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.
J González  I Vidal  A Bernardo  R Martin 《Biochimie》1988,70(12):1791-1797
The kinetic mechanism of diacetyl and 2,3-pentanedione reduction by diacetyl reductase from Staphylococcus aureus was investigated. The shape of the primary double reciprocal plots, the product inhibition pattern, and the features of the inhibition by a substrate analogue (acetone) show that diacetyl is reduced via an Ordered Bi-Bi mechanism, and 2,3-pentanedione by an Ordered Bi-Bi or Theorell-Chance mechanism. NADH is the leading substrate in both reactions. Affinity constants for the coenzyme and the substrates and inhibition constants for NAD, acetoin, and acetone were also calculated. This enzyme has a high affinity for NADH; Km (31-50 microM) and Ks (20-27 microM) for this compound are around one-tenth of the NADH intracellular concentration. Therefore, it must operate in vivo saturated with the coenzyme. This condition is not adequate to play the role, formerly proposed for diacetyl reductases, of regulating the equilibrium between oxidized and reduced forms of pyridine-nucleotides.  相似文献   

2.
The kinetic mechanism of diacetyl and 2,3-pentanedione reduction by diacetyl reductase from Staphylococcus aureus was investigated. The shape of the primary double reciprocal plots, the product inhibition pattern, and the features of the inhibition by a substrate analogue (acetone) show that diacetyl is reduced via an Ordered Bi-Bi mechanism, and 2,3-pentanedione by an Ordered Bi-Bi or Theorell-Chance mechanism. NADH is the leading substrate in both reactions.Affinity constants for the coenzyme and the substrates and inhibition constants for NAD, acetoin, and acetone were also calculated. This enzyme has a high affinity for NADH; Km (31–50 μM) and Ks (20–27 μM) for this compound are around one-tenth of the NADH intracellular concentration. Therefore, it must operate in vivo saturated with the coenzyme. This condition is not adequate to play the role, formerly proposed for diacetyl reductases, of regulating the equilibrium between oxidized and reduced forms of pyridine-nucleotides.  相似文献   

3.
A new method, faster, milder and more efficient than the one previously described [Bryn, K., Hetland, O. & Stormer, F. C. (1971) Eur. J. Biochem, 18, 116-119], for purification of diacetyl(acetoin) reductase from Enterobacter aerogenes is proposed. The experiments carried out with the electrophoretically pure preparations obtained by this procedure show that the enzyme (a) produces L-glycols from the corresponding L-alpha-hydroxycarbonyls by reversible reduction of their oxo groups and also reduces the oxo group of uncharged alpha-dicarbonyls converting them into L-alpha-hydroxycarbonyls, and (b) is specific for NAD. This is a new enzyme for which we suggest the systematic name of L-glycol: NAD+ oxidoreductase and the recommended name of L-glycol dehydrogenase(NAD). The molecular mass, pI, affinity for substrates and pH profiles of this enzyme are also described.  相似文献   

4.
Two 2,3-butanediol dehydrogenases (enzymes 1 and 2; molecular weight of each, 170,000) have been partially purified from Lactococcus lactis subsp. lactis (Streptococcus diacetylactis) D10 and shown to have reductase activity with either diacetyl or acetoin as the substrate. However, the reductase activity with 10 mM diacetyl was far greater for both enzymes (7.0- and 4.7-fold for enzymes 1 and 2, respectively) than with 10 mM acetoin as the substrate. In contrast, when acetoin and diacetyl were present together, acetoin was the preferred substrate for both enzymes, with enzyme 1 showing the more marked preference for acetoin. meso-2,3-Butanediol was the only isomeric product, with enzyme 1 independent of the substrate combinations. For enzyme 2, both the meso and optical isomers of 2,3-butanediol were formed with acetoin as the substrate, but only the optical isomers were produced with diacetyl as the substrate. With batch cultures of strain D10 at or near the point of citrate exhaustion, the main isomers of 2,3-butanediol present were the optical forms. If the pH was sufficiently high (>pH 5), acetoin reduction occurred over time and was followed by diacetyl reduction, and meso-2,3-butanediol became the predominant isomer. Interconversion of the optical isomers into the meso isomer did occur. The properties of 2,3-butanediol dehydrogenases are consistent with diacetyl and acetoin removal and the appearance of the isomers of 2,3-butanediol.  相似文献   

