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.
A method is described for the detection of 2-keto-4-hydroxyglutarate aldolase activity after electrophoresis of the enzyme on polyacrylamide gels. When gels are incubated with substrate (2-keto-4-hydroxyglutarate), activity is seen as a yellow-colored band due to interaction of the product )glyoxylate) with ortho-aminobenzaldehyde and glycine. Positive results have been obtained using either crude cell-free preparations or homogeneous enzyme from Escherichia coli as well as with highly purified samples of aldolase from bovine liver or kidney extracts. The method is potentially applicable to other aldolases that liberate an aliphatic aldehyde as a product; modifications and limitations of the procedure for detecting fructose 1,6-diphosphate aldolase, 2-keto-3-deoxy-6-phosphogluconate aldolase, and 2-deoxyribose-5-phosphate aldolase activities have been explored.  相似文献   

2.
Treatment of homogeneous preparations of Escherichia coli 2-keto-4-hydroxyglutarate aldolase with 1,2-cyclohexanedione, 2,3-butanedione, phenylglyoxal, or 2,4-pentanedione results in a time- and concentration-dependent loss of enzymatic activity; the kinetics of inactivation are pseudo-first order. Cyclohexanedione is the most effective modifier; a plot of log (1000/t 1/2) versus log [cyclohexanedione] gives a straight line with slope = 1.1, indicating that one molecule of modifier reacts with each active unit of enzyme. The kinetics of inactivation are first order with respect to cyclohexanedione, suggesting that the loss of activity is due to modification of 1 arginine residue/subunit. Controls establish that this inactivation is not due to modifier-induced dissociation or photoinduced structural alteration of the aldolase. The same Km but decreased Vmax values are obtained when partially inactivated enzyme is compared with native. Amino acid analyses of 95% inactivated aldolase show the loss of 1 arginine/subunit with no significant change in other amino acid residues. Considerable protection against inactivation is provided by the substrates 2-keto-4-hydroxyglutarate and pyruvate (75 and 50%, respectively) and to a lesser extent (40 and 35%, respectively) by analogs like 2-keto-4-hydroxybutyrate and 2-keto-3-deoxyarabonate. In contrast, formaldehyde or glycolaldehyde (analogs of glyoxylate) under similar conditions show no protective effect. These results indicate that an arginine residue is required for E. coli 2-keto-4-hydroxyglutarate aldolase activity; it most likely participates in the active site of the enzyme by interacting with the carboxylate anion of the pyruvate-forming moiety of 2-keto-4-hydroxyglutarate.  相似文献   

3.
The interaction of bromopyruvate with the active site of 2-keto-3-deoxygluconate-6-P aldolase ofPseudomonas saccharophila was investigated. The reagent inactivates the enzyme, exhibiting saturation kinetics and competition with pyruvate. The minimal inactivation half-time was 6 min, equivalent to a first-order rate constant of 0.115 min?1. The concentration of bromopyruvate giving the half-maximal inactivation rateK inact was 50 mM. TheK s value of pyruvate as a competitive inhibitor was 0.85 mM. The enzyme asymmetrically detritiates (3RS)-[3? 3H 2 ]bromopyruvate, forming, in water, (3S)-[3-3H,H]bromopyruvate. This stereochemistry is also exhibited by 2-keto-6-deoxygalactonate-6-P aldolase isolated from the same organism as well as the 2-keto-3-deoxygluconate-5-P aldolase ofP. putida. Over a range of [3-3H]bromopyruvate concentrations affecting the inactivation rate, the ratio of nanomoles reagent catalytically turned over per unit of enzyme inactivated remained constant at 14:1, providing evidence that both catalysis and alkylation occur at the same protein site.  相似文献   

