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.
Kazuo Mukai  Yuji Okauchi 《Lipids》1989,24(11):936-939
A kinetic study of the reaction between a tocopheroxyl radical and unsaturated fatty acid esters has been undertaken. The rates of allylic hydrogen abstraction from various unsaturated fatty acid esters (ethyl oleate2, ethyl linoleate3, ethyl linolenate4, and ethyl arachidonate5) by the tocopheroxyl radical (5,7-diisopropyltocopheroxyl6) in benzene have been determined spectrophotometrically. The second-order rate constants, k3, obtained are 1.04×10−5 M−1s−1 for2, 1.82×10−2 M−1s−1 for3, 3.84×10−2 M−1s−1 for4, and 4.83×10−2 M−1s−1 for5 at 25.0°C. Thus, the rate constants, kabstr/H, given on an available hydrogen basis are k3/4=2.60×10−6 M−1s−1 for2, k3/2=9.10×10−3 M−1s−1 for3, k3/4=9.60×10−3 M−1s−1 for4, and k3/6=8.05×10−3 M−1s−1 for5. The kabstr/H values obtained for the polyunsaturated fatty acid esters3,4, and5 containing H-atoms activated by two π-electron systems are similar to each other, and are about three orders of magnitude higher than that for the ethyl oleate2 containing H-atoms activated by a single π-system. From these results, it is suggested that the prooxidant effect of α-tocopherol in edible oils and fats may be induced by the above hydrogen abstraction reaction.  相似文献   

2.
A kinetic study of the prooxidant effect of α-tocopherol was performed. The rates of allylic hydrogen abstraction from various unsaturated fatty acid esters (ethyl stearate 1, ethyl oleate 2, ethyl linoleate 3, ethyl linolenate 4, and ethyl arachidonate 5) by α-tocopheroxyl radical in toluene were determined, using a double-mixing stopped-flow spectrophotometer. The second-order rate constants (k p) obtained are <1 × 10−2 M−1 s−1 for 1, 1.90 × 10−2 M−1 s−1 for 2, 8.33 × 10−2 M−1 s−1 for 3, 1.92 × 10−1 M−1 s−1 for 4, and 2.43 × 10−1 M−1 s−1 for 5 at 25.0 °C. Fatty acid esters 3, 4, and 5 contain two, four, and six –CH2– hydrogen atoms activated by two π-electron systems (–C=C–CH2–C=C–). On the other hand, fatty acid ester 2 has four –CH2– hydrogen atoms activated by a single π-electron system (–CH2–C=C–CH2–). Thus, the rate constants, k abstr/H, given on an available hydrogen basis are k p/4 = 4.75 × 10−3 M−1 s−1 for 2, k p/2 = 4.16 × 10−2 M−1 s−1 for 3, k p/4 = 4.79 × 10−2 M−1 s−1 for 4, and k p/6 = 4.05 × 10−2 M−1 s−1 for 5. The k abstr/H values obtained for 3, 4, and 5 are similar to each other, and are by about one order of magnitude higher than that for 2. From these results, it is suggested that the prooxidant effect of α-tocopherol in edible oils, fats, and low-density lipoproteins may be induced by the above hydrogen abstraction reaction.  相似文献   

3.
Ozone/hydrogen peroxide batch treatment was utilized to study the degradation of the steroidal hormone estrone (E1). The competition kinetics method was used to determine the rate constants of reaction for direct ozone and E1, and for hydroxyl radicals and E1 at three pH levels (4, 7, and 8.5), three different molar O3/H2O2 ratios (1:2, 2:1, and 4:1) and a temperature about 20°C. The average second-order rate constants for direct ozone-E1 reaction were determined as 6.2?×?103?±?3.2?×?103 M?1s?1, 9.4?×?105?±?2.7?×?105 M?1s?1, and 2.1?×?107?±?3.1?×?106 M?1s?1 at pH 4, 7, and 8.5, respectively. It was found that pH had the greatest influence on the reaction rate, whereas O3/H2O2 ratio was found to be slightly statistically significant. For the hydroxyl radical-E1 reaction, apparent rate constants ranged from 1.1?×?1010 M?1s?1 to 7.0?×?1010 M?1s?1 with an average value of 2.6?×?1010 M?1s?1. Overall, O3/H2O2 is shown to be an effective treatment for E1.  相似文献   

