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 203 毫秒
1.
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   

2.
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   

3.
Di-cysteine substituted hypocrellin B (DCHB) is a new water-soluble photosensitizer with significantly enhanced red absorption at wavelengths longer than 600 nm over the parent compound hypocrellin B (HB). The photosensitizing properties (Type I and/or Type II mechanisms) of DCHB have been investigated in dimethylsulfoxide (DMSO) and aqueous solution (pH 7.4) using electron paramagnetic resonance (EPR) and spectrophotometric methods. In anaerobic DMSO solution, the semiquinone anion radical of DCHB (DCHB•−) is predominantly photoproduced via self-electron transfer between excited- and ground-state DCHB species. The presence of an electron donor significantly promotes the formation of the reduced form of DCHB. When a deoxygenated aqueous solution of DCHB and an electron donor are irradiated with 532 nm light, the hydroquinone of DCHB (DCHBH2) is formed via the disproportionation of the first-formed DCHB•− and second electron transfer to DCHB•− from the electron donor. When oxygen is present, singlet oxygen (1O2), superoxide anion radical (O2•−) and hydroxyl radical (OH) are produced. The quantum yield of 1O2 generation by DCHB photosensitization is estimated to be 0.54 using Rose Bengal as a reference, a little lower than that of HB (0.76). The superoxide anion radical is also significantly enhanced by the presence of electron donors. Moreover, (O2•−) upon disproportionation generated H2O2 and ultimately the highly reactive OH via the Haber-Weiss reaction pathway. The efficiency of (O2•−) generation by DCHB is obviously enhanced over that of HB. These findings suggest that the photodynamic actions of DCHB may proceed via Type I and Type II mechanisms and that this new photosensitizer retains photosensitizing activity after photodynamic therapy-oriented chemical modification.  相似文献   

4.
The triethylsilane radical R3Si, produced by radiolysis in an airfree methanol/silane-system, acts as a specific scavenger for the CH3O and CH2OH transients with rate constants, k14(R3Si + CH3O) = 1.1 x 108 dm3 mol-1s-1 and k15(R3Si + CH2OH) = 0.7 x 108 dm3 mol-1s-1, resulting in R3Si—OCH3 (triethylmethoxysilane) and R3Si—CH2OH (triethylsilylmethanol). By increasing the silane concentration (range: 10-2-6 mol dm-3 R3SiH) the formation of the otherwise major products of methanol radiolysis, formaldehyde and glycol, is successively reduced to nil. The yield of R3Si—CH2OH, studied under the same conditions, passes a maximum at about 0.8 mol dm-3 R3SiH and then also diminishes. On the other hand, the yield of R3Si—OCH3 is increased correspondingly and reaches an interpolated value of G = 3.75 ± 0.1 at 4 mol dm-3R3SiH. This indicates that the radical CH3O (G = 3.75 ± 0.1) is the primary radiolytic transient of methanol in addition to H, e-sol etc., but not CH2OH species. The latter one is obviously formed by the secondary reaction: CH3O + CH3OH→ CH3OH + CH2OH.  相似文献   

5.
Ab initio calculations were performed for special points of the minimal energy pathways (MEP) of the nucleophilic addition reactions of the isolated H anion, LiH molecule and Li+/H ion pair to acetylene (A) and methylacetylene (MA) molecules, proceeding in accordance (M) and against (aM) the Markovnikov's rule. All structural parameters were optimized using the restricted Hartree–Fock (RHF) method. For the addition of H, the 6-31++G* basis set was used and for the reactions of LiH and Li+/H the 6-31G* basis set with the subsequent recalculation of single point energies, taking into account of electron correlation energy by means of the second-order Möller–Plesset perturbation theory at the MP2/6-31++G** level. The results of calculations demonstrate, that the energy characteristics of both M- and aM-additions with H do not differ sufficiently (0.1–1.2 kcal/mol for the activation energies (ΔEa) and the reaction heats (ΔQ)). The substitution of the H atom by the CH3 group in A molecule results in practically the same values of ΔQ and ΔEa. On the contrary, for the LiH molecule and Li+/H ionic pair, the M-addition is favorable (charge control). It is found that the presence of electrophile decreases the activation energy by 3–5 kcal/mol as compared with the addition of the isolated hydride ion H.  相似文献   

