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 78 毫秒
1.
The kinetics of NADH oxidation by the outer membrane electron transport system of intact beetroot (Beta vulgaris L.) mitochondria were investigated. Very different values for Vmax and the Km for NADH were obtained when either antimycin A-insensitive NADH-cytochrome c activity (Vmax= 31 ± 2.5 nmol cytochrome c (mg protein)?1 min?1; Km= 3.1 ± 0.8 μM) or antimycin A-insensitive NADH-ferricyanide activity (Vmax= 1.7 ± 0.7 μmol ferricyanide (mg protein)?1 min?1; Km= 83 ± 20 μM) were measured. As ferricyanide is believed to accept electrons closer to the NADH binding site than cytochrome c, it was concluded that 83 ± 20 μM NADH represented a more accurate estimate of the binding affinity of the outer membrane dehydrogenase for NADH. The low Km determined with NADH-cytochrome c activity may be due to a limitation in electron flow through the components of the outer membrane electron transport chain. The Km for NADH of the externally-facing inner membrane NADH dehydrogenase of pea leaf (Pisum sativum L. cv. Massey Gem) mitochondria was 26.7 ± 4.3 μM when oxygen was the electron acceptor. At an NADH concentration at which the inner membrane dehydrogenase should predominate, the Ca2+ chelator, ethyleneglycol-(β-aminoethylether)-N,N,-tetraacetic acid (EGTA), inhibited the oxidation of NADH through to oxygen and to the ubiquinone-10 analogues, duroquinone and ubiquinone-1, but had no effect on the antimycin A-insensitive ferricyanide reduction. It is concluded that the site of action of Ca2+ involves the interaction of the enzyme with ubiquinone and not with NADH.  相似文献   

2.
Inside-out submitochondrial particles from both potato (Solanum tuberosum L. cv. Bintje) tubers and pea (Pisum sativum L. cv. Oregon) leaves possess three distinct dehydrogenase activities: Complex I catalyzes the rotenone-sensitive oxidation of deamino-NADH, NDin(NADPH) catalyzes the rotenone-insensitive and Ca2+-dependent oxidation of NADPH and NDin(NADH) catalyzes the rotenone-insensitive and Ca2+-independent oxidation of NADH. Diphenylene iodonium (DPI) inhibits complex I, NDin(NADPH) and NDin (NADH) activity with a Ki of 3.7, 0.17 and 63 µM, respectively, and the 400-fold difference in Ki between the two NDin made possible the use of DPI inhibition to estimate NDin (NADPH) contribution to malate oxidation by intact mitochondria. The oxidation of malate in the presence of rotenone by intact mitochondria from both species was inhibited by 5 µM DPI. The maximum decrease in rate was 10–20 nmol O2 mg?1 min?1. The reduction level of NAD(P) was manipulated by measuring malate oxidation in state 3 at pH 7.2 and 6.8 and in the presence and absence of an oxaloacetate-removing system. The inhibition by DPI was largest under conditions of high NAD(P) reduction. Control experiments showed that 125 µM DPI had no effect on the activities of malate dehydrogenase (with NADH or NADPH) or malic enzyme (with NAD+ or NADP+) in a matrix extract from either species. Malate dehydrogenase was unable to use NADP+ in the forward reaction. DPI at 125 µM did not have any effect on succinate oxidation by intact mitochondria of either species. We conclude that the inhibition caused by DPI in the presence of rotenone in plant mitochondria oxidizing malate is due to inhibition of NDin(NADPH) oxidizing NADPH. Thus, NADP turnover contributes to malate oxidation by plant mitochondria.  相似文献   

3.
Exogenous NAD+ stimulated the rotenone-resistant oxidation of all the NAD+-linked tricarboxylic acid-cycle substrates in mitochondria from Jerusalem artichoke (Helianthus tuberosus L.) tubers. The stimulation was not removed by the addition of EGTA, which is known to inhibit the oxidation of exogenous NADH. It is therefore concluded that added NAD+ gains access to the matrix space and stimulates oxidation by the rotenone-resistant NADH dehydrogenase located on the matrix surface of the inner membrane. Added NAD+ stimulated the activity of malic enzyme and displaced the equilibrium of malate dehydrogenase; both observations are consistent with entry of NAD+ into the matrix space. Analysis of products of malate oxidation showed that rotenone-resistant oxygen uptake only occurred when the concentration of oxaloacetate was low and that of NADH was high. Thus it is proposed that the concentration of NADH regulates the activity of the two internal NADH dehydrogenases. Evidence is presented to suggest that the rotenone-resistant NADH dehydrogenase is engaged under conditions of high phosphorylation potential, which restricts electron flux through the rotenone-sensitive dehydrogenase (coupled to ATP synthesis).  相似文献   