4.
Honglian Shi  Etsuo Niki 《Lipids》1998,33(4):365-370
Owing to increasing evidence showing the importance of lipid peroxidation in oxidative stress in vivo, the role and evaluation of antioxidants have received much attention. Ginkgo biloba extract (GBE), well-known as an efficient drug against diseases induced by free radicals, has been suggested to exert its effect by antioxidant action. A method was established to determine the activity of GBE as a hydrogen donor by stoichiometric and kinetic studies, and GBE was compared with several other antioxidants such as α-tocopherol, propyl gallate, and two kinds of flavonoids which are found in GBE, quercetin, and kaempferol. It was found that there were 6.62×1019 active hydrogens in 1g of GBE. Stoichiometric studies showed that one molecule of α-tocopherol reacted with one molecule of galvinoxyl radical. For quercetin, kaempferol and propyl gallate, the experimental stoichiometric numbers were 4.0, 1.9, and 3.1, respectively. The rates of reaction of antioxidants with galvinoxyl in ethanol were determined spectrophotometrically, using a stopped-flow technique. The second-order rate constant, k 2, obtained at 25°C was 0.13 (g/l)−1s−1 for GBE and 5.9×103, 2.1×103, 1.2×104, and 2.4×103 M−1s−1 for quercetin, kaempferol, propyl gallate, and α-tocopherol, respectively. The second-order rate constant, k 2′, on the molar basis of active hydroxyl groups in the tested substances obtained at 25°C decreased in the order of propyl gallate > α-tocopherol> quercetin>GBF∼kaempferol. This is the first study on GBE as an antioxidant which reports both stoichiometric and kinetic results.  相似文献   

5.
The chemical stability of perfluorinated and non‐perfluorinated low temperature fuel cell model compounds (MCs) against attack by hydroxyl radicals, HO, is compared using a competition kinetics approach in aqueous solutions at ambient temperature. HO radicals were generated in situ by UV photolysis of hydrogen peroxide in the electron spin resonance (ESR) resonator. Acetic acid (AA), trifluoroacetic acid (TFAA), methanesulfonic acid (MSA), trifluorosulfonic acid (TFSA), and perfluoro(2‐ethoxyethane)sulfonic acid (PFEESA) were chosen as MCs, while the rate constants of 5,5‐dimethyl‐1‐pyrroline‐N‐oxide (DMPO) and methanol (CH3OH) served as reference for the determination of relative rate constants by means of steady state ESR signal amplitudes. In decreasing order the rate constants are: kMSA = (4.8 ± 0.2) × 107 M–1 s–1, kAA = (4.2 ± 0.3) × 107 M–1 s–1, kPFEESA = (3.7 ± 0.1) × 106 M–1 s–1, kTFAA = (7.9 ± 0.2) × 105 M–1 s–1, and kTFSA < 1.0 × 105 M–1 s–1. Applying these results to perfluorinated fuel cell membranes like Nafion®, the main points of attack by HO are concluded to be the ether groups of the side chains, followed by the remaining carboxyl groups from the manufacturing process of the polymers.  相似文献   

6.
Oxidation kinetics of selected pharmaceutical compounds and their degradation during ozonation of secondary treated municipal wastewater effluent (MWWE) was investigated. The apparent second-order rate constants for the reaction between chlorotetracycline (CTC), enrofloxacin (ENR), gemfibrozil (GEM) and ozone ranged between 6.82 – 52.7 × 104 M?1s?1. The measured second-order hydroxyl radical rate constants were several orders of magnitude higher at 8.4 × 109 – 13.1 × 109 M?1s?1 with a reactivity sequence of GEM > CTC > ENR. Overall degradation of CTC, ENR and GEM in secondary treated municipal wastewater effluent was >76 % at ozone doses of 0.33 mg O3/mg DOC or higher.  相似文献   