4.
Surface glycan switching is often observed when micro-organisms transition between different biotic and abiotic niches, including biofilms, although the advantages of this switching to the organism are not well understood. Bacillus cereus grown in a biofilm-inducing medium has been shown to synthesize an unusual cell wall polysaccharide composed of the repeating subunit →6)Gal(α1–2)(2-R-hydroxyglutar-5-ylamido)Fuc2NAc4N(α1–6)GlcNAc(β1→, where galactose is linked to the hydroxyglutarate moiety of FucNAc-4-amido-(2)-hydroxyglutarate. The molecular mechanism involved in attaching 2-hydroxyglutarate to 4-amino-FucNAc has not been determined. Here, we show two genes in B. cereus ATCC 14579 encoding enzymes involved in the synthesis of UDP-FucNAc-4-amido-(2)-oxoglutarate (UDP-Yelosamine), a modified UDP-sugar not previously reported to exist. Using mass spectrometry and real time NMR spectroscopy, we show that Bc5273 encodes a C4″-aminotransferase (herein referred to as Pat) that, in the presence of pyridoxal phosphate, transfers the primary amino group of l-Glu to C-4″ of UDP-4-keto-6-deoxy-d-GlcNAc to form UDP-4-amino-FucNAc and 2-oxoglutarate. Pat also converts 4-keto-xylose, 4-keto-glucose, and 4-keto-2-acetamido-altrose to their corresponding UDP-4-amino-sugars. Bc5272 encodes a carboxylate-amine ligase (herein referred as Pyl) that, in the presence of ATP and Mg(II), adds 2-oxoglutarate to the 4-amino moiety of UDP-4-amino-FucNAc to form UDP-Yelosamine and ADP. Pyl is also able to ligate 2-oxoglutarate to other 4-amino-sugar derivatives to form UDP-Yelose, UDP-Solosamine, and UDP-Aravonose. Characterizing the metabolic pathways involved in the formation of modified nucleotide sugars provides a basis for understanding some of the mechanisms used by bacteria to modify or alter their cell surface polysaccharides in response to changing growth and environmental challenges.  相似文献   

5.
(R)-ethyl-3-hydroxyglutarate with highly optical purity (≥99%) can be used as a novel precursor for synthesis of chiral side chain of rosuvastatin. In this study, a novel synthesis route of (R)-ethyl-3-hydroxyglutarate by whole microorganism cells from racemic ethyl 4-cyano-3-hydroxybutyate was created. A strain ZJB-0910 capable of transforming racemic β-hydroxy aliphatic nitrile was isolated by employing a screening method based on a colorimetric reaction of Co2+ ion with ammonia, and identified as Rhodococcus erythropolis based on its morphology, physiological tests, Biolog, and the 16S rDNA sequence. After cultivation in a sterilized medium with composition of 20 g glucose, 5 g yeast extract, 0.5 g KH2PO4, 0.5 g K2HPO4, 0.2 g MgSO4·7H2O per liter at 30°C and 150 rpm for 48 h, the whole cells of R. erythropolis ZJB-0910 were prepared as a catalyst in (R)-enantioselective hydrolysis of racemic ethyl 4-cyano-3-hydroxybutyate for synthesis of (R)-ethyl-3-hydroxyglutarate, without bearing hydrolase activity for the ester bond of ethyl 4-cyano-3-hydroxybutyate. Under the optimized biotransformation conditions of pH 7.5, 30°C, and 20 mM substrate concentration, (R)-ethyl-3-hydroxyglutarate with 46.2% yield (ee > 99%) was afforded, and its chemical structure was determined by ESI-MS, NMR, and IR. The apparent Michaelis constant K m and maximum rate V max for this biocatalytic reaction were 0.01 M and 85.6 μmol min−1 g−1, respectively.  相似文献   

