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.
The transient absorption bands (λmax = 330, 525 nm, kf = 5 × 109 dm3 mol−1 s−1) obtained on pulse radiolysis of N2O‐saturated neutral aqueous solution of 4,4′‐thiodiphenol (TDPH) are due to the reaction of TDPH with ·OH radicals and are assigned to phenoxyl radical formed on fast deprotonation of the solute radical cation. The reaction of specific one‐electron oxidants (Cl2·−, Br2·−, N3·, TI2+, CCl3OO·) with TDPH also produced similar transient absorption bands. The phenoxyl radicals are also produced on pulse radiolysis of N2‐saturated solution of TDPH in 1,2‐dichloroethane. The nature of transient absorption spectrum obtained on reaction of ·OH radicals with TDPH is not affected in acidic solutions, showing that OH‐adduct is not formed in neutral solutions. The oxidation potential for the formation of phenoxyl radical is determined to be 0.98 V. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 603–610, 1999  相似文献   

2.
The free radical scavenging activity of hydroxytyrosol (HTyr) and tyrosol (Tyr) has been studied in aqueous and lipid solutions, using the density functional theory. Four mechanisms of reaction have been considered: single electron transfer (SET), sequential electron proton transfer (SEPT), hydrogen transfer (HT), and radical adduct formation. It was found that while SET and SEPT do not contribute to the overall reactivity of HTyr and Tyr toward ·OOH and ·OCH3 radicals, they can be important for their reactions with ·OH, ·OCCl3, and ·OOCCl3. The ·OOH-scavenging activity of HTyr and Tyr was found to take place exclusively by HT, and it is also predicted to be the main mechanism for their reactions with ·OCH3. HT is proposed as the main mechanism for the scavenging activity of HTyr and Tyr when reacting with other ·OR and ·OOR radicals, provided that R is an alkyl or an alkenyl group. The major products of reaction are predicted to be the phenoxyl radicals. In addition, Tyr was found to be less efficient than HTyr as free radical scavenger. Moreover, while HTyr is predicted to be a good peroxyl scavenger, Tyr is predicted to be only moderately for that purpose.  相似文献   

3.
The mechanism of the radiation-thermal conversion of lignin including the formation of aromatic radical cations and their fragmentation resulting in the appearance of phenoxyl radicals is considered. The multipath formation of phenoxyl radicals occurs with the participation of the reactions of molecules with electrons and small radicals (?Н and ?СН3) and electronic excitation relaxation processes. Phenoxyl radicals are characterized by smaller thermal stability in comparison with that of parent macromolecules. The further thermally stimulated decomposition of these radicals results in the release of monohydric and dihydric phenols from a polymeric chain. The most effective liberation of phenols takes place on the surface of lignin particles, whereas the formation of wood charcoal with the participation of unsaturated products dominates in the bulk. The formation of dihydric phenols is intensified in the presence of alkanes in the irradiated sample; this fact is indicative of an important role of ?Н and ?СН3 radicals in the formation of monomeric phenol products.  相似文献   

4.
《Analytical letters》2012,45(9):2025-2038
Abstract

A simple and highly sensitive method to quantify the rates of production of phenoxyl radicals in enzyme reaction is described. This method employs the peroxidase‐catalyzed reaction between chlorophenols and hydroperoxide to generate phenoxyl free radicals, which can enhance dimerization of L‐tyrosine. The product, dityrosine, was monitored fluorometrically at the excitation/emission wavelength of 320/410 nm and the initial rate of accelerated‐accumulation of dityrosine represents the formation rate of phenoxyl free radicals. With this method, the phenoxyl radicals generated in oxidation of chlorophenols with hydrogen peroxide, catalyzed by horseradish peroxidase, were investigated. Phenoxyl radicals generated from as low as 5.0×10?9 M 2,4‐dichlorophenol, for example, can be readily detected with a relative standard deviation of 2.6% for 9 replicated determination. The detection limits of phenoxyl radicals produced by various chlorophenols are 4.2×10?9, 1.1×10?9, 1.0×10?10, 2.8×10?8, and 1.1×10?7 M for 2‐chlorophenol, 4‐chlorophenol, 2,4‐dichlorophenol, 2,4,6‐trichlorophenol, and 2,3,4,6‐tetrachlorophenol, respectively. The possible pathway of the reaction is proposed. The protocol is suitable for quantification of free radicals in enzyme reaction and shows promise in being applied to biological systems.  相似文献   