7.
The kinetic model of asymmetric reduction of 3-oxo-3-phenylpropionic acid ethyl ester using Saccharomyces cerevisiae CGMCC No.2266 with 10% glucose as co-substrate to realize the regeneration of NADPH was established. The effect factors on reduction, the type and the content of co-substrate and coenzyme, and the changes of the substrate and product content vs. time during the reaction process were investigated. The results indicate that 10% glucose can increase the reaction conversion from 23.0% to 98.4% and NADPH is reducer. The reduction process conforms with sequence mechanisms. The model parameters are as follows: v m =5.0×10−4 mol·L−1·h−1, k1=1.5×10−6 mol·L−1·h−1, k2=3.0×10−3 mol·L−1·h−1. The kinetic model is in good agreement with the experimental data.  相似文献   

8.
The water vapor (WVP) and oxygen (O2P) permeabilities of beeswax (BW), candelilla wax (CnW), carnauba wax (CrW) and microcrystalline wax (MW), formed as freestanding films, were determined. CnW and CrW both had small values for O2P (0.29 and 0.26 g·m−1·sec−1·Pa−1 × 10−14, respectively), which are less than half the value for high-density polyethylene and about a decade greater than the value for polyethylene terephthalate. O2P values for BW and MW were about 6−9× greater than those of CnW and CrW. WVP of CnW was 0.18 g·m−1·sec−1·Pa−1 × 10−12, which is about one-half the value for CrW and MW and about one-third the value for BW. The WVP of CnW was somewhat less than that of polypropylene and somewhat greater than that of high-density polyethylene. Differences in permeabilities among the wax films are attributed mainly to differences in chemical composition and crystal type as determined by X-ray diffraction.  相似文献   

9.
Potassium and acid ferrierites were impregnated with boron species by wet and incipient wetness techniques. All samples display a medium-intensity band at 3,450–3,470 cm−1 associated to Si−OH···O groups corresponding to boron-containing units. The 1,398–1,404 cm−1 band assigned to the B–O stretching in BO3 units does not appear on boron–potassium–ferrierite prepared by wet impregnation. Catalytic performance during the linear butene skeletal isomerization was measured. At 300 °C, boron impregnated by incipient wetness technique on acid ferrierite reduces both linear butene conversion at a short time and isobutene yield in all time range. Boron–potassium–ferrierite prepared by wet impregnation has a suitable acidity to promote isobutene production. At 450 °C, this sample shows the best performance, being the isobutene yield 1.7 times higher than the acid-ferrierite one and reaching the highest isobutene selectivity (92%). This performance is maintained with time. Both isobutene yield and by-product distribution are strongly affected by temperature; dimer intermediates are formed. Finally, both kinds of hydroxyl groups corresponding to 3,466 and 3,635 cm−1 bands influence the isobutene production whereas BO3 sites are inactive for this reaction.  相似文献   

10.
A novel cyano-bridged heterotrinuclear FeIII/CoII/FeIII complex, [Co(phen)2][Fe(phen)(CN)4]2·4H2O (1) (phen = 1,10-phenanthroline), has been synthesized under solvothermal conditions, and characterized by elemental analysis, FI-IR spectrum and single-crystal X-ray diffraction. The structure analysis reveals complex 1 is constructed into a 3D supramolecular architecture through abundant intermolecular non-covalent interactions. According to the supramolecular self-assembly of complex 1, five types of π–π stacking models of phenanthroline ligands, hydrogen bonds formed by four lattice water molecules, and C–H···N weak interactions involving cyano groups are all observed. Electronic absorption spectroscopy and fluorescence titration studies of the interaction between complex 1 and calf thymus DNA are suggestive of an intercalative binding mode with an intrinsic binding constant of 7.95 × 103 M−1 and a linear Stern–Volmer quenching constant of 1.033 × 105 M−1.  相似文献   