6.
The reactions of two triphenyl methane (TPM) dyes—crystal violet (CV+) and malachite green (MG+)—with N3 and OH radicals were studied by pulse radiolytic kinetic spectrophotometry. The rate constants for the reaction of the cationic dyes (D+) with N3 are (9.0±0.6)×109 and (3.0±0.2)×109 dm3 mol−1 s−1 respectively and those for the reaction with OH are obtained as (8.0±0.6)×109 and (1.1±0.1)×109 dm3 mol−1 s−1 respectively. The transient spectra resulting from the oxidation of the dyes were characterized. The time-resolved spectra indicate that the reaction with OH radicals initially generates an adduct which subsequently dissociates to form the radical dication D•2+. The D•2+ species decay by further reaction with the parent dye.  相似文献   

7.
The reactions of hydroxyl radical, hydrogen atom and hydrated electron intermediates of water radiolysis with N-isopropylacrylamide (NIPAAm) were studied by pulse radiolysis in dilute aqueous solutions. OH, H and eaq react with NIPAAm with rate coefficient of (6.9±1.2)×109, (6.6±1)×109, and (1.0±0.2)×1010 mol−1 dm3 s−1. In OH and H radical addition to the double bond mainly -carboxyalkyl type radicals form, (OHCH2CHC(N-i-C3H7)O and CH3CHC(N-i-C3H7)O). In reaction of eaq oxygen atom centered radical anion is produced (CH2CHC(N-i-C3H7)O), the anion undergoes reversible protonation with pKa=8.7. There is also an irreversible protonation on the β-carbon atom that produces the same radical as forms in H atom reaction (CH3CHC(N-i-C3H7)O). The -carboxyalkyl type radicals at low NIPAAm concentration (0.1–1 mmol dm−3) mainly disappear in self-termination reactions, 2kt,m=8.4×108 mol−1 dm3 s−1. At higher concentrations the decay curves reflect the competition of the self-termination and radical addition to monomer (propagation). The termination rate coefficient of oligomer radicals containing a few monomer units is 2kt≈2×108 mol−1 dm3 s1.  相似文献   

8.
The rate constants at which oxidizing and reducing radicals react with the dinuclear iron(III) complex Fe2O(ttha)2− were measured in neutral aqueous solution. The rate constants for reduction of the complex by ·CO2.− CH3.CHOH and O2.− were found to be comparable with rate constants previously measured in mononuclear iron(III) polyaminocarboxylate systems. Fe2O(ttha)2− reacts slowly with O2.− (k8 = (1.2 ± 0.2) × 104 dm3 mol−1 s−1) and, hence, is a relatively poor catalyst for the dismutation of superoxide radical. The hydrated electron reduces the complex at a diffusion-controlled rate in a process which consumes one proton: eaq + Fe2O(ttha)2− → Fe2III,IIO(ttha)3− The reduction by carbon-centered radicals produces a (III,II) mixed-valence complex with an absorption spectrum different from that of the Fe2(II,III) species produced from reduction by the hydrated electron. The oxidizing radicals .OH and ·CO3 appear to act as reductants of the complex via ligand oxidation rather than by oxidation of the Fe2IIIO core to Fe2III,IVO. In the former case ligand attack appears to occur mainly at the methylene carbon of a glycinate group. The decarboxylation product, CO2, was detected by its aquation reaction in the presence of a pH sensitive dye, bromthymol blue.  相似文献   

9.
A complete electrochemical study and a novel electroanalytical procedure for bromhexine quantitation are described. Bromhexine in methanol/0.1 mol L−1 Britton–Robinson buffer solution (2.5/97.5) shows an anodic response on glassy carbon electrode between pH 2 and 7.5. By DPV and CV, both peak potential and current peak values were pH-dependent in all the pH range studied. A break at pH 5.5 in EP versus pH plot revealing a protonation–deprotonation (pKa) equilibrium of bromhexine was observed. Spectrophotometrically, an apparent pKa value of 4.3 was also determined.

An electrodic mechanism involving the oxidation of bromhexine via two-electrons and two-protons was proposed. Controlled potential electrolysis followed by HPLC–UV and GC–MS permitted the identification of three oxidation products: N-methylcyclohexanamine, 2-amino-3,5-dibromobenzaldehyde and 2,4,8,10-tetrabromo dibenzo[b,f][1,5] diazocine.

