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 31 毫秒
1.
A series of quinone-based compounds were tested for their ability to act as external electron acceptors in the 1-dehydrogenation of-α-methyl-hydrocortisone-21-acetate, with polyurethane-entrapped Arthrobacter simplex cells in buffer-saturated n-decan-1-ol. This organic solvent was needed to solubilize the steroid substrate. In aqueous medium, the conversion with free cells virtually stopped after one hour, probably due to substrate limitation. All the tested quinones acted as external electron acceptors, increasing the bioconversion rate. The process kinetics were complex. However, when keeping the concentration of one of the substrates (steroid or quinone) constant and varying that of the other, Michaelis-Menten kinetics provided a reasonably good model for the initial reaction rates, and apparent kinetic constants were estimated. The most effective of the tested external electron acceptors were 2,6-dimethyl-p-benzoquinone and menadione. Mass transfer limitations seemed to appear after some hours of reaction, with low concentrations of the more efficient quinones, when the biocatalyst microenvironment was quinone- and possibly oxygen-depleted. Monosodium glutamate was included with the cells in the immobilisation foam, as an activity-stabilizing agent. It was observed that some of the quinones apparently formed complexes with this glutamate, thereby influencing the kinetics of the process. The catalytic half-life of the system depended on the quinone concentration and optimal values (60-80 h) were observed at 1 mM levels of 2,6-dimethyl-p-benzoquinone or menadione. Quinone toxicity, direct or through the formation of peroxides in the aerobic reoxidation process, may be at the origin of enzyme deactivation.  相似文献   

2.
The reaction between p-benzoquinone and H2O2 in slightly alkaline solutions yields three major quinoid products that accumulate in the reaction mixture: (a) 2,3-epoxy-p-benzoquinone, (b) 2-hydroxy-p-benzoquinone and (c) p-benzohydroquinone. The reaction is accompanied by photoemission, probably originating from excited triplet 2-hydroxy-p-benzoquinone. These products originate from hydrogen peroxide and hydroxide nucleophilic addition to the C2?C3 double bond, as well as secondary redox interactions. The hydroxy substituent and the epoxide ring exert a substantial influence on the electronic distribution in the p-benzoquinone molecule leading to a decrease in the half-wave potential, as compared to the parent p-benzoquinone. The generation of electronically excited states is the result of reactions secondary to the nucleophilic additions involving 2-hydroxy-p-benzosemiquinone, H2O2 and hydroxyl radical. The process involves the primary oxidation of 2-hydroxy-p-benzosemiquinone by hydrogen peroxide, followed by oxidation of the semiquinone by hydroxyl radical leading to the formation of the electronically excited quinone. The decay of the excited triplet to the ground state is accompanied by photoemission with maximal intensity at 485–530 nm. Thermodynamic calculations along with an observed increase of photoemission intensity in anaerobiosis point to the triplet (n, π*) multiplicity of the excited state. The efficiency of chemiluminescence could be calculated as 10?8 photons/2-hydroxy-p-benzoquinone molecule formed. Photoemission arising from the p-benzoquinone/H2O2 reaction was inhibited efficiently by addition of GSH to the reaction mixture. This may be due to deactivation of the triplet quinone by a 2-glutathionyl-p-benzohydroquinone adduct, involving thioether α-hydrogen atom-transfer to the triplet ketone.  相似文献   

3.
In this study, the abilities of two microbial consortia (Y and F) to degrade aliphatic–aromatic hydrocarbon mixtures were investigated. Y consortium preferentially degraded the aromatic hydrocarbon fractions in kerosene, while F consortium preferentially degraded the aliphatic hydrocarbon fractions. Degradation experiments were performed under aerobic conditions in sealed bottles containing liquid medium and n-octane or n-decane as representative aliphatic hydrocarbons or toluene, ethylbenzene or p-xylene as representative aromatic hydrocarbons (all at 100 mg/l). Results demonstrated that the Y consortium degraded p-xylene more rapidly than n-octane. It degraded toluene, ethylbenzene and p-xylene more rapidly than decane. In comparison, the F consortium degraded n-octane more rapidly than toluene, ethylbenzene or p-xylene, and n-decane more rapidly than toluene, ethylbenzene or p-xylene. 16S rRNA gene sequencing revealed that the Y consortium was dominated by Betaproteobacteria and the F consortium by Gammaproteobacteria, and in particular Pseudomonas. This could account for their metabolic differences. The substrate preferences of the two consortia showed that the aliphatic–aromatic hydrocarbon binary mixtures, especially the n-decane–toluene/ethylbenzene/p-xylene pairs, reflected their degradation ability of complex hydrocarbon compounds such as kerosene. This suggests that aliphatic–aromatic binary systems could be used as a tool to rapidly determine the degradation preferences of a microbial consortium.  相似文献   