5.
At near neutral pH (approx. 5.5), the OH-adduct of chlorogenic acid (CGA), formed on pulse radiolysis of N2O-saturated aqueous CGA solutions (λ max = 400 and 450 nm) with k = 9 × 109 dm3 mol−1 s−1, rapidly eliminates water (k = 1 × 103 s−1) to give a resonance-stabilized phenoxyl type of radical. Oxygen rapidly adds to the OH-adduct of CGA (pH 5.5) to form a peroxyl type of radical (k = 6 × 107 dm3 mol−1 s−1). At pH 10.5, where both the hydroxyl groups of CGA are deprotonated, the rate of reaction of · OH radicals with CGA was essentially the same as at pH 5.5, although there was a marked shift in the absorption maximum to approx. 500 nm. The CGA phenoxyl radical formed with more specific one-electron oxidants, viz., Br 2 ·− and N 3 · radicals show an absorption maximum at 385 and 500 nm, k ranging from 1–5.5 × 109 dm3 mol−1 s−1. Reactions of other one-electron oxidants, viz., NO 2 · , NO· and CCl3OO· radicals, are also discussed. Repair rates of thymidine, cytidine and guanosine radicals generated pulse radiolytically at pH 9.5 by CGA are in the range of (0.7–3) × 109 dm3 mol−1 s−1.  相似文献   

6.
The oxidative splitting process of cis-syn 1,3-dimethyluracil cyclobutane dimer(DMUD) in aqueous solution was investigated using pulse radiolysis technique.The results indicated that DMUD can be splitted into 1,3-dimethyluracil(DMU) by OH radicals(OH) and Br2 radical anions(Br2^-),but not by azide radicals(N3^).The oxidative mechanisms that an H-abstracted from DMUD for OH oxidative splitting and an electron transfer from DMUD to Br2-,were suggested.Related kinetic parameters were determined.  相似文献   

7.
《European Polymer Journal》1987,23(5):383-388
Trialkyl phosphites react with cyanoisopropylperoxyl radicals, generated by thermolysis of azobis(isobutyronitrile) in the presence of oxygen, to give the corresponding phosphates with rate constants of the order of 103 M−1 sec−1 at 65°C. Phenyl phosphites are oxidized also. A small amount of cyanoisopropyl phosphite is formed by substitution of the phosphite by alkyloxyl radicals leading to phenoxyl radicals. Sterically hindered aryl phosphites react with cyanoisopropylperoxyl radicals to yield the corresponding phosphates and alkoxyl radicals which in a second step react with phosphite by substitution releasing a sterically hindered phenoxyl radical. Therefore, sterically hindered phosphites are capable of acting as chain-terminating primary antioxidants. Because the rate constants of reaction of these phosphites with peroxyl radicals are only in the range of 102 M−1 sec−1 and 100 times smaller than those of phenols, phosphites should be less active as primary antioxidants than phenols.  相似文献   