4.
Alpha-chymotrypsin was made more hydrophilic by modifying 11 (out of 16) ε-amino groups with pyromellitic dianhydride. The hydrophilic preparation was precipitated with n-propanol. This preparation gave significantly higher initial rates at the optimum aw (127.51 nmol mg?1 min?1 in n-octane and 21.30 nmol mg?1 min?1 in acetonitrile at aw=0.33) compared with the lyophilized preparation (53.50 nmol mg?1 min?1 in n-octane and 0.26 nmol mg?1 min?1 in acetonitrile at aw=0.97). FT-IR showed that the precipitate of modified alpha-chymotrypsin has a higher content of alpha-helices and beta-sheets compared to the lyophilized powder.  相似文献   

5.
The proton magnetic resonance spectra of the dihydronicotinamide ring of αNADH3 and the nicotinamide ring of αNAD+ are reported and the proton absorptions assigned. The absolute assignment of the C4 methylene protons of αNADH is based on the generation of specifically deuterium-labeled (pro-S) B-deuterio-αNADH from enzymatically prepared B-deuterio-βNADH. The C4 proton absorption of αNAD+ is assigned by oxidation of B-deuterio-αNADH by the A specific, yeast alcohol dehydrogenase to yield 4-deuterio-αNAD+.The epimerization of either αNADH or βNADH yields an equilibrium ratio of approximately 9:1 βNADH to αNADH. The rate of epimerization of αNADH to βNADH at 38 °C in 0.05, pH 7.5, phosphate buffer is 3.1 × 10?3 min?1, corresponding to a half-life of 4 hr. Four related dehydrogenases, yeast and horse liver alcohol dehydrogenase and chicken M4 and H4 lactate dehydrogenase, are shown to oxidize αNADH to αNAD+ at rates three to four orders of magnitude slower than for βNADH. By using specifically labeled B-deuterio-αNADH the enzymatic oxidation by yeast alcohol dehydrogenase has been shown to occur with the identical stereospecificity as the oxidation of βNADH. The nonenzymatic epimerization of αNADH to βNADH and the enzymatic oxidation αNADH are discussed as a possible source of αNAD+in vivo.  相似文献   

6.
Nitrite-driven anaerobic ATP synthesis in barley and rice root mitochondria   总被引:4,自引:0,他引:4  
Mitochondria isolated from the roots of barley (Hordeum vulgare L.) and rice (Oryza sativa L.) seedlings were capable of oxidizing external NADH and NADPH anaerobically in the presence of nitrite. The reaction was linked to ATP synthesis and nitric oxide (NO) was a measurable product. The rates of NADH and NADPH oxidation were in the range of 12–16 nmol min−1 mg−1 protein for both species. The anaerobic ATP synthesis rate was 7–9 nmol min−1 mg−1 protein for barley and 15–17 nmol min−1 mg−1 protein for rice. The rates are of the same order of magnitude as glycolytic ATP production during anoxia and about 3–5% of the aerobic mitochondrial ATP synthesis rate. NADH/NADPH oxidation and ATP synthesis were sensitive to the mitochondrial inhibitors myxothiazol, oligomycin, diphenyleneiodonium and insensitive to rotenone and antimycin A. The uncoupler FCCP completely eliminated ATP production. Succinate was also capable of driving ATP synthesis. We conclude that plant mitochondria, under anaerobic conditions, have a capacity to use nitrite as an electron acceptor to oxidize cytosolic NADH/NADPH and generate ATP.  相似文献   