4.
The relationship of exhaled ethane and n-pentane to exhaled NO, carbonylated proteins, and indoor/outdoor atmospheric pollutants were examined in order to evaluate ethane and n-pentane as potential markers of airway inflammation and/or oxidative stress. Exhaled NO and carbonylated proteins were found to have no significant associations with either ethane (p = 0.96 and p = 0.81, respectively) or n-pentane (p = 0.44 and 0.28, respectively) when outliers were included. In the case where outliers were removed n-pentane was found to be inversely associated with carbonylated proteins. Exhaled hydrocarbons adjusted for indoor hydrocarbon concentrations were instead found to be positively associated with air pollutants (NO, NO2 and CO), suggesting pollutant exposure is driving exhaled hydrocarbon concentrations. Given these findings, ethane and n-pentane do not appear to be markers of airway inflammation or oxidative stress.  相似文献   

5.
The quinones 1,4-naphthoquinone (NQ), methyl-1,4-naphthoquinone (MNQ), trimethyl-1,4-benzoquinone (TMQ) and 2,3-dimethoxy-5-methyl-1,4-benzoquinone (UQ-0) enhance the rate of nitric oxide (NO) reduction by ascorbate in nitrogen-saturated phosphate buffer (pH 7.4). The observed rate constants for this reaction were determined to be 16±2,215±6,290±14 and 462±18?M-1?s-1, for MNQ, TMQ, NQ and UQ-0, respectively. These rate constants increase with an increase in quinone one-electron redox potential at neutral pH, E71. Since NO production is enhanced under hypoxia and under certain pathological conditions, the observations obtained in this work are very relevant to such conditions.  相似文献   

6.
7.
The brown-rot basidiomycete Gloeophyllum trabeum uses a quinone redox cycle to generate extracellular Fenton reagent, a key component of the biodegradative system expressed by this highly destructive wood decay fungus. The hitherto uncharacterized quinone reductase that drives this cycle is a potential target for inhibitors of wood decay. We have identified the major quinone reductase expressed by G. trabeum under conditions that elicit high levels of quinone redox cycling. The enzyme comprises two identical 22-kDa subunits, each with one molecule of flavin mononucleotide. It is specific for NADH as the reductant and uses the quinones produced by G. trabeum (2,5-dimethoxy-1,4-benzoquinone and 4,5-dimethoxy-1,2-benzoquinone) as electron acceptors. The affinity of the reductase for these quinones is so high that precise kinetic parameters were not obtainable, but it is clear that kcat/Km for the quinones is greater than 108 M−1 s−1. The reductase is encoded by a gene with substantial similarity to NAD(P)H:quinone reductase genes from other fungi. The G. trabeum quinone reductase may function in quinone detoxification, a role often proposed for these enzymes, but we hypothesize that the fungus has recruited it to drive extracellular oxyradical production.  相似文献   

8.
The bacterium, Rhodococcus erythropolis H-2, which can utilize dibenzothiophene (DBT) as a sole source of sulfur in the presence of hydrocarbon, was isolated from soil samples. When this strain was cultivated in a medium containing 0.27 mM DBT and 40% n-tetradecane, DBT was metabolized stoichiometrically to 2-hydroxybiphenyl within 1 day. This strain grew in the presence of n-octane and longer-carbonchain hydrocarbons, but not with n-hexane, styrene, p-xylene, cyclooctane or toluene. DBT degradation proceeded in the resting cell system with lyophilized cells of this strain. The addition of n-tetradecane enhanced the reaction rate, the optimal concentration being 40%. DBT degradation occurred in the reaction mixture even in the presence of 70% n-tetradecane, whereas at concentrations above 80% n-tetradecane suppressed the degradation.  相似文献   