6.
The purpose of the present work was to progress in our understanding of the pathophysiology of L-2-hydroxyglutaric aciduria, due to a defect in L-2-hydroxyglutarate dehydrogenase, by creating and studying a mouse model of this disease. L-2-hydroxyglutarate dehydrogenase-deficient mice (l2hgdh -/-) accumulated L-2-hydroxyglutarate in tissues, most particularly in brain and testis, where the concentration reached ≈ 3.5 μmol/g. Male mice showed a 30% higher excretion of L-2-hydroxyglutarate compared to female mice, supporting that this dicarboxylic acid is partially made in males by lactate dehydrogenase C, a poorly specific form of this enzyme exclusively expressed in testes. Involvement of mitochondrial malate dehydrogenase in the formation of L-2-hydroxyglutarate was supported by the commensurate decrease in the formation of this dicarboxylic acid when down-regulating this enzyme in mouse l2hgdh -/- embryonic fibroblasts. The concentration of lysine and arginine was markedly increased in the brain of l2hgdh -/- adult mice. Saccharopine was depleted and glutamine was decreased by ≈ 40%. Lysine-α-ketoglutarate reductase, which converts lysine to saccharopine, was inhibited by L-2-hydroxyglutarate with a Ki of ≈ 0.8 mM. As low but significant activities of the bifunctional enzyme lysine-α-ketoglutarate reductase/saccharopine dehydrogenase were found in brain, these findings suggest that the classical lysine degradation pathway also operates in brain and is inhibited by the high concentrations of L-2-hydroxyglutarate found in l2hgdh -/- mice. Pathological analysis of the brain showed significant spongiosis. The vacuolar lesions mostly affected oligodendrocytes and myelin sheats, as in other dicarboxylic acidurias, suggesting that the pathophysiology of this model of leukodystrophy may involve irreversible pumping of a dicarboxylate in oligodendrocytes. Neurobehavioral testing indicated that the mice mostly suffered from a deficit in learning capacity. In conclusion, the findings support the concept that L-2-hydroxyglutaric aciduria is a disorder of metabolite repair. The accumulation of L-2-hydroxyglutarate exerts toxic effects through various means including enzyme inhibition and glial cell swelling.  相似文献   

7.
An H2O2-requiring enzyme system was found in the extracellular medium of ligninolytic cultures of Phanerochaete chrysosporium. The enzyme system generated ethylene from 2-keto-4-thiomethyl butyric acid (KTBA), and oxidized a variety of lignin model compounds including the diarylpropane 1-(4′-ethoxy-3′-methoxyphenyl) 1,3-dihydroxy-2-(4″-methoxyphenyl)propane (I), a β-ether dimer 1-(4′-ethoxy-3′-methoxyphenyl)glycerol-β-guaiacyl ether (IV) and an olefin 1-(4′-ethoxy-3′-methoxy-phenyl)1,2-propene (VI). The products found were equivalent to the metabolic products previously isolated from intact ligninolytic cultures. In addition, the enzyme system partially degraded 14C-ring labeled lignin. The enzyme was not found in high nitrogen (N) cultures, nor in cultures of a ligninolytic mutant strain which is incapable of metabolizing lignin.  相似文献   

8.
Hydroperoxides of polyunsaturated fatty acids can be transformed to epoxyalcohols and keto fatty acids by metal enzymes, hematin, and various catalysts. In the current study, we used hematin to transform 9-hydroperoxy-10E,12Z-octadecadienoic acid and 13-hydroperoxy-9Z,11E-octadecadienoic acid to epoxyalcohols (with trans epoxide configuration) and to keto fatty acids. The products were separated by normal phase high-performance liquid chromatography (NP-HPLC) and analyzed using postcolumn addition of isopropanol/water and online negative ion electrospray ionization mass spectrometry (MS). The tandem MS (MS/MS) spectra were studied using analogs prepared from [9,10,12,13-2H4]linoleic acid (18:2n−6) and from α-linolenic acid (18:3n−3). We also studied the MS/MS spectra of epoxyalcohols formed from 11-hydroperoxy- and 8-hydroperoxy-9Z,12Z-octadecadienoic acids. Results were confirmed by MS/MS analysis of a series of authentic standards. MS/MS ions of 9-keto-10E,12Z-octadecadienoic acid and 13-keto-9Z,11E-octadecadienoic acid could be explained by keto-enol tautomerism. MS/MS spectra of regioisomeric allylic epoxyalcohols differed in relative intensities of characteristic ions. The MS/MS spectra of the epoxyalcohols with 1-hydroxy-2,3-epoxy-4Z-pentene or 3-hydroxy-1,2-epoxy-4Z-pentene elements were virtually identical and showed two characteristic ions that differed by 30 in m/z values (CH(OH)). The results suggested that epoxide migration (Payne rearrangement) occurred during collision-induced dissociation. We conclude that regioisomeric allylic epoxyalcohols can be identified by their MS/MS spectra, whereas regioisomeric epoxyalcohols can be identified by MS/MS in combination with their retention times on NP-HPLC.  相似文献   