7.
Mitochondria were isolated from 7-day-old wheat roots (Triticum vulgare Vill. cv. Svenno Spring Wheat) grown in either a full-strength culture medium (100%) or in the same medium diluted 100 times (1%). Outer membrane integrity was assayed using the cytochrome c reduction assay. This indicated about 20% damage. Using an oxygen electrode the respiration of the mitochondria was measured with either malate or succinate as the substrate (both 40 mM). KCN (3 mM) and salicylhydroxamic acid (SHAM, 1 mM) were used as inhibitors. The properties of the isolated mitochondria (STATE 3 rate, ADP/O ratio, and KCN-sensitivity) depend upon the ionic concentration of the growth medium of the roots. In the mitochondria isolated from roots grown in the 1% medium (1% mitochondria) there is a synergistic effect of KCN and SHAM. This means that electrons can be shifted from one pathway to the other when only one of the inhibitors is added. This flexibility between the electron pathways is almost nil in the mitochondria isolated from roots grown in the 100% medium (100% mitochondria). The maximal capacity of the alternative electron pathway (= rate in the presence of KCN) is higher in 1% (40 nmol O2 min?1 (mg protein)?1) than in 100% mitochondria (20 nmol O2 min?1 (mg protein)?1. In 100% mitochondria the alternative pathway seems to be operating at maximal capacity in the absence of KCN with both substrates and in both STATES 3 and 4. In 1% mitochondria the alternative pathway functions at >50% of its capacity in the absence of KCN.  相似文献   

8.
We analyzed the effect of lysophosphatidylcholine (lysoPC) on the activity of the plasma membrane (PM) H+-AT-Pase measured at pH 6.3 or 7.5 in inside-out PM vesicles isolated from germinating radish seeds. LysoPC stimulated PM H+-ATPase at both pHs, but the dependence of the effect on lysoPC concentration was different: at pH 6.3 maximal stimulation was observed with 40 to 200 μg ml?1 lysoPC, while at pH 7.5 a sharp peak of activation was observed at about 50 μg ml?1 lysoPC, higher concentrations becoming dramatically inhibitory; this inhibitory effect was considerably reduced in the presence of 10% (v/v) glycerol. In trypsin-treared PM lysoPC stimulated the H+-ATPase activity assayed at pH 6.3, but only marginally that assayed at pH 7.5. LysoPC increased both Vmax (from 190 to 280nmol min?1 mg?1 prot) and apparent KM (from 0.15 to 0.3 mM) of the H+-ATPase at pH 6.3, while it increased Vmax (from 120 to 230 nmol min?1 mg?1 prot) and decreased apparent Km (from 0.8 to 0.4 mM) at pH 7.5. Low concentrations of Nacetylimidazole (10 to 50 mM), which modifies tyrosine residues, abolished the stimulation by lysoPC of the PM H+-ATPase activity at pH 7.5, but not that observed at pH 6.3. These results indicate that lysoPC influences the PM H+-ATPase through different mechanisms, and that its effect can only partly be ascribed to its ability to hamper the inhibitory interaction of the regulatory C-terminal domain with the catalytic site. N-acety-limidazole did not affect the stimulation of PM H+-ATPase by controlled trypsin treatment or by fusicoccin, indicating that the requirement for the tyrosine residue(s) modified by low Nacetylimidazole concentrations is specific for lysoPC-induced displacement of the C-terminal domain.  相似文献   

9.
Crude extracts of Clostridium thermoaceticum DSM 521 contain various AMAPORs (artificial mediator accepting pyridine nucleotide oxidoreductases). The specific activities of this mixture of AMAPORs is about 8–9 U mg?1 protein (µmoles mg?1 min?1) for NADPH and 3–4 U mg?1 protein for NADH formation with reduced methylviologen (MV++) as electron donor. These AMAPOR-activities are only slightly oxygen sensitive. The reoxidation of NADPH and NADH with carboxamido-methylviologen is catalysed by crude extracts with 2.0 and 1.6 U mg?1 protein, respectively. The same crude extracts also catalyse the dehydrogenation of reduced pyridine nucleotides with suitable quinones such as anthraquinone-2,6-disulphonate. The reduced quinone can be reoxidised by dioxygen.

The Km-values of these enzymes for the pyridine nucleotides and also for the artificial electron mediators are in a suitable range for preparative transformations.