DPV at pH 2 was selected as optimal pH for analytical purposes. Repeatability, reproducibility and selectivity parameters were adequate to quantify bromhexine in pharmaceutical forms. The recovery was 94.50 ± 2.03% and the detection and quantitation limits were 1.4 × 10−5 and 1.6 × 10−5 mol L−1, respectively. Furthermore, the DPV method was applied successfully to individual tablet assay in order to verify the uniformity content of bromhexine. No special treatment of sample were required due to excipients do not interfered with the analytical signal. Finally the method was not time-consuming and less expensive than the HPLC one.  相似文献   


10.
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   

11.
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   

12.
Ionizing radiation inactivated purified chicken intestinal aminopeptidase in media saturated with gases in the order N2O>N2>air. The D37 values in the above conditions were 281, 210 and 198 Gy, respectively. OH radical scavengers such as t-butanol and isopropanol effectively nullified the radiation-induced damage in N2O. The radicals (SCN)2•−, Br2•− and I2•− inactivated the enzyme, pointing to the involvement of aromatic amino acids and cysteine in its catalytic activity. The enzyme exhibited fluorescence emission at 340 nm which is characteristic of tryptophan. The radiation-induced loss of activity was accompanied by a decrease in the fluorescence of the enzyme suggesting a predominant influence on tryptophan residues. The enzyme inhibition was associated with a marked increase in the Km and a decrease in the Vmax and kcat values, suggesting an irreversible alteration in the catalytic site. The above observations were confirmed by pulse radiolysis studies.  相似文献   

13.
UV spectra and kinetics for the reactions of alkyl and alkylperoxy radicals from methyl tert-butyl ether (MTBE) were studied in 1 atm of SF6 by the pulse radiolysis-UV absorption technique. UV spectra for the radical mixtures were quantified from 215 to 340 nm. At 240 nm. σR = (2.6 ± 0.4) × 10−18 cm2 molecule−1 and σRO2 = (4.1 ± 0.6) × 10−18 cm2 molecule−1 (base e). The rate constant for the self-reaction of the alkyl radicals is (2.5 ± 1.1) × 10−11 cm3 molecule−1 s−1. The rate constants for reaction of the alkyl radicals with molecular oxygen and the alkylperoxy radicals with NO and NO2 are (9.1 ± 1.5) × 10−13, (4.3 ± 1.6) × 10−12 and (1.2 ± 0.3) × 10−11 cm3 molecule−1 s−1, respectively. The rate constants given above refer to reaction at the tert-butyl side of the molecule.  相似文献   

14.
Low energy (<3 eV) electrons impact to gas phase Adenine generates the dehydrogenated Adenine negative fragments, (A–H), and an H-atom neutral radical counterpart. Within the energy range of 0.7–2.8 eV, production of (A–H) arises from Dissociative Electron Attachment (DEA). In addition, a sharp peak is observed at near 0 eV. This peak is identified to arise from dissociative electron transfer reaction of SF6 (from the calibration gas) with Adenine.  相似文献   

15.
The effects of (+)-catechin adsorption to the alumina surface were studied by ζ-potential and surface free energy determination. The presence of catechin causes essential changes in the alumina ζ-potential, which at the concentration slightly higher than 10−5 M reverses from the positive into negative one. At constant concentration of catechin (10−3 M), the effect on ζ-potential of alumina as a function of pH appears in a drastic shift of the isoelectric point, from pH 8.4 to 4.6, and the equilibrium is established practically within 2 h. This is probably due to relatively low pKa=4.6 for catechin 3′-OH group deprotonation. At high alkaline environment (pH≥10), even in the presence of catechin in the solution, the hydroxyl OH ions play principal role in the surface charge formation for the alumina. At such pH catechin molecule is double negatively charged and hence its adsorption on highly negatively charged alumina surface is rather restricted. Nevertheless, various dimeric forms of catechin, which are formed at the alkaline pH, probably adsorb on the alumina surface. This appears in small increase in apolar surface free energy component at natural and alkaline pH. On the other hand, at acidic pH 3 small increase of the electron acceptor interaction is observed. This may result from increased number of hydroxyl groups on the alumina surface originating from the adsorbed molecules of catechin, which are mostly undissociated at this pH. The interactions of catechin with alumina surface seems to be also of some specific nature, because neither changes in the ultraviolet–visible (UV–vis) absorbance (Part I) nor in the ζ-potentials had occurred in the silica suspensions in which also catechin was present.  相似文献   