9.
Citroylformic acid-γ-lactone (CFA, 1-keto-2,4-dihydroxy-4-carboxyadipenoic acid(2–3)-1,4-lactone), isolated from a commercial batch of oxaloacetate, inhibited conversion of rat liver apotyrosine aminotransferase (EC 2.6.1.5) to holoenzyme. Using partially purified enzyme, the Ki was determined to be less than 0.7 mm. A more definitive Ki was difficult to obtain because at pH 7 CFA had a half-life of about 2 hr. Inhibition of the enzyme by CFA was stereospecific and reversible; the S (?) stereoisomer was approximately 10 times more inhibitory than its R(+) antipode, and over 90% of inhibited enzyme was recoverable after overnight dialysis. Preineubation of apotyrosine aminotransferase with its coenzyme (pyridoxal phosphate) prevented inhibition by CFA, and a substantial fraction of enzyme that had been inhibited by CFA could be readily reactivated by addition of high concentrations of pyridoxal phosphate. Studies with inhibitor analogs indicated that both a partially unsaturated lactone ring and a stereospecific carboxymethyl group are required for maximal inhibitory activity. The sodium salts of citroylformic acid and oxalopyruvic acid, formed by the hydrolysis of their respective lactones, were not inhibitory; 1-keto-2,4-dihydroxy-4-carboxyadipic acid-γ-lactone and little inhibitory activity, and 1-keto-2,4-dihydroxyglutarenoic acid-γ-lactone and 1-keto-2,4-dihydroxybutene-γ-lactone were somewhat better inhibitors than the R(+) stereoisomer of CFA. The possibility that CFA is a naturally occurring biological substance is discussed.  相似文献   

10.
The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases   总被引:2,自引:0,他引:2  
Mutations in isocitrate dehydrogenases (IDHs) have a gain-of-function effect leading to R(−)-2-hydroxyglutarate (R-2HG) accumulation. By using biochemical, structural and cellular assays, we show that either or both R- and S-2HG inhibit 2-oxoglutarate (2OG)-dependent oxygenases with varying potencies. Half-maximal inhibitory concentration (IC50) values for the R-form of 2HG varied from approximately 25 μM for the histone Nɛ-lysine demethylase JMJD2A to more than 5 mM for the hypoxia-inducible factor (HIF) prolyl hydroxylase. The results indicate that candidate oncogenic pathways in IDH-associated malignancy should include those that are regulated by other 2OG oxygenases than HIF hydroxylases, in particular those involving the regulation of histone methylation.  相似文献   