8.
In a model system of initiated oxidation of methyl oleate, the antioxidant activities of 3-hydroxy-2-ethyl-6-methylpyridinesuccinate (mexidol) and bis[3-(3,5-di-tert-butyl-4-hydroxyphenyl)propyl] disulfide (SO-4) were studied and compared with those of α-tocopherol and 1-hydroxy-2,6-di-tert-butyl-4-methylbenzene (dibunol). A linear pattern of dependence of the inhibitory effect on the concentration of compounds was established. The ability of antioxidants to decompose hydroperoxides and inhibit their accumulation was revealed. The combined inhibitory effects of SO-4 with mexidol, α-tocopherol, and phospholipids were described for the first time. The rate constant for disproportionation of the SO-4 phenoxyl radicals, k 9 = 0.90·103 L mol−1 s−1, was determined by steady-state photolysis. The rate constant k 10 eff for the reactions of SO-4 phenoxyl radicals with lipids characterized by different unsaturation degrees were determined for methyl oleate, linolic and arachidonic acids containing one, two, and four multiple bonds, and phospholiopids containing polyunsaturated fatty acids. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1330–1337, August, 2006.  相似文献   

9.
The bimolecular rate constants for the reactions of sulfate radicals with epicatechin (EC), epicatechingallate (ECG), and epigallocatechingallate (EGCG) were found to be (1.46 ± 0.06) × 109, (1.20 ± 0.08) × 109, and (1.04 ± 0.07) × 109, respectively. The activation energy [EA = 9 ± 3 kJ mol?1] and preexponential factor [A = (4.8 ± 0.6) × 1010] for the reaction of EC with the sulfate radical were measured in the temperature range 288–303 K. The phenoxyl radicals of EC (λmax = 310 nm) were obtained both by the reaction of this flavonoid with the sulfate radicals and by photoionization. The measured bimolecular rate constants for the reactions of the dihydrogen phosphate radicals with EC, ECG, and EGCG were (7.8 ± 0.9) × 108, (8.5 ± 0.4) × 108, and (6.8 ± 0.4) × 108, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 391–396, 2010  相似文献   

10.
Addition of the ·P(O)(OPri)2, Me·, Et·, ·But, and Cl3C· radicals to the (ν2-C60)Os(CO)-(PPh3)2(CNBut) complex (1) was studied by ESR spectroscopy. The spectral parameters of the spin-adducts of these radicals with complex 1 were determined. The predominant direction of the attack by the ·P(O)(OPri)2, ·But, and Cl3C· radicals are the cis-1 and cis-2 bonds of the fullerene molecule. The stability of the spin-adducts depends substantially on the nature of the added radical. The addition rate constants of the ·P(O)(OPri)2, ·But, and Cl3C· radicals to complex 1 and the dimerization rate constants for these spin-adducts were determined. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 301–307, February, 2008.  相似文献   

11.
Photolysis of pyrophosphate and tripolyphosphate ions in aqueous solution is proposed to produce electron detachment with formation of pyrophosphate and tripolyphosphate radicals (with quantum yields <0.1 at 266 nm), respectively. Formation of P2O7·3− is observed after photolysis of both polyphosphate ions, because the decomposition of P3O10·4− yields P2O7·3−. The latter radicals further react with hydroxyl ions (k = 1.4 × 106 M−1s−1) generating HO· radicals. The reaction of the solvated electrons with molecular oxygen produces O2·−. The rate constants for the reaction of SO4·− radicals with P2O74− and P3O105− were also measured. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 111–117, 2000  相似文献   

12.
Reactions of eaq -, H-atom and OH radicals with 3-pyridine methanol (3-PM) and 3-pyridine carboxaldehyde (3-PCA) have been studied at various pHs using pulse radiolysis technique. eaq - was found to be highly reactive with both 3-PM and 3-PCA (k approx. 1010 dm3 mol 1 s-1). Semi-reduced species formed in both cases were strongly reducing in nature. In the case of 3-PM, electron addition leads to the formation of pyridinyl radicals whereas in the case of 3-PCA, PyCHOH type radicals are formed. At pH 6.8, H-atom reaction with 3-PCA also gives semi-reduced species (PyCHOH), whereas at pH 1, H-atoms add to the ring. (CH3)2 ·COH radicals were found to transfer electron to 3-PCA at all the pH values tested and by making use of changes in the absorption spectra, pK a values of the semi-reduced species were determined to be 4.5 and 10.6. OH radicals were found to undergo addition reaction with 3-PCA, whereas in the case of 3-PM they reacted by H-abstraction as well as addition reaction. By following the yield of methylviologen radical cation formed by electron transfer reaction, it was estimated that approx. 50% of OH radicals react with 3-PM by H-atom abstraction at pH 6.8, giving reducing radicals, whereas at pH 3.2, where 3-PM is in the protonated form, the same is only about 10%. At pH 13, O radical anions were found to react exclusively by H-atom abstraction. Reaction of SO4 radicals with 3-PCA was found to give a species identical to the one formed by one electron reduction of nicotinic acid at acidic pH values.  相似文献   