11.
A novel real-time method was developed to evaluate the bisphenol A degradation kinetics by molecular ozone and radical pathway using the stopped-flow technique. The second-order kinetics was determined under pseudo-first-order conditions for the molecular pathway by the absolute rate constant method and for the radical pathway by the Rct concept involving the hydroxyl radical and ozone ratio. Bisphenol A degradation by ozone was performed and evaluated at a pH ranging from 2 to 10. At pH?4?M?1?s?1 and for the radical pathway at pH?>?10, the constant was 3.43?×?109?M?1?s?1. To validate the method, ciprofloxacin degradation kinetics was determined at pH 8 by radical pathway, in 4.55?×?109?M?1?s?1. The method permits the determination of kinetic parameters for the design of chemical reactors; avoiding the generation of undesirable reactions and by-products in the degradation of emerging compounds.  相似文献   

12.
The 1O2 quenching rate constants (k Q ) of α-tocopherol (α-Toc) and carotenoids such as β-carotene, astaxanthin, canthaxanthin, and lycopene in liposomes were determined in light of the localization of their active sites in membranes and the micropolarity of the membrane regions, and compared with those in ethanol solution. The activities of α-Toc and carotenoids in inhibiting 1O2-dependent lipid peroxidation (reciprocal of the concentration required for 50% inhibition of lipid peroxidation: [IC50]−1) were also measured in liposomes and ethanol solution and compared with their k Q values. The k Q and [IC50]−1 values were also compared in two photosensitizing systems containing Rose bengal (RB) and pyrenedodecanoic acid (PDA), respectively, which generate 1O2 at different sites in membranes. The k Q values of α-Toc were 2.9×108M−1s−1 in ethanol solution and 1.4×107 M−1s−1 (RB system) or 2.5×106 M−1s−1 (PDA system) in liposomes. The relative [IC50]−1 value of α-Toc in liposomes was also five times higher in the RB system than in the PDA-system. In consideration of the local concentration of the OH-group of α-Toc in membranes, the k Q value of α-Toc in liposomes was recalculated as 3.3×106 M−1s−1 in both the RB and PDA systems. The k Q values of all the carotenoids tested in two photosensitizing systems were almost the same. The k Q value of α-Toc in liposomes was 88 times less than in ethanol solution, but those of carotenoids in liposomes were 600–1200 times less than those in ethanol solution. The [IC50]−1 value of α-Toc in liposomes was 19 times less than that in ethanol solution, whereas those of carotenoids in liposomes were 60–170 times less those in ethanol solution. There were no great differences (less than twice) in the k q and [IC50]−1 values of any carotenoids. The k Q values of all carotenoids were 40–80 times higher than that of α-Toc in ethanol solution but only six times higher that of α-Toc in liposomes. The [IC50]−1 values of carotenoid were also higher than that of α-Toc in ethanol solution than in liposomes, and these correlated well with the k Q values.  相似文献   

13.
A new cyano-bridged copper (I) complex {[Cu(phen)CN][Cu(phen)][Cu(CN)2]} n (1) (phen = 1,10-phenanthroline) was synthesized under the solvothermal conditions, and characterized by elemental analysis, FI-IR spectra and single-crystal X-ray diffraction. The structure analysis reveals that complex 1 contains two [Cu(phen)CN][Cu(phen)][Cu(CN)2] subunits, being, respectively, constructed into two chains with different cyano backbones. The resulting chains offer a double curvy chain by Cu···Cu weak interactions and π–π stacking interactions. It is exciting that the double curvy chains adopt anti-parallel array and result in a novel 3D architecture through π–π stacking interactions. According to the supramolecular self-assembly of complex 1, four types of stacking mode of phenanthroline moieties are given and discussed. Absorption and fluorescence titration studies of complex 1 with calf thymus DNA are suggestive of the intercalation binding mode with a intrinsic binding constant of 7.52 × 103 M−1 and a linear Stern-Volmer quenching constant of 1.02 × 105 M−1.  相似文献   