11.
Two pure peroxidase isoenzymes B1 and D4 were isolated from the upper parts of 10-day-old wheat seedlings by means of gel and ion-exchange chromatography. Their MWs were 85000 and 24000 respectively. B1 was unstable and under various conditions it was converted to another isoenzyme, electrophoretically identical with D4. B1 contains about 40% of neutral sugars: 17.2% arabinose, 15.3% galactose, 5% glucose and traces of mannose. D4 is free of neutral sugars. None of the isoenzymes contained amino sugars. B1 oxidizes ferulic and p-coumaric acids. This oxidation has two pH optima of 4.4 and 5.4–5.6 and is inhibited by high concentrations of substrates, cyanide and azide. B1 oxidizes IAA in the presence of phenolic cofactor and Mn2+ ions. IAA oxidation has two pH optima of 4.5 and 5.6 and is inhibited by high substrate concentration, cyanide and azide, and by a number of indole derivatives. The main products of IAA oxidation are 3-methyleneoxindole and indole-3-methanol. o- and p- diphenols induce a lag period prior to IAA oxidation. Ferulic acid is oxidized during this lag period, probably to a dimer. B1 is able to produce H2O2 from oxygen. Mn2+ ions, a phenolic cofactor and an electron donor (IAA or NADH) are needed. B1 oxidizes α-keto-γ- methylmercaptobutyric acid to ethylene. D4 has a low peroxidatic activity and is inactive as an IAA oxidase. Thus B1 is probably an active cell wall-bound peroxidase isoenzyme, whereas D4 is its decomposition product.  相似文献   

12.
Mitochondrial NADPH-dependent isocitrate dehydrogenase, IDH2, and cytosolic IDH1, catalyze reductive carboxylation of 2-oxoglutarate. Both idh2 and idh1 monoallelic mutations are harbored in grade 2/3 gliomas, secondary glioblastomas and acute myeloid leukemia. Mutant IDH1/IDH2 enzymes were reported to form an oncometabolite r-2-hydroxyglutarate (2HG), further strengthening malignancy. We quantified CO2-dependent reductive carboxylation glutaminolysis (RCG) and CO2-independent 2HG production in HTB-126 and MDA-MB-231 breast carcinoma cells by measuring 13C incorporation from 1-13C-glutamine into citrate, malate, and 2HG. For HTB-126 cells, 13C-citrate, 13C-malate, and 13C-2-hydroxyglutarate were enriched by 2-, 5-, and 15-fold at 5 mM glucose (2-, 2.5-, and 13-fold at 25 mM glucose), respectively, after 6 h. Such enrichment decreased by 6% with IDH1 silencing, but by 30–50% upon IDH2 silencing while cell respiration and ATP levels rose up to 150%. Unlike 2HG production RCG declined at decreasing CO2. At hypoxia (5% O2), IDH2-related and unrelated 13C-accumulation into citrate and malate increased 1.5–2.5-fold with unchanged IDH2 expression; whereas hypoxic 2HG formation did not. 13C–2HG originated by ∼50% from other than IDH2 or IDH1 reactions, substantiating remaining activity in IDH1&2-silenced cells. Relatively high basal 12C–2HG levels existed (5-fold higher vs. non-tumor HTB-125 cells) and 13C–2HG was formed despite the absence of any idh2 and idh1 mutations in HTB-126 cells. Since RCG is enhanced at hypoxia (frequent in solid tumors) and 2HG can be formed without idh1/2 mutations, we suggest 2HG as an analytic marker (in serum, urine, or biopsies) predicting malignancy of breast cancer in all patients.  相似文献   

