 s−1. The formal oxidation–reduction potential of VA was 0.63 V versus SCE at pH 7.0. The kinetics of VA interaction with reduced glucose oxidase (GO) was explored in the electrocatalytical system. A typical electrocatalytical wave was generated in the presence of the VA and glucose. An apparent kox calculated by using the Nicholson–Shain function was 1.85×106 M−1 s−1 at pH 7.0 and 25 °C. Glucose and l-lactate bioelectrodes were prepared by adsorbing the GO and l-lactate oxidase (LO) onto the VA-modified graphite electrode. The electrode was poised at 0.6 V versus SCE and linear response was obtained over the range of 4–20 mM glucose and 2–12 mM l-lactate, respectively.  相似文献   

13.
A sensitive and specific spectrofluorimetric assay was developed for the determination of 2-hydroxynicotinic acid (1) and its major metabolite, the N-1-riboside (11) in blood and urine. Both compounds exhibit strong fluorescence in acidic media. Thin-layer chromatography is employed to separate the drug from its metabolite; the compounds are eluted from the silica gel into methanol—0.1 M hydrochloric acid (80+20). The sensitivity is 0.04–0.05 μg of I and 0.10–0.12 μg of II per ml of blood or urine. The two compounds can also be determined by differential pulse polarography, which is especially suitable for urine. The fluorimetric method was applied to the determination of blood levels and the urinary excretion of the drug in man after single oral 500-mg doses.  相似文献   

14.

In this study, an oxadiazole multi-wall carbon nanotube-modified glassy carbon electrode (OMWCNT−GCE) was used as a highly sensitive electrochemical sensor for hydrazine determination. The surface charge transfer rate constant, k s, and the charge transfer coefficient, α, for electron transfer between GCE and electrodeposited oxadiazole were calculated as 19.4 ± 0.5 s−1 and 0.51, respectively at pH = 7.0. The obtained results indicate that hydrazine peak potential at OMWCNT−GCE shifted for 14, 109, and 136 mV to negative values as compared with oxadiazole-modified GCE, MWCNT−GCE, and activated GCE surface, respectively. The electron transfer coefficient, α, and the heterogeneous rate constant, k′, for the oxidation of hydrazine at OMWCNT−GCE were also determined by cyclic voltammetry measurements. Two linear dynamic ranges of 0.6 to 10.0 μM and 10.0 to 400.0 μM and detection limit of 0.17 μM for hydrazine determination were evaluated using differential pulse voltammetry. In addition, OMWCNT−GCE was shown to be successfully applied to determine hydrazine in various water samples.

  相似文献   

15.
A differential pulse polarographic method for the determination of the herbicide thiazopyr has been developed. The polarographic study of thiazopyr exhibited two well-defined cathodic peaks within the pH range of 1.0 to 8.0. The variation of pH and polarographic parameters indicated that the optimum conditions under which thiazopyr could be reduced were a pH 7.0 BR buffer solution, a reduction peak potential of ?1270 mV (vs. SCE), scan rate of 5 mV s?1, pulse amplitude of 50 mV with pulse duration of 50 ms at an ambient temperature of 25 ± 3°C. The main reduction peak was characterised by cyclic voltammetry as being irreversible and diffusion-controlled. A linear relationship between the peak current and the concentration of thiazopyr was obtained in the range of 0.43–38.6 µg mL?1, with a detection limit of 0.127 µg mL?1. The proposed method was successfully applied to the determination of thiazopyr in spiked fruit juice and soil samples. The mean recoveries of the 19.8 µg g?1 and 3.96 µg mL?1 thiazopyr spiked to soil and orange juice were 20.2 ± 1.0 µg g?1 and 3.84 ± 0.12 µg mL?1, at 95% confidence level, respectively. The sufficiently good recoveries and low relative standard deviation (RSD) data confirm the high accuracy and precision of the proposed method. The interferences effects of several commonly used pesticides and inorganic species were also studied. Interfering effects were eliminated either by providing selectivity with pH, or using EDTA as complexing agent.  相似文献   

16.
The selective interactions of azide radicals with glutathione (GSH) have been quantitatively studied in buffered neutral aqueous solutions using γ and pulse radiolysis. The sulfur centered GS̽ and GSSG⨪ radicals are produced in pulse and γ radiolysis. Kinetic experiments and simulation allowed to estimate the rate constant of N̽3 with GSH which has been found to be equal to (9.5 ± 0.5) × 106M−1 s−1 at pH 7. In steady state radiolysis, we have found GSSG as the final product formed with an initial G value of 2.9 × 10−7 mol J−1.  相似文献   