Furthermore the crude extract of C. thermoaceticum contains about 2.5 U mg?1 protein of an NADP+-dependent formate dehydrogenase (FDH), which is suitable for NADPH and/or MV++ regeneration. The regeneration of MV++ with FDH and formate as electron donor proceeds with a specific activity of about 5 U mg?1 protein of the crude extract. The reduced viologen in turn reduces NAD(P)+ by AMAPOR. The formate dehydrogenase is sensitive to oxygen.

Examples of compounds which have been prepared by combination of AMAPORs or formate dehydrogenase with an oxidoreductase are: (S)-3-hydroxycarboxylates, esters of (S)-3-hydroxycarboxylates, (1R,2S)-1-hydroxypropane-tricarboxylate (Ds-(+)-isocitrate), Ls-(-)-isocitrate and 6-phosphogluconate.  相似文献   

10.
11.
In this article we compare the kinetic behavior toward pyridine nucleotides (NAD+, NADH) of NAD+-malic enzyme, pyruvate dehydrogenase, isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and glycine decarboxylase extracted from pea (Pisum sativum) leaf and potato (Solanum tuberosum) tuber mitochondria. NADH competitively inhibited all the studied dehydrogenases when NAD+ was the varied substrate. However, the NAD+-linked malic enzyme exhibited the weakest affinity for NAD+ and the lowest sensitivity for NADH. It is suggested that NAD+-linked malic enzyme, when fully activated, is able to raise the matricial NADH level up to the required concentration to fully engage the rotenone-resistant internal NADH-dehydrogenase, whose affinity for NADH is weaker than complex I.  相似文献   

12.
Abstract: The oxidation of 4-aminobutyric acid (GABA) by nonsynaptosomal mitochondria isolated from rat forebrain and the inhibition of this metabolism by the branched-chain fatty acids 2-methyl-2-ethyl caproate (MEC) and 2, 2-dimethyl valerate (DMV) were studied. The rate of GABA oxidation, as measured by O2 uptake, was determined in medium containing either 5 or 100 mM-[K+]. The apparent Km for GABA was 1.16 ± 0.19 mM and the Vmax in state 3 was 23.8 ± 5.5 ng-atoms O2. min?1. mg protein?1 in 5 mM-[K+]. In a medium with 100 mM-[K+] the apparent Km was 1.11 ± 0.17 mM and Vmax was 47.4 ± 5.7 ng-atoms O2. min?1. mg protein?1. The Km for MEC was determined to be 0.58 ± 0.24 or 0.32 ± 0.08 mM, in 5 or 100 mM-[K+], respectively. For DMV, the Ki was 0.28 ± 0.05 or 0.34 ± 0.06 mM, in 5 or 100 mM-[K+] medium, respectively. The O2 uptake of the mitochondria in the presence of GABA was coupled to the formation of glutamate and aspartate; the ratio of oxygen uptake to the rate of amino acid formation was close to the theoretical value of 3. Neither the [K2] nor any of the above inhibitors had any effect on this ratio. The metabolism of exogenous succinic semialdehyde (SSA) by these same mitochondria was also examined. The Vmax for utilization of oxygen in the presence of SSA was much greater than that found with exogenously added GABA, indicating that the capacity for GABA oxidation by these mitochondria is not limited by SSA dehydrogenase. In addition, the branched-chain fatty acids did not inhibit the metabolism of exogenously added SSA. Thus, the inhibitors examined apparently act by competitively inhibiting the GABA transaminase system of the mitochondria.  相似文献   

13.
Cell-free extracts of crotonate-grown cells of the syntrophic butyrate-oxidizing bacteriumSyntrophospora bryantii contained high hydrogenase activities (8.5–75.8 µmol · min–1 mg–1 protein) and relatively low formate dehydrogenase activities (0.04–0.07 µmol · min–1 mg–1 protein). The K M value and threshold value of the hydrogenase for H2 were 0.21 mM and 18 µM, respectively, whereas the K M value and threshold value of the formate dehydrogenase for formate were 0.22 mM and 10 µM, respectively. Hydrogenase, butyryl-CoA dehydrogenase and 3-OH-butyryl-CoA dehydrogenase were detected in the cytoplasmic fraction. Formate dehydrogenase and CO2 reductase were membrane-bound, likely located at the outer aspect of the cytoplasmic membrane. Results suggest that during syntrophic butyrate oxidation H2 is formed intracellularly while formate is formed at the outside of the cell.  相似文献   