13.
Extracts of Pseudomonas B4 grown with l-β-lysine (3,6-diaminohexanoate) as the main energy source are shown to contain a 3-keto-6-acetamidohexanoate cleavage enzyme that converts 3-keto-6-acetamidohexanoate and acetyl · CoA reversibly to 4-acetamidobutyryl · CoA and acetoacetate. The enzyme catalyzes the third step in β-lysine degradation. In unfractionated extracts cleavage enzyme activity is generally assayed spectrophotometrically by coupling the forward reaction with excess 4-acetamidobutyryl · CoA thiolesterase, derived from the same organism, and measuring the rate of CoASH formation by reaction with 5,5-dithiobis(2-nitrobenzoic acid). Enzyme freed of thiolesterase is conveniently assayed by using 4-acetamidobutyryl · CoA and acetoacetate as substrates and measuring acetyl · CoA formation by means of citrate synthase reaction in the presence of 5,5-dithiobis(2-nitrobenzoic acid). The cleavage enzyme has been purified 38-fold to a specific activity of 237 mU/mg. The stoichiometry, equilibrium constant, molecular weight, and various kinetic properties of the enzymatic reaction have been determined. The substrate specificity of the Pseudomonas enzyme differs markedly from that of the analogous 3-keto-5-aminohexanoate cleavage enzyme of Clostridium subterminale strain SB4 and is broader. In the forward reaction 3-ketohexanoate can replace 3-keto-6-acetamidohexanoate, and propionyl · CoA can replace acetyl · CoA as a substrate. In the backward reaction, 4-acetamidobutyryl · CoA can be replaced by any of several CoA thiolesters including the butyryl, valeryl, 4-propionamidobutyryl, 3-acetamidopropionyl, and β-alanyl derivatives, and acetoacetate can be replaced by 2-methylacetoacetate. The products of these reactions have been characterized. Unlike the cleavage enzyme of Clostridium subterminale strain SB4, the Pseudomonas enzyme is not stimulated by Co2+ or Mn2+ and is not inhibited by EDTA, 5,5-dithiobis(2-nitrobenzoic acid), or p-chloromercuribenzoate. Tracer experiments indicate that carbon atoms 1 and 2 of acetoacetate are derived from carbon atoms 1 and 2 of 3-keto-6-acetamidohexanoate, and carbon atoms 3 and 4 of acetoacetate are derived from the acetyl group of acetyl · CoA. The cleavage enzyme is not formed in detectable amounts when Pseudomonas B4 is grown in a peptone-yeast extract medium.  相似文献   

14.
Aldolases catalyze the reversible reactions of aldol condensation and cleavage and have strong potential for the synthesis of chiral compounds, widely used in pharmaceuticals. Here, we investigated a new Class II metal aldolase from the p-hydroxyphenylacetate degradation pathway in Acinetobacter baumannii, 4-hydroxy-2-keto-heptane-1,7-dioate aldolase (AbHpaI), which has various properties suitable for biocatalysis, including stereoselectivity/stereospecificity, broad aldehyde utilization, thermostability, and solvent tolerance. Notably, the use of Zn2+ by AbHpaI as a native cofactor is distinct from other enzymes in this class. AbHpaI can also use other metal ion (M2+) cofactors, except Ca2+, for catalysis. We found that Zn2+ yielded the highest enzyme complex thermostability (Tm of 87 °C) and solvent tolerance. All AbHpaI•M2+ complexes demonstrated preferential cleavage of (4R)-2-keto-3-deoxy-D-galactonate ((4R)-KDGal) over (4S)-2-keto-3-deoxy-D-gluconate ((4S)-KDGlu), with AbHpaI•Zn2+ displaying the highest R/S stereoselectivity ratio (sixfold higher than other M2+ cofactors). For the aldol condensation reaction, AbHpaI•M2+ only specifically forms (4R)-KDGal and not (4S)-KDGlu and preferentially catalyzes condensation rather than cleavage by ∼40-fold. Based on 11 X-ray structures of AbHpaI complexed with M2+ and ligands at 1.85 to 2.0 Å resolution, the data clearly indicate that the M2+ cofactors form an octahedral geometry with Glu151 and Asp177, pyruvate, and water molecules. Moreover, Arg72 in the Zn2+-bound form governs the stereoselectivity/stereospecificity of AbHpaI. X-ray structures also show that Ca2+ binds at the trimer interface via interaction with Asp51. Hence, we conclude that AbHpaI•Zn2+ is distinctive from its homologues in substrate stereospecificity, preference for aldol formation over cleavage, and protein robustness, and is attractive for biocatalytic applications.  相似文献   

15.
Fructose-1,6-bisphosphatase (FBPase) and sedoheptulose-1,7-bisphosphatase (SBPase) were identified and purified from the Crassulacean acid metabolism (CAM) plant, Kalanchoë daigremontiana. FBPase and SBPase showed respective molecular weights of 180,000 and 76,000, and exhibited immunological cross-reactivity with their counterparts from chloroplasts of C3 (spinach) and C4 (corn) plants. Based on Western blot analysis, FBPase was composed of four identical 45,000-dalton subunits and SBPase of two identical 38,000-dalton subunits. Immunological evidence, together with physical properties, indicated that both enzymes were of chloroplast origin.