13.
The intermolecular interactions existing at three different sites between phenylacetylene and LiX (X = OH, NH2, F, Cl, Br, CN, NC) have been investigated by means of second‐order Møller?Plesset perturbation theory (MP2) calculations and quantum theory of “atoms in molecules” (QTAIM) studies. At each site, the lithium‐bonding interactions with electron‐withdrawing groups (? F, ? Cl, ? Br, ? CN, ? NC) were found to be stronger than those with electron‐donating groups (? OH and ? NH2). Molecular graphs of C6H5C?CH···LiF and πC6H5C?CH···LiF show the same connectional positions, and the electron densities at the lithium bond critical points (BCPs) of the πC6H5C?CH···LiF complexes are distinctly higher than those of the σC6H5C?CH···LiF complexes, indicating that the intermolecular interactions in the C6H5C?CH···LiX complexes can be mainly attributed to the π‐type interaction. QTAIM studies have shown that these lithium‐bond interactions display the characteristics of “closed‐shell” noncovalent interactions, and the molecular formation density difference indicates that electron transfer plays an important role in the formation of the lithium bond. For each site, linear relationships have been found between the topological properties at the BCP (the electron density ρb, its Laplacian ?2ρb, and the eigenvalue λ3 of the Hessian matrix) and the lithium bond length d(Li‐bond). The shorter the lithium bond length d(Li‐bond), the larger ρb, and the stronger the π···Li bond. The shorter d(Li‐bond), the larger ?2ρb, and the greater the electrostatic character of the π···Li bond. © 2012 Wiley Periodicals, Inc.  相似文献   

14.
One‐electron reduction of phenosafranine (PS+ 3,7‐diamino‐5‐phenylphenazinium chloride), a phenazinium dye, has been studied in homogeneous aqueous and sodium dodecyl sulfate (SDS) micellar media, using the pulse radiolysis technique. The various reducing radicals employed for the study in homogeneous aqueous medium were eaq?, H˙, CO2,˙?, and isopropyl ketyl radicals (CH3)2 ˙COH. Semireduced species generated by these reactions have been characterized by their absorption spectra, decay kinetics, and pKa. The one‐electron reduction potential of PS+ was determined at pH 7 in homogeneous aqueous solution employing nitrobenzene (NB/NB˙?) as the standard couple. One‐electron reduction in SDS micellar medium and a detailed spectrophotometric investigation of the parent dye in this surfactant system was carried out in order to understand the dye–surfactant interactions in the micellar and premicellar media.© 2001 John Wiley & Sons, Inc. Int J Chem Kinet 34: 56–66, 2002  相似文献   

15.
The reactions of 5,5-dimethyl-3-oxo-1-pyrroline 1-oxide (3-oxo-DMPO, 1) with NH2OH and N2H4 afforded oxime (2a) and hydrazone (2b), respectively. The reaction products were studied as spin traps for the short-lived radicals HO·, Ph·, PhCO2 ·, NC(Me2)C·, and NC(Me2)CO·. The nitroxides generated in the reactions of the above-mentioned short-lived radicals with nitrones 1 and 2a,b were characterized by ESR spectroscopy. Of these nitrones, oxime 2a is the most effective radical trap.  相似文献   