5.
NADP+-dependent dihydrodiol dehydrogenase (trans-1,2-dihydrobenzene-1,2-diol: NADP+ oxidoreductase, EC 1.3.1.20) activity in the cytosol of guinea-pig testis was separated into two major and two minor peaks by Q-Sepharose chromatography; one minor form was immunologically cross-reacted with hepatic aldehyde reductase. The two major enzyme forms were purified to homogeneity. One form, which had the highest amount in the tissue, was a monomeric protein with a molecular weight of 32,000 and isoelectric point of 4.2, showed strict specificity for benzene dihydrodiol and NADP+, and reduced pyridine aldehydes, glyceraldehyde and diacetyl at low rates. Another form, with a molecular weight of 36,000 and isoelectric point of 5.0, oxidized n-butanol, glycerol and sorbitol as well as benzene dihydrodiol in the presence of NADP+ or NAD+, and exhibited much higher reductase activity towards various aldehydes, aldoses and diacetyl. The pI 5.0 form was more sensitive to inhibition by sorbinil and p-chloromercuriphenyl sulfonate than the pI 4.2 form and was activated by sulfate ion. The two enzymes did not catalyze the oxidation of hydroxysteroids and xenobiotic alicyclic alcohols and were immunologically different from hepatic 17 beta-hydroxysteroid-dihydrodiol dehydrogenase. The results indicate that guinea-pig testis contains at least two dihydrodiol dehydrogenases distinct from the hepatic enzymes, one of which, the pI 5.0 enzyme form, may be identical to aldose reductase.  相似文献   

6.
Based on requirements for acetate or lipoic acid for aerobic (but not anaerobic) growth, Lactococcus lactis subsp. lactis mutants with impaired pyruvate catabolism were isolated following classical mutagenesis. Strains with defects in one or two of the enzymes, pyruvate formate-lyase (PFL), lactate dehydrogenase (LDH) and the pyruvate dehydrogenase complex (PDHC) were obtained. Growth and product formation of these strains were characterized. A PFL-defective strain (requiring acetate for anaerobic growth) displayed a two-fold increase in specific lactate production compared with the corresponding wild-type strain when grown anaerobically. LDH defective strains directed 91-96% of the pyruvate towards alpha-acetolactate, acetoin and diacetyl production when grown aerobically in the presence of acetate and absence of lipoic acid (a similar characteristic was observed in an LDH and PDHC defective strain in the presence of both acetate and lipoic acid) and more than 65% towards formate, acetate and ethanol production under anaerobic conditions. Another strain with defective PFL and LDH was strictly aerobic. However, a variant with strongly enhanced diacetyl reductase activities (NADH/NAD+ dependent diacetyl reductase, acetoin reductase and butanediol dehydrogenase activities) was selected from this strain under anaerobic conditions by supplementing the medium with acetoin. This strain is strictly aerobic, unless supplied with acetoin.  相似文献   