14.
《Biomarkers》2013,18(4):267-272
Abstract

Sulphonamide hypersensitivity reactions are believed to be mediated through reactive intermediates derived from oxidation of the paraamino group to form sulphonamide hydroxylamines. Sulphamethoxazole hydroxylamine (SMX-HA) can be acetylated by N-acetyltransferase (NAT) enzymes to form an acetoxy metabolite (acetoxySMX). In the current studies, acetoxySMX was found to be not toxic over the concentration range of 0 to 500 μM towards a human lymphoblastoid cell line (RPMI 1788) or a human hepatoma cell line (HepG2). Further, transient expression of NAT1 in COS-1 cells or stable transfection of NAT1 andNAT2 in HepG2 cells did not alter the toxicity of SMX-HA in vitro. The activity of NAT1 in isolated mononuclear leucocytes (a reflection of systemic NAT1 activity) determined with paraaminobenzoic acid as a substrate was not different between controls (n = 11) or patients with a known hypersensitivity reaction (n = 5) (4.1 ±1.2 nmol min?1mg?1 vs 5.7 ± 1.4 nmol min?1 mg?1). Thus, acetoxy SMX is unlikely to play a significant role in mediating SMX hypersensitivity reactions anda constitutive deficiency in NAT1 activity is not a common finding in patients susceptible to SMX hypersensitivity reactions.  相似文献   

15.
The inhibition of neuraminidase from Clostridium chauvoei (jakari strain) with partially purified methanolic extracts of some plants used in Ethnopharmacological practice was evaluated. Extracts of two medicinal plants, Tamarindus indicus and Combretum fragrans at 100–1000 μg/ml, both significantly reduced the activity of the enzyme in a dose-dependent fashion (P < 0.001).

The estimated IC50 values for Tamarindus indicus and Combretum fragrans were 100 and 150 μ/ml respectively. Initial velocity studies conducted, using fetuin as substrate revealed a non-competitive inhibition with the Vmax significantly altered from 500 μmole min?1 mg?1 to 240μmole min?1 mg?1 and 340 μmole min?1 mg?1 in the presence of Tamarindus indicus and Combretum fragrans respectively. The KM remained unchanged at 0.42 mM. The computed Index of physiological efficiency was reduced from 1.19 min?1 to 0.57 min?1 and 0.75 min?1 with Tamarindus indicus and Combretum fragrans as inhibitors respectively.  相似文献   

16.
Submitochondrial particles (SMP) were isolated from potato ( Solanum tuberosum L. cv. Bintje) tubers. The SMP were 91% inside-out and they were able to form a membrane potential, as monitored by oxonol VI, with succinate, NADH and NADPH. The pH dependence and kinetics of NADH and NADPH oxidation by these SMP was studied using three different electron acceptors – O2, duroquinone and ferricyanide. In addition, the SMP were solubilized, fractionated by non-denaturing polyacrylamide gel electrophoresis, and the gels were stained for NAD(P)H dehydrogenase activity and specificity at different pH using Nitro Blue Tetrazolium. From the results we conclude that there are at least two distinct NAD(P)H dehydrogenases on the inner surface of the inner membrane: (1) Complex 1 which oxidizes NADH and deamino-NADH in a rotenone-sensitive manner, (O2 as acceptor) with optimum activity at pH 8 and a very low Km(NADH) of 3 μ M . It also oxidizes NADPH and deamino-NADPH in a rotenone-sensitive manner, but with a pH optimum at pH 5.8 and a very high Km(NADPH) of more than 1 m M . This complex is found as a broad, diffuse band at the top of the gels. (2) A second dehydrogenase which oxidizes NADH in a rotenone-insensitive manner with optimum activity at pH 6.2 and a higher Km(NADH) of 14 μ M . It also oxidizes NADPH in a rotenone-insensitive manner with an activity optimum at pH 6.8 and low Km(NADPH) of 25 μ M . This dehydrogenase does not oxidize deamino-NAD(P)H. One of the sharp bands around the middle of the native gels may be caused by this dehydrogenase indicating that it has a relatively low molecular mass compared to Complex I. Several other NAD(P)H dehydrogenase bands were observed on the gels which we cannot yet assign.  相似文献   