16.
The Arrhenius parameters of the bimolecular rate constants for the decay of several phenoxyl radicals in aqueous solution were measured. The p-halophenoxyl radicals (F, Cl, and Br) decay in a diffusion controlled reaction as the activation energies are the same as that of diffusion of water (16 ± 1.5 kJ · mol?1). The A factors are 1012.2 ± 0.2. For alkyl and alkoxy substituted phenoxyl, slightly higher activation energies were found (19.5 ? 21.9 kJ · mol?1). © 1993 John Wiley & Sons, Inc.  相似文献   

17.
Excessive reactive oxygen species (ROS) can oxidatively damage DNA to cause severe biological consequences. In the study, a natural flavonoid, myricitrin (myricetin‐3‐O‐α‐L‐rhamnopyranoside), was found to have a protective effect against hydroxyl‐induced DNA damage (IC50 159.86 ± 54.24 μg/mL). To investigate the mechanism, it was determined by various antioxidant assays. The results revealed that myricitrin could effectively scavenge ·OH, ·O2?, DPPH· (1,1‐diphenyl‐2‐picrylhydrazyl radical), and ABTS+· (2,2′‐Azino‐bis(3‐ethylbenzothiazoline‐6‐sulfonic acid) radicals (IC50 values were respectively 69.71 ± 5.93, 69.71 ± 5.93, 25.34 ± 2.14, and 1.71 ± 0.09 μg/mL), and bind Cu2+ (IC50 27.33 ± 2.36 μg/mL). Based on the mechanistic analysis, it can be concluded that: (i) myricitrin can effectively protect against hydroxyl‐induced DNA oxidative damage via ROS scavenging and deoxynucleotide radicals repairing approaches. Both approaches can be attributed to its antioxidant. From a structure‐activity relationship viewpoint, its antioxidant ability can be attributed to the ortho‐dihydroxyl moiety, and ultimately to the stability of its oxidized form ortho‐benzoquinone; (ii) its ROS scavenging is mediated via metal‐chelating, and direct radical‐scavenging which is through donating hydrogen (H·) and electron (e); and (iii) its protective effect against DNA oxidative damage may be primarily responsible for the pharmacological effects, and offers promise as a new therapeutic reagent for diseases from DNA oxidative damage.  相似文献   

18.
Square‐planar nickel(II) complexes of salen ligands, N,N′‐bis(3‐tert‐butyl‐(5R)‐salicylidene)‐1,2‐cyclohexanediamine), in which R=tert‐butyl ( 1 ), OMe ( 2 ), and NMe2 ( 3 ), were prepared and the electronic structure of the one‐electron‐oxidized species [ 1 – 3 ]+. was investigated in solution. Cyclic voltammograms of [ 1 – 3 ] showed two quasi‐reversible redox waves that were assigned to the oxidation of the phenolate moieties to phenoxyl radicals. From the difference between the first and second redox potentials, the trend of electronic delocalization 1 +.> 2 +.> 3 +. was obtained. The cations [ 1 – 3 ]+. exhibited isotropic g tensors of 2.045, 2.023, and 2.005, respectively, reflecting a lower metal character of the singly occupied molecular orbital (SOMO) for systems that involve strongly electron‐donating substituents. Pulsed‐EPR spectroscopy showed a single population of equivalent imino nitrogen atoms for 1 +., whereas two distinct populations were observed for 2 +.. The resonance Raman spectra of 2 +. and 3 +. displayed the ν8a band of the phenoxyl radicals at 1612 cm?1, as well as the ν8a bands of the phenolates. In contrast, the Raman spectrum of 1 +. exhibited the ν8a band at 1602 cm?1, without any evidence of the phenolate peak. Previous work showed an intense near‐infrared (NIR) electronic transition for 1 +.ν1/2=660 cm?1, ε=21 700 M ?1 cm?1), indicating that the electron hole is fully delocalized over the ligand. The broader and moderately intense NIR transition of 2 +.ν1/2=1250 cm?1, ε=12 800 M ?1 cm?1) suggests a certain degree of ligand‐radical localization, whereas the very broad NIR transition of 3 +.ν1/2=8630 cm?1, ε=2550 M ?1 cm?1) indicates significant localization of the ligand radical on a single ring. Therefore, 1 +. is a Class III mixed‐valence complex, 2 +. is Class II/III borderline complex, and 3 +. is a Class II complex according to the Robin–Day classification method. By employing the Coulomb‐attenuated method (CAM‐B3LYP) we were able to predict the electron‐hole localization and NIR transitions in the series, and show that the energy match between the redox‐active ligand and the metal d orbitals is crucial for delocalization of the radical SOMO.  相似文献   