9.
In addition to α-tocopherolquinone and 5α-stigmast-9(11)-en-3β-ol, two new quinones have been isolated from the seeds of Costus speciosus and characterized as 6-methyl dihydrophytylplastoquinone (2,5,6-trimethyl-3-(3,7,11,15-tetramethylhexadecyl)-1,4-benzoquinone) and dihydrophytylplastoquinone (5,6-dimethyl-3-(3,7,11,15-tetramethylhexadecyl)-1,4-benzoquinone) respectively by physical data and chemical studies.  相似文献   

10.
p-Benzoquinone non-enzymatically reacted with d-glucosamine at physiological pH and moderate temperature. The reaction of p-benzoquinone with glucosamine was signaled by changes in the UV and visible spectra. The reactivity proceeded fastest at pH values above 7, with a sharp drop from pH 6.5 to 7.0, and the reaction was negligible in acidic conditions. The order of reactivity of amino sugars was d-mannosamine > d-glucosamine > d-galactosamine. From the reaction mixture, four conversion products were isolated and none was toxic to Escherichia coli even at 500–700 g ml–1, while p-benzoquinone was cytotoxic to E. coli at 20 g ml–1. Chitosan could react with p-benzoquinone efficiently and remove this toxicant in aqueous solution.  相似文献   

11.
We have investigated the correlation between DNA adduct levels and aryl hydrocarbon hydroxylase (AHH) activity in peripheral lymphocyte samples obtained from 42 lung cancer patients. DNA adducts and AHH activity were determined by the 32P-postlabelling technique and the fluorometric method, respectively. The mean +/- SD of DNA adduct level was 0.88 +/- 0.37 (ranged from 0.22 to 1.90) per 108 nucleotides. The geometric means of non-induced and 3-methylcholanthrene (MC)-induced AHH activity, as well as AHH inducibility (MC-induced AHH activity/non-induced AHH activity) were 0.029, 0.228 pmol min-1 10-6 cells, and 7.776, respectively. There was no statistically significant correlation between DNA adduct levels and non-induced or MC-induced AHH activity. A tendency of positive correlation was found between DNA adduct levels and AHH inducibility for the all subjects (n = 42, r = 0.25, p = 0.11). Such a positive correlation reached statistical significance in the subjects with squamous cell carcinoma (n = 13, r = 0.70, p < 0.01). In addition, similar correlation of DNA adducts with AHH inducibility was also observed in the GSTM1 present genotype (n = 17, r = 0.44, p = 0.07) and GSTP1-AA genotype (n = 29, r = 0.37, p = 0.05) individuals. These findings suggest that DNA adduct levels are mediated by CYP1A1 enzyme, and AHH inducibility may be a more relevant indicator than specific AHH activity for explaining the variation of DNA adduct levels in lymphocytes.  相似文献   

12.
Synopsis Out of a total of eleven bifunctional reagents tested as fixatives for light microscopic immunohistochemistry, four were found satisfactory when applied in the vapour phase to freeze-dried blocks. These were diethylpyrocarbonate, as observed in carlier studies, dimethyladipimidate,p-benzoquinone, and diacetyl. Adequate but less satisfactory liquid-phase fixation was provided by three reagents (dimethyladipimidate, dimethylsuberimidate,p-benzoquinone).None of the eleven reagents gave optimal preservation of antigenicity and structure when tested at the ultrastructural level. Encouraging results were obtained withp-benzoquinone, however.  相似文献   

13.
14.
Glucose oxidase was immobilized on the surface of a graphite electrode by irreversible adsorption. An electrocatalytic steady-state current for the oxidation of D-glucose was observed using this electrode in the presence of p-benzoquinone as an electron transfer mediator. The electrocatalytic current at 0.5 V vs. SCE was analyzed as a function of the concentrations of D-glucose and p-benzoquinone, and the maximum current, Ismax, and the Michaelis constants (K1 and K2 for D-glucose and p-benzoquinone, respectively) of the electrocatalysis were determined. The dependence of the current on the electrode potential, pH, and temperature was also investigated. The results indicate that the kinetics of the immobilized enzyme are essentially the same as those of the enzyme in the solubilized state. The effect of various electron transfer mediators on the electrocatalytic current was also examined and evaluated in terms of Ismax, K1, and K2 values.  相似文献   