17.
The decomposition of meta-phenylphenol (m-PP) and para-phenylphenol (p-PP) in a heterogeneous gas-liquid system using ozone was investigated. The influence of different reaction parameters such as ozone and PP isomers concentration as well as pH and temperature of the reaction mixture on the PP decay rate was determined. The second-order rate constants for the direct reaction of molecular ozone, determined in a homogeneous system, were (5.85 ± 0.35) × 102 M?1 s?1 and (8.90 ± 0.33) × 102 M?1 s?1 for m-PP and p-PP, respectively. The rate constants for the reaction of m-PP and p-PP with ozone increased with increasing pH. The reaction rate constants with ozone were found to be (1.75 ± 0.02) × 109 M?1 s?1 and (1.86 ± 0.02) × 109 M?1 s?1 for m-PP and p-PP anions, respectively.  相似文献   

18.
《Analytical letters》2012,45(15):2860-2871
Abstract

A voltammetric study of the oxidation of disopyramide has been carried out using a glassy carbon electrode. The electrochemical oxidation of disopyramide was investigated by cyclic, differential pulse, and square wave voltammetry. The oxidation of disopyramide is an irreversible, diffusion‐controlled process. The diffusion coefficient of disopyramide was calculated in pH 7.0 phosphate buffer to be D disopyramide=3.8×10?6 cm2 s?1. The oxidation of disopyramide is also pH dependent and for electrolytes with pH between 4 and 7 occurs with the transfer of one electron and one proton. In alkaline electrolytes, two consecutive charge transfer reactions are observed: both oxidation reactions involve the transfer of two electrons but only the first also involves the transfer of two protons. Two procedures for the analytical determination of disopyramide in pH 7.0 phosphate buffer were developed and compared and a detection limit LOD=1.27 µM was obtained.  相似文献   

19.
Laser flash photolysis combined with competition kinetics with SCN? as the reference substance has been used to determine the rate constants of OH radicals with three fluorinated and three chlorinated ethanols in water as a function of temperature. The following Arrhenius expressions have been obtained for the reactions of OH radicals with (1) 2‐fluoroethanol, k1(T) = (5.7 ± 0.8) × 1011 exp((?2047 ± 1202)/T) M?1 s?1, (2) 2,2‐difluoroethanol, k2(T) = (4.5 ± 0.5) × 109 exp((?855 ± 796)/T) M?1 s?1, (3) 2,2,2‐trifluoroethanol, k3(T) = (2.0 ± 0.1) × 1011 exp((?2400 ± 790)/T) M?1 s?1, (4) 2‐chloroethanol, k4(T) = (3.0 ± 0.2) × 1010 exp((?1067 ± 440)/T) M?1 s?1, (5) 2, 2‐dichloroethanol, k5(T) = (2.1 ± 0.2) × 1010 exp((?1179 ± 517)/T) M?1 s?1, and (6) 2,2,2‐trichloroethanol, k6(T) = (1.6 ± 0.1) × 1010 exp((?1237 ± 550)/T) M?1 s?1. All experiments were carried out at temperatures between 288 and 328 K and at pH = 5.5–6.5. This set of compounds has been chosen for a detailed study because of their possible environmental impact as alternatives to chlorofluorocarbon and hydrogen‐containing chlorofluorocarbon compounds in the case of the fluorinated alcohols and due to the demonstrated toxicity when chlorinated alcohols are considered. The observed rate constants and derived activation energies of the reactions are correlated with the corresponding bond dissociation energy (BDE) and ionization potential (IP), where the BDEs and IPs of the chlorinated ethanols have been calculated using quantum mechanical calculations. The errors stated in this study are statistical errors for a confidence interval of 95%. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 174–188, 2008  相似文献   

20.
The purpose of this study was to assess the stability of the polyfructan levan under different pH solution conditions by monitoring changes in the levan physicochemical properties, such as molar mass (M), root mean square radius (r rms ), hydrodynamic radius (r h ), structure factor (r rms /r h ), and aggregation state with respect to solution pH and hydrolysis time. A commercial levan produced from Z. Mobilis was characterized using asymmetric flow field-flow fractionation (AF4) in combination with online multiangle light scattering (MALS) and differential refractive index (dRI) detection. Under neutral pH solution conditions the levan was found to have a M ranging from 105 to 5?×?107 g/mol, a r rms ranging from ~25 to 100 nm and a r h from ~3 to 151 nm. Two populations were observed in the sample. One population with a M less than 106 g/mol which represented ~60 % of the sample and a second population with an ultrahigh M up to 5?×?107 g/mol, which comprised ~40 % of the sample. The measured r rms /r h structure factor decreased from 1.8 to 0.65 across the AF4 fractogram indicating that early eluting low M levan species had a random coil configuration and late eluting high M species had more homogeneous spherical structures. The measured apparent density values decreased from 80 to 10 kg/m3 across the elution profile and suggest that the observed second population also contains aggregates. The stability of levan in different pH conditions ranging from 1.3 to 8.5 was assessed by tracking changes in the average M and r h , and monitoring the formation of fructose over 1 week. The onset of levan acid hydrolysis was observed to occur sooner at lower pH conditions and no hydrolysis was observed for pH 5.5 and higher.  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