19.
The self‐assembly of NiCl2·6H2O with a diaminodiamide ligand 4,8‐diazaundecanediamide (L‐2,3,2) gave a [Ni(C9H20N4O2)(Cl)(H2O)] Cl·2H2O ( 1 ). The structure of 1 was characterized by single‐crystal X‐ray diffraction analysis. Structural data for 1 indicate that the Ni(II) is coordinated to two tertiary N atoms, two O atoms, one water and one chloride in a distorted octahedral geometry. Crystal data for 1: orthorhombic, space group P 21nb, a = 9.5796(3) Å, b = 12.3463(4) Å, c = 14.6305(5) Å, Z = 4. Through NH···Cl–Ni (H···Cl 2.42 Å, N···Cl 3.24 Å, NH···Cl 158°) and OH···Cl–Ni contacts (H···Cl 2.36 Å, O···Cl 3.08 Å, OH···Cl 143°), each cationic moiety [Ni(C9H20N4O2) (Cl)(H2O)]+ in 1 is linked to neighboring ones, producing a charged hydrogen‐bonded 1D chainlike structure. Thermogrametric analysis of compound 1 is consistent with the crystallographic observations. The electronic absorption spectrum of Ni(L‐2,3,2)2+ in aqueous solution shows four absorption bands, which are assigned to the 3A2g3T2g, 3T2g1Eg, 3T2g3T1g, and 3A2g3T1g transitions of triplet‐ground state, distorted octahedral nickel(II) complex. The cyclic volammetric measurement shows that Ni2+ is more easily reduced than Ni(L‐2,3,2)2+ in aqueous solution.  相似文献   

20.
We report an investigation on intermolecular interactions in R? CN ··· H? OCH3 (R = H, CH3, F, Cl, NO2, OH, SH, SCH3, CHO, COCH3, CH2Cl, CH2F, CH2OH, CH2COOH, CF3, SCOCH3, SCF3, OCHF2, CH2CF3, CH2OCH3, and CH2CH2OH) complexes using density functional theory. The calculations were conducted on B3LYP/6‐311++G** level of theory for optimization of geometries of complexes and monomers. An improper hydrogen bonding (HB) in the H3CO? H ··· NC? R complexes was observed in that N atom of the nitriles functions acts as a proton acceptor. Furthermore, quantum theory of “Atoms in Molecules” (AIM) and natural bond orbital (NBO) method were applied to analyze H‐bond interactions in respective complexes. The electron density (ρ) and Laplacian (?2ρ) properties, estimated by atoms in molecules calculations, indicate that H ··· N bond possesses low ρ and positive ?2ρ values, which are in agreement with partially covalent character of the HBs, whereas O? H bonds have negative ?2ρ values. In addition, the weak intermolecular force due to dipole–dipole interaction (U) is also considered for analysis. The examination of HB in these complexes by quantum theory of NBO method fairly supports the ab initio results. Natural population analysis data, the electron density, and Laplacian properties, as well as, the ν(O? H) and γ(O? H) frequencies of complexes, calculated at the B3LYP/6‐311++G** level of theory, are used to evaluate the HB interactions. The calculated geometrical parameters and conformational analysis in water phase solution show that the H3CO? H ··· NC? R complexes in water are more stable than that in gas phase. The obtained results demonstrated a strong influence of the R substituent on the properties of complexes. Numerous correlations between topological, geometrical, thermodynamic properties, and energetic parameters were also found. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  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