Kalanchoë FBPase and SBPase could be activated by thioredoxin f reduced chemically by dithiothreitol or photochemically by a reconstituted Kalanchoë ferredoxin/thioredoxin system. Both enzymes were activated synergistically by reduced thioredoxin f and thier respective substrates.

Kalanchoë FBPase could be partially activated by Mg2+ at concentrations greater than 10 millimolar; however, such activation was considerably less than that observed in the presence of reduced thioredoxin and Ca2+, especially in the pH range between 7.8 and 8.3. In contrast to FBPase, Kalanchoë SBPase exhibited an absolute requirement for a dithiol such as reduced thioredoxin irrespective of Mg2+ concentration. However, like FBPase, increased Mg2+ concentrations enhanced the thioredoxin-linked activation of this enzyme.

In conjunction with these studies, an NADP-linked malate dehydrogenase (NADP-MDH) was identified in cell-free preparations of Kalanchoë leaves which required reduced thioredoxin m for activity.

These results indicate that Kalanchoë FBPase, SBPase, and NADP-MDH share physical and regulatory properties with their equivalents in C3 and C4 plants. In contrast to previous evidence, all three enzymes appear to have the capacity to be photoregulated in chloroplasts of CAM plants, thereby providing a means for the functional segregation of glucan synthesis and degradation.

  相似文献   

16.
A number of bacteria belonging to the genera Proteus, Providencia, Pseudomonas and Erwinia have been tested for their capacity to oxidize l-amino acids to their corresponding α-keto acids. Members of the Proteus and the Providencia genera were active towards various l-amino acids. Immobilized cell preparations of Providencia sp. PCM 1298 were shown to form up to 80 mg α-keto-γ-methiol butyric acid from l-methionine per g of gel preparation (containing 4% w/w cells) per day. The productivity was highly dependent on the size of the beads. Oxygen appeared to be the rate-limiting substrate and oxygen transfer rates of 3–4 μmol cm?2h?1were calculated. The entrapment of activated charcoal to remove H2O2 formed during the oxidation extended the half-life of the immobilized biocatalyst considerably. A decrease in l-amino acid oxidase [l-amino acid: oxygen oxidoreductase (deaminating); EC 1.4.3.2] activity during operation could be compensated for by reinoculation of the alginate-entrapped cells in fresh growth medium, allowing use of these preparations of immobilized bacterial cells for more than one month.  相似文献   

17.
In the course of screening for antioxidative carotenoids from bacteria, we isolated and identified a novel carotenoid, OH-chlorobactene glucoside hexadecanoate (4), and rare carotenoids, OH-chlorobactene glucoside (1), OH-γ-carotene glucoside (2) and OH-4-keto-γ-carotene glucoside hexadecanoate (3) from Rhodococcus sp. CIP. The singlet oxygen (1O2) quenching model of these carotenoids showed potent antioxidative activities IC50 14.6 μM for OH-chlorobactene glucoside hexadecanoate (4), 6.5 μM for OH-chlorobactene glucoside (1), 9.9 μM for OH-γ-carotene glucoside (2) and 7.3 μM for OH-4-keto-γ-carotene glucoside hexadecanoate (3).  相似文献   