14.
Evidence is presented that cardiolipin, a naturally occurring phospholipid, inhibits the aggregatory effect of platelet-activating factor (paf) on rabbit plateletsin vitro. Bovine heart cardiolipin was shown to inhibit the aggregation of washed rabbit platelets induced by 1×10−10 M and 2×10−10 M paf with IC50 values (doses for half-maximal inhibition) of 8.4±0.8×10−7 M and 2.6±0.6×10−6 M, respectively. Phosphonocardiolipin was also able to inhibit platelet aggregation induced by 1× 10−10 M paf with an IC50 value of 3±1×10−7M. Both compounds, in concentrations up to 1×10−5 M, were unable to aggregate washed rabbit platelets and failed to inhibit the aggregation induced by 0.9 and 1.8 μM adenosine diphosphate or 0.2–1.0 μM arrchidonic acid. By contrast, the acetylated derivative of cardiolipin exerted an aggregatory effect on aspirin-treated rabbit platelets in the presence of creatine phosphate/creatine phosphokinase. This aggregation was inhibited by the specific paf antagonists BN 52021 and WEB 2086. Also, platelets treated with acetyl-cardiolipin were insensitive to the aggregatory effect of paf. Phosphatidic acid, phosphatidylglycerol,bis(dipalmitoylglycero)phosphate and their phosphono analogues were totally inactive. Similar data were obtained when platelet-rich plasma was used instead of washed rabbit platelets. Our results support the hypothesis that the effect of cardiolipin is mediated through specific paf receptors that act on the rabbit platelet membrane.  相似文献   

15.
A kinetic study of the prooxidant effect of vitamin E (tocopherol, TocH) has been carried out. The rates of hydrogen abstraction (k−1) from methyl linoleate hydroperoxide (ML-OOH) by α-tocopheroxyl (α-Toc.) (1) and eight types of alkyl substituted Toc. radicals, (2–9) in benzene solution have been determined spectrophotometrically. The results show that the rate constants decrease as the total electron-donating capacity of the alkyl substituents on the aromatic ring of Toc. increases. The k−1 value (5.0×10−1M−1s−1) obtained for α-Toc. (1) was found to be about seven orders of magnitude lower than the k1 value (3.2×106M−1s−1) for the reaction of α-TocH with peroxyl radical, which is well known as the usual radical-scavenging reaction of α-TocH. The above reaction rates (k−1) obtained were compared with those (k3) of methyl linoleate with Toc. (1–9) in benzene solution. The rates (k−1) were found to be about six times larger than those (k3) of the corresponding Toc.. The results suggest that both reactions may relate, to the prooxidant effect of α-TocH at high concentrations in foods and oils. The effect of the phytyl side chain on the reaction rate, of Toc. in micellar dispersions has also been studied. We have measured the rate constant, k−1, for the reaction of phosphatidylcholine hydroperoxide with a Toc. radical in benzene,tert-butanol and in Triton X-100 micellar dispersions, and compared the observed k−1 values with the corresponding values for ML-OOH.  相似文献   

16.
The influence of sodium fluoride additives on the physicochemical properties of glasses in the Na2O-B2O3 systems is investigated. The introduction of sodium fluoride into the Na2O · 2B2O3 and Na2O · 3B2O3 glasses leads to an increase in the electric conductivity. The temperature-concentration dependence of the electric conductivity has been investigated. It is shown that, in glasses of the NaF-Na2O · 2B2O3 system, an increase in the volume concentration of sodium ions from 2.4 × 10−2 to ∼3 × 10−2 mol/cm3 is accompanied by an insignificant decrease in the activation energy Eσ from 1.44 to 1.38 eV and a sharp (by a factor of ∼30) increase in the electric conductivity. In glasses of the NaF-Na2O · 3B2O3 system, an increase in the concentration of sodium ions from 1.8 × 10−2 to ∼2.3 × 10−2 mol/cm3 brings about an increase in the electric conductivity by a factor of approximately 100 and an increase in Eσ from 1.6 to 1.83 eV. A further increase in the concentration of sodium ions (up to 2.5 × 10−2 mol/cm3) virtually does not affect the electric conductivity and Eσ. At the same concentration of sodium ions (∼2.46 × 10−2 mol/cm3) in the 9.8NaF · 90.2[Na2O · 2B2O3] and 57.1NaF · 42.9[Na2O · 3B2O3] glasses, the electric conductivity and the activation energy are considerably higher in the glass with a larger fluorine content. The regularities revealed are interpreted in the framework of the microinhomogeneous glass structure.  相似文献   