7.
A NAD (P)-linked alcohol dehydrogenase was isolated from the soluble extract of the strictly respiratory bacterium Alcaligenes eutrophus N9A. Derepression of the formation of this enzyme occurs only in cells incubated under conditions of restricted oxygen supply for prolonged times. The purification procedure included precipitation by cetyltrimethylammonium bromide and ammonium sulfate and subsequent chromatography on DEAE-Sephacel, Cibacron blue F3G-A Sepharose and thiol-Sepharose. The procedure resulted in a 120-fold purification of a multifunctional alcohol dehydrogenase exhibiting dehydrogenase activities for 2,3-butanediol, ethanol and acetaldehyde and reductase activities for diacetyl, acetoin and acetaldehyde. During purification the ratio between 2,3-butanediol dehydrogenase and ethanol dehydrogenase activity remained nearly constant. Recovering about 20% of the initial 2,3-butanediol dehydrogenase activity, the specific activity of the final preparation was 70.0 U X mg protein-1 (2,3-butanediol oxidation) and 2.8 U X mg protein-1 (ethanol oxidation). The alcohol dehydrogenase is a tetramer of a relative molecular mass of 156000 consisting of four equal subunits. The determination of the Km values for different substrates and coenzymes as well as the determination of the pH optima for the reactions catalyzed resulted in values which were in good agreement with the fermentative function of this enzyme. The alcohol dehydrogenase catalyzed the NAD (P)-dependent dismutation of acetaldehyde to acetate and ethanol. This reaction was studied in detail, and its possible involvement in acetate formation is discussed. Among various compounds tested for affecting enzyme activity only NAD, NADP, AMP, ADP, acetate and 2-mercaptoethanol exhibited significant effects.  相似文献   

8.
Aliphatic chain compounds at least four carbons long with vicinal carbonyl groups in the 2,3 positions were detected by the Westerfeld test. Acetoin, which has one carbonyl group and an adjacent hydroxyl group, gave positive results, but methyl action (3-hydroxy-3-methyl-2-butanone) was negative, and subsequent tests supported the conclusion that acetoin is oxidized to diacetyl by alpha-naphthol during the Westerfeld test in the absence or presence of air. 2,3-Pentanedione and 2,3-heptanedione gave positive results, but equimolar concentrations of these compounds gave maximal absorbancy readings that were only 35% (2,3-pentanedione) and 31% (2,3-heptanedione) of those obtained with diacetyl or acetoin. Negative results were obtained with pyruvic acid, 2,3-butylene glycol, and carbon ring compounds (1,2-cyclohexanedione, alloxan, and 3,4-dihydroxy-3-cyclobutene-1,2-dione). alpha-Naphtho could not be replaced in the test by beta-naphthol, 1,2,3,4,-tetrahydroxy-1-naphthol, or 5,6,7,8-tetrahydroxy-1-naphthol. Creatine could not be replaced by arginine, guanidine . HCl, or guanidinoacetic acid.  相似文献   

9.
The mechanism of the formation of 2,3-butanediol isomers in Bacillus polymyxa was studied. We proposed a new model with NADPH-linked diacetyl reductase (S-acetoin forming) and R(−)-2,3-butanediol dehydrogenase. The two enzymes were separated by Blue Sepharose CL-6B and their stereospecificities were identified using all of the pure isomers of 2,3-butanediol (R(−), S(+)m, and meso), acetoin (R(−) and S(+)) and the separation and measurement of these isomers. The presence of acetoin or butanediol racemase was not confirmed in our experiments.  相似文献   

10.
Aliphatic chain compounds at least four carbons long with vicinal carbonyl groups in the 2,3 positions were detected by the Westerfeld test. Acetoin, which has one carbonyl group and an adjacent hydroxyl group, gave positive results, but methyl action (3-hydroxy-3-methyl-2-butanone) was negative, and subsequent tests supported the conclusion that acetoin is oxidized to diacetyl by alpha-naphthol during the Westerfeld test in the absence or presence of air. 2,3-Pentanedione and 2,3-heptanedione gave positive results, but equimolar concentrations of these compounds gave maximal absorbancy readings that were only 35% (2,3-pentanedione) and 31% (2,3-heptanedione) of those obtained with diacetyl or acetoin. Negative results were obtained with pyruvic acid, 2,3-butylene glycol, and carbon ring compounds (1,2-cyclohexanedione, alloxan, and 3,4