16.
The transient optical absorption bands formed at λmax=340 and 435 nm, on reaction of OH radicals in aerated acidic aqueous solutions of 1,1,1-trifluoro-2-iodoethane at low and high solute concentration, have been assigned to monomer and dimer radical cations, respectively. The deprotonation of the solute radical cations is the rate-determining step for the decay of the dimer radical cations. The stability constant for the dimer radical cation is determined to be 50 dm3 mol−1 at 25°C. The dimer radical cation is a strong one-electron oxidant. Quantum chemical calculations and experimental results confirm that fluorine reduces the electron density at iodine and the OH-radical-induced oxidation of fluoroiodoalkanes becomes a difficult process compared to iodoalkanes.  相似文献   

17.
The role of photocatalysts in improving the photochemical synthesis of fatty amines has been examined. Ab initio SCF and CI calculations for the addition of the aminyl radical NH2 to ethylene are reported. In contrast with previous theoretical results a low activation energy and a rather small exothermicity were determined. Thus an explanation for the known inactivity of amino radicals in addition reactions can be found in the possibility of a retroaddition. The calculations indicate that the role of photocatalysts could be explained by a significant stabilization of the addition photoproduct blocking the reversibility of the process. The involvement of the photocatalyst with the primary photoproduct NH2 should also be considered in view of a significant increase in the exothermicity observed.  相似文献   

18.
Analytical results of anion determination by suppressed ion chromatography are significantly affected by calibration curve calculation. In this paper, as expected, eluent pKa is shown to influence calibration linearity in the range 1–20 mg/l sulfate, with A carbonate-hydrogencarbonate mixture producing a larger non-linearity than NaOH. Evidence is given for very large errors (about 30–40%) in estimating sample sulfate concentration when linear regression is used instead of a quadratic calibration curve. This study was performed following a 24 run full factorial experimental design, including eluent pKa, counterion type, solution composition and current level for background suppression as main variables.  相似文献   

19.
2,3-Dicarboxypyrazine (2,3-H2DCPz), unlike 2,3-dicarboxypyridine (2,3-H2DCPy), does not exist as a zwitterion in the solid state. This is due to the remarkably low base strength of nitrogen in the pyrazine ring relative to that of pyridine. The decrease in the base strength of pyrazine relative to pyridine may result from greater disruption of the aromatic π-system of pyrazine on protonation than is the case with pyridine. We propose that 2,3-H2DCPz has a structure of Cs symmetry in the solid state in which one carboxyl group forms an internal hydrogen bond to carbonyl of adjacent carboxyl in the same molecule, and the second carboxyl forms an intermolecular hydrogen bond to carbonyl in an adjacent molecule. The monoanion in 2,3-NaHDCPz has a strong O–H–O covalent three-center hydrogen bond between carboxylate groups. When saturated solutions of 2,3-H2DCPz are treated with F, one additional diacid molecule is taken into solution for each 1.5 F added. As is the case with 2,3-H2DCPy, the high solubility and acid strength of 2,3-H2DCPz (pKa=2.87) leads to formation of the hydrogen-rich H2F3 ion instead of HF2.  相似文献   

20.
Infrared spectroscopy and matrix isolation technique have been used to study the 1 : 1 complexes formed between 2,4,5-trichlorophenol (TCP), pentachlorophenol (PCP) or 2-chloro-4,6-dinitrophenol (CNP) and trimethylamine (TMA) isolated in solid argon. The results were analyzed in relation to the type of complex formed. Depending on the proton-donor ability of the phenol three different types of hydrogen bonded complexes have been identified in argon matrices. The weakest phenol in the series, TCP (pKa = 6.72), forms a strong molecular hydrogen bonded complex with TMA as indicated by the broad ν(OHN) absorption with a maximum at 2490 cm−1 and a band at 811 cm−1 due to the νs(C3N) mode of the perturbed amine. The strongest phenol, CNP (pKa = 2.01), interacts with TMA in an argon matrix to form ionic complex with the proton transferred to the base molecule. This is evidenced by the presence of the ν(NH+---O) absorption between 3000−1800 cm−1, by the νas(C3N+) and νs(C3N+) absorptions due to the protonated amine and by numerous product bands due to the relatively strongly perturbed modes of the phenol ring. The interaction between TMA and a phenol of intermediate strength, PCP (pKa = 4.74), in solid argon probably leads to the formation of two types of hydrogen bonded complexes: an ionic complex with the proton transferred to the amine molecule and a pseudosymmetric one with the proton more or less equally shared between the phenol and amine molecules. In this case the protonic absorption consists of two broad features situated in the 3000–1600 cm−1 and 950–400 cm−1 regions due to the ν(NH+O) and ν(OHN) modes, respectively.  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