17.
The influence of sodium fluoride additives on the physicochemical properties of glasses in the Na2O-B2O3 systems is investigated. The introduction of sodium fluoride into the Na2O · 2B2O3 and Na2O · 3B2O3 glasses leads to an increase in the electric conductivity. The temperature-concentration dependence of the electric conductivity has been investigated. It is shown that, in glasses of the NaF-Na2O · 2B2O3 system, an increase in the volume concentration of sodium ions from 2.4 × 10−2 to ∼3 × 10−2 mol/cm3 is accompanied by an insignificant decrease in the activation energy Eσ from 1.44 to 1.38 eV and a sharp (by a factor of ∼30) increase in the electric conductivity. In glasses of the NaF-Na2O · 3B2O3 system, an increase in the concentration of sodium ions from 1.8 × 10−2 to ∼2.3 × 10−2 mol/cm3 brings about an increase in the electric conductivity by a factor of approximately 100 and an increase in Eσ from 1.6 to 1.83 eV. A further increase in the concentration of sodium ions (up to 2.5 × 10−2 mol/cm3) virtually does not affect the electric conductivity and Eσ. At the same concentration of sodium ions (∼2.46 × 10−2 mol/cm3) in the 9.8NaF · 90.2[Na2O · 2B2O3] and 57.1NaF · 42.9[Na2O · 3B2O3] glasses, the electric conductivity and the activation energy are considerably higher in the glass with a larger fluorine content. The regularities revealed are interpreted in the framework of the microinhomogeneous glass structure.  相似文献   

18.
Recent occurrences of pharmaceutical antibiotics in surface water, drinking water, and wastewater systems have gained significant attention due to their potential threats to human health. This study determined the absolute second-order rate constants of ozone with two amine-based antibiotics, namely, lincomycin and spectinomycin, using the stopped-flow technique under controlled ionic strength, buffer, and temperature. Results indicate that ozone reacts quickly with the selected antibiotics, and the reaction rate significantly depends on solution pH. For lincomycin, ozone attacks its free amine group and sulfur group with absolute rate constants of 2.76 × 106 M?1·s?1 (for neutral form) and 3.26 × 105 M?1·s?1 (for monoprotonated form), respectively. For spectinomycin, ozone attacks two free amine groups with absolute rate constants of 1.27 × 106 M?1·s?1 (for neutral form) and 3.30 × 105 M?1·s?1 (for monoprotonated form), respectively. These rate constants have been corrected to zero ionic strength. Protonated amine is nonreactive toward ozone. Model prediction indicates that lincomycin and spectinomycin can be effectively transformed by ozonation processes around neutral pH.  相似文献   

19.
This study investigated the oxidation and reaction kinetics of biocide o-phenylphenol (o-PP) during ozonation. The second-order rate constants for direct reaction of molecular ozone with o-PP were determined in homogenous system using classical and competition kinetics. Obtained values of the second-order rate at pH 2 are equal to (3.79 ± 0.23)?×?102 M?1 s?1 and (4.42 ± 0.64)?×?102 M?1 s?1, for the two methods, respectively. The rate constant for the dissociated form of o-PP was also determined. It was found that the rate constants for the reaction of o-PP with ozone increase significantly with increasing pH.  相似文献   

20.
The ac response of polyaniline thin films on platinum electrodes was measured at different dc potentials during the N2-fixation in methanol + LiClO4 electrolyte with 0.03 mol L−1 H2SO4 for the first time. The optimum film thickness was found to be 1.5 μm, N2-pressure 50 bar and an optimum electrolysis potential of −0.12 V (NHE). The diffusion coefficients for N2 into the polymer film was found to be (5 ± 2)×10−9 cm2 s−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