18.
Treatment of pure 2-keto-4-hydroxyglutarate aldolase from Escherichia coli, a "lysine-type," Schiff-base mechanism enzyme, with the substrate analog bromopyruvate results in a time- and concentration-dependent loss of enzymatic activity. Whereas the substrates pyruvate and 2-keto-4-hydroxyglutarate provide greater than 90% protection against inactivation by bromopyruvate, no protective effect is seen with glycolaldehyde, an analog of glyoxylate. Inactivation studies with [14C] bromopyruvate show the incorporation of 1.1 mol of 14C-labeled compound/enzyme subunit; isolation of a radioactive peptide and determination of its amino acid sequence indicate that the radioactivity is associated with glutamate 45. Incubation of the enzyme with excess [14C]bromopyruvate followed by denaturation with guanidine.HCl allow for the incorporation of carbon-14 at cysteines 159 and 180 as well. Whereas the presence of pyruvate protects Glu-45 from being esterified, it does not prevent the alkylation of these 2 cysteine residues. The results indicate that Glu-45 of E. coli 2-keto-4-hydroxyglutarate aldolase is essential for catalytic activity, most likely acting as the amphoteric proton donor/acceptor that is required as a participant in the overall mechanism of the reaction catalyzed.  相似文献   

19.
Isopycnic sucrose gradient separation of rat liver organelles revealed the presence of two distinct branched-chain α-keto acid decarboxylase activities; a mitochondrial activity, which decarboxylates the three branched-chain α-keto acids and requires CoA and NAD+ and a cytosolic activity, which decarboxylates α-ketoisocaproate, but not α-ketoisovalerate, or α-keto-β-methylvalerate. The latter enzyme does not require added CoA or NAD+. Assay conditions for the cytosolic α-ketoisocaproate decarboxylase activity were optimized and this activity was partially characterized. In rat liver cytosol preparations this activity has a pH optimum of 6.5 and is activated by 1.5 m ammonium sulfate. The decarboxylase activity has an apparent Km of 0.03 mm for α-ketoisocaproate when optimized assay conditions are employed. Phenylpyruvate is a very potent inhibitor. α-Ketoisovalerate, α-keto-β-methylvalerate, α-ketobutyrate, and α-ketononanoate also inhibit the α-ketoisocaproate decarboxylase activity. The data indicate that the soluble α-ketoisocaproate decarboxylase is an oxidase. Rat liver cytosol preparations consumed oxygen when either α-ketoisocaproate or α-keto-γ-methiolbutyrate were added. None of the other α-keto acids tested stimulated oxygen consumption. 1-14C-Labeled α-keto-γ-methiolbutyrate is also decarboxylated by cytosol preparations. The α-ketoisocaproate oxidase was purified 20-fold from a 70,000g supernatant fraction of a rat liver homogenate. In these preparations the activity was increased 4-fold by the addition of dithiothreitol, ferrous iron, and ascorbate. The major product of this enzyme activity is β-hydroxyisovalerate. Isovalerate is not a free intermediate in the reaction. The data indicate an alternative pathway for metabolism of α-ketoisocaproate which produces β-hydroxyisovalerate.  相似文献   

20.
Ouabain-treated turtle bladders bathed on both surfaces by identical HCO3?/CO2-containing, Cl?-free Na+ media exhibit a short-circuit current (Isc) and transepithelial potential (p.d.) serosa electronegative to mucosa. Addition of 3-isobutyl-1-methylxanthine (IBMX), an inhibitor of cyclic nucleotide phosphodiesterase, rapidly reverses the direction of the Isc and p.d.. The IBMX-induced reversal of Isc and p.d. is (1) dependent on the presence of HCO3? (and CO2) in the serosal bathing fluid, (2) independent of Na+ and other ions in the bathing medium, (3) decreased by inhibitors of carbonic anhydrase or oxidative metabolism, (4) increased by the serosal addition of cyclic AMP or the disulfonic stilbene, SITS. The results constitute evidence that the reversed Isc elicited by IBMX represents electrogenic secretion of HCO3?.  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