17.
A particulate adenylate cyclase was identified in the excitable ciliary membrane from Paramecium tetraurelia. MnATP was preferentially used as substrate, the Km was 67 μM, Vmax was 1 nmol cAMP.min?1.mg?1, a marked temperature optimum of 37°C was observed. Adenylate cyclase was not inhibited by 100 μM EGTA or 100 μM La3+, whereas under these conditions guanylate cyclase activity was abolished. Fractionation of ciliary membrane vesicles by a Percoll density gradient yielded two vesicle populations with adenylate cyclase activity. In contrast, calmodulin/Ca-dependent guanylate cyclase was associated with vesicles of high buoyant density only.  相似文献   

18.
Plant (and fungal) mitochondria contain multiple NAD(P)H dehydrogenases in the inner membrane all of which are connected to the respiratory chain via ubiquinone. On the outer surface, facing the intermembrane space and the cytoplasm, NADH and NADPH are oxidized by what is probably a single low-molecular-weight, nonproton-pumping, unspecific rotenone-insensitive NAD(P)H dehydrogenase. Exogenous NADH oxidation is completely dependent on the presence of free Ca2+ with aK 0.5 of about 1 µM. On the inner surface facing the matrix there are two dehydrogenases: (1) the proton-pumping rotenone-sensitive multisubunit Complex I with properties similar to those of Complex I in mammalian and fungal mitochondria. (2) a rotenone-insensitive NAD(P)H dehydrogenase with equal activity with NADH and NADPH and no proton-pumping activity. The NADPH-oxidizing activity of this enzyme is completely dependent on Ca2+ with aK 0.5 of 3 µM. The enzyme consists of a single subunit of 26 kDa and has a native size of 76 kDa, which means that it may form a trimer.  相似文献   

19.
The ferricyande assay for Type I NADH dehydrogenase (high molecular weight soluble form) was evaluated. A turnover number of 4.2 × 105 min?1, based on Vmax(ferricyanide) and FMN content, and Km(ferricyanide) of 2.2 mM were determined for this enzyme. Inclusion of a NAD-recycling system consisting of alcohol dehydrogenase and ethanol is suggested for determination of Km(NADH). This Km was found to be 17 μ M in contrast to earlier reported values of around 100 μ M.  相似文献   

20.
Desulfotomaculum acetoxidans oxidizes acetate to CO2 with sulfate. This organism metabolizes acetate via a pathway in which C1 units rather than tri- and dicarboxylic acids are intermediates. We report here that cell extracts of D. acetoxidans catalyzed an exchange between CO2 and the carboxyl group of acetate at a rate of 90 nmol · min-1 · mg-1 protein which is sufficient to account for the in vivo acetate oxidation rate of 250 nmol · min-1 · mg-1 protein. The reaction was strictly dependent on both ATP and coenzyme A. The extracts contain high activities of acetate kinase (6.3 U · mg-1 protein) and phosphotransacetylase (60 U · mg-1 protein). These findings indicate that acetyl-CoA rather than acetyl-phosphate or acetate is the substrate of the carbon-carbon cleavage activity. Exchange was only observed in the presence of strong reducing agents such as Ti3+. Interestingly, the cell extracts also catalyzed the reduction of CO2 to CO with Ti3+ as electron donor (120 nmol · min-1 · mg-1 protein). Carbon monoxide dehydrogenase and other oxidoreductases involved in acetate oxidation were found to be partially associated with the membrane fraction suggesting a membrane localization of these enzymes.Abbreviations MOPS Morpholinopropane sulfonic acid - Tricine N-tris(hydroxymethyl)-methylglycine - DTT d,l-1,4-Dithiothreitol - DMN 2,3-Dimethyl-1,4-naphthoquinone - MVOX Methyl viologen, oxidized - APS Adenosinephosphosulfate - SRB Sulfate reducing bacteria - U mol product formed per min  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