15.
The photoacoustic technique was used to measure energy storage by cyclic electron transfer around photosystem I in intact Chlamydomonas reinhardtii cells illuminated with far-red light (>715 nm). The in-vivo cyclic pathway was characterized by investigating the effects of various chemicals on energy storage. Participation of plastoquinone and ferredoxin in the cyclic electron flow was confirmed by the complete suppression of energy storage in the presence of the plastoquinol antagonist 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIB) and the ferredoxin inhibitors/competitors methylviologen, phenylmercuric acetate and p-benzoquinone. Two alternative electron cycles are demonstrated to operate in vivo. One cycle is sensitive to antimycin A, myxothiazol and 2-(n-heptyl)-4-hydroxyquinoline N-oxide (HQNO) and is catalyzed by ferredoxin which reduces plastoquinone through a route involving cytochrome b 6 and its protonmotive Q-cycle. The other cycle is unaffected by the above-mentioned inhibitors but is sensitive to N-ethylmaleimide (NEM), an inhibitor of the ferredoxin-NADP reductase, and 2-monophosphoadenosine-5-diphosphoribose (PADR), an analogue of NADP, showing that the electron recycling was mediated by NADPH. Possibly, electrons enter the plastoquinone pool through the action of a NAD(P)H dehydrogenase, which is insensitive to classical inhibitors of the mitochondrial NADH dehydrogenase. Loss of energy storage by photosystem-I-driven cyclic electron transfer in farred light was observed only when antimycin A, myxothiazol or HQNO was used in combination with NEM or PADR. Analysis of the light-intensity dependence and the rate of in-vivo cyclic electron transfer in the presence of various inhibitors indicates that the NADPH-dependent electron-cycle is the preferential cyclic pathway in Chlamydomonas cells illuminated with far-red light.Abbreviations Amax maximal photothermal signal - Cyt cytochrome - DBMIB 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone - DCMU (diuron) 3-(3,4-dichlorophenyl)-1,1-dimethylurea - ES photochemical energy storage - FNR ferredoxin NADP+ reductase - HQNO 2-(n-heptyl)-4-hydroxyquinoline N-oxide - NEM N-ethylmaleimide - P700 reaction-center pigment of PSI - PADR 2-monophosphoadenosine-5-diphosphoribose - pBQ p-benzoquinone - PMA phenylmercuric acetate We are very grateful to Dr. M.-H. Montane (Cadarache, Saint-Paul-lez-Durance, France) for her advice in the electroporation experiments.  相似文献   

16.
The functional state of the PS II population localized in the stroma exposed non-appressed thylakoid region was investigated by direct analysis of the PS II content of isolated stroma thylakoid vesicles. This PS II population, possessing an antenna size typical for PS II, was found to have a fully functional oxygen evolving capacity in the presence of an added quinone electron acceptor such as phenyl-p-benzoquinone. The sensitivity to DCMU for this PS II population was the same as for PS II in control thylakoids. However, under more physiological conditions, in the absence of an added quinone acceptor, no oxygen was evolved from stroma thylakoid vesicles and their PS II centers were found to be incapable to pass electrons to PS I and to yield NADPH. By comparison of the effect of a variety of added quinone acceptors with different midpoint potentials, it is concluded that the inability of PS II in the stroma thylakoid membranes to contribute to NADPH formation probably is due to that QA of this population is not able to reduce PQ, although it can reduce some artificial acceptors like phenyl-p-benzoquinone. These data give further support to the notion of a discrete PS II population in the non-appressed stroma thylakoid region, PS II, having a higher midpoint potential of QA than the PS II population in the appressed thylakoid region, PS II. The physiological significance of a PS II population that does not produce any NADPH is discussed.Abbreviations pBQ p-benzoquinone - Chl chlorophyll - DCBQ 2,6-dichloro-p-benzoquinone - DCIP 2,6-dichloroindophenol - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - DMBQ 2,5-dimethyl-p-benzoquinone - DQ duroquinone(tetramethyl-p-benzoquinone) - FeCN ferricyanide (potassium hexacyanoferrat) - MV methylviologen - NADPH,NADP+ reduced or oxidized form of nicotinamide adenine dinucleotide phosphate respectively - PpBQ phenyl-p-benzoquinone - PQ plastoquinone - PS II photosystem II - PS I photosystem I - QA primary quinone acceptor of PS II - QB secondary quinone acceptor of PS II - E microEinstein  相似文献   

17.
Corynebacterium cultures, isolated from soil by the use of n-paraffin as the sole carbon source have been shown to bring about significant oxidation of a new compound, dl-cistetrahydro-2-oxo-4-n-pentyl-thieno-(3,4-d)-imidazoline (dl-TOPTI), in the presence of hydrocarbon.

The co-oxidation products were isolated from culture broth and identified as dl-biotinol and dl-biotin by IR, NMR and MS spectroscopies.  相似文献   

18.
Succinate:menaquinone oxidoreductase from Corynebacterium glutamicum, a high-G+C, Gram-positive bacterium, was purified to homogeneity. The enzyme contained two heme B molecules and three polypeptides with apparent molecular masses of 67, 29 and 23 kDa, which corresponded to SdhA (flavoprotein), SdhB (iron–sulfur protein), and SdhC (membrane anchor protein), respectively. In non-denaturating polyacrylamide gel electrophoresis, the enzyme migrated as a single band with an apparent molecular mass of 410 kDa, suggesting that it existed as a trimer. The succinate dehydrogenase activity assayed using 2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone and 2,6-dichloroindophenol as the electron acceptor was inhibited by 2-n-heptyl-4-hydroxyquinoline N-oxide (HQNO), and the Dixon plots were biphasic. In contrast, the succinate dehydrogenase activity assayed using phenazine methosulfate and 2,6-dichloroindophenol was inhibited by p-benzoquinone and not by HQNO. These findings suggested that the C. glutamicum succinate:menaquinone oxidoreductase had two quinone binding sites. In the phylogenetic tree of SdhA, Corynebacterium species do not belong to the high-G+C group, which includes Mycobacterium tuberculosis and Streptomyces coelicolor, but are rather close to the group of low-G+C, Gram-positive bacteria such as Bacillus subtilis. This situation may have arisen due to the horizontal gene transfer.  相似文献   

19.
 Degradation of tetrachloroguaiacol is catalyzed by an extracellular enzyme, the laccase of the white-rot fungus Coriolus versicolor. This enzyme catalyzes the dechlorination of tetrachloroguaiacol and release of chloride ions. The pathway for the degradation was deduced from the intermediates produced by purified laccase and 18O-labeling experiments. The first step is demethylation. The resulting tetrachlorocatechol is dechlorinated to give 2,3,5-trichloro-6-hydroxy-p-benzoquinone, 2,5-dichloro-3,6-dihydroxy-p-benzoquinone, and dichloro-6-hydroxy-p-benzoquinone. Isotopic experiments established the mechanism of dechlorination of tetrachloroguaiacol by laccase. The laccase-catalyzed dechlorination is not caused by oxidative coupling but by nucleophilic substitution in which Cl- is released by water from cation radicals generated by laccase. Received: 25 August 1995/Received revision: 27 October 1995/Accepted: 20 November 1995  相似文献   

20.
The sodium ion-translocating NADH:quinone oxidoreductase (Na+-NQR) from the pathogen Vibrio cholerae exploits the free energy liberated during oxidation of NADH with ubiquinone to pump sodium ions across the cytoplasmic membrane. The Na+-NQR consists of four membrane-bound subunits NqrBCDE and the peripheral NqrF and NqrA subunits. NqrA binds ubiquinone-8 as well as quinones with shorter prenyl chains (ubiquinone-1 and ubiquinone-2). Here we show that the quinone derivative 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIB), a known inhibitor of the bc1 and b6f complexes found in mitochondria and chloroplasts, also inhibits quinone reduction by the Na+-NQR in a mixed inhibition mode. Tryptophan fluorescence quenching and saturation transfer difference NMR experiments in the presence of Na+-NQR inhibitor (DBMIB or 2-n-heptyl-4-hydroxyquinoline N-oxide) indicate that two quinone analog ligands are bound simultaneously by the NqrA subunit with very similar interaction constants as observed with the holoenzyme complex. We conclude that the catalytic site of quinone reduction is located on NqrA. The two ligands bind to an extended binding pocket in direct vicinity to each other as demonstrated by interligand Overhauser effects between ubiquinone-1 and DBMIB or 2-n-heptyl-4-hydroxyquinoline N-oxide, respectively. We propose that a similar spatially close arrangement of the native quinone substrates is also operational in vivo, enhancing the catalytic efficiency during the final electron transfer steps in the Na+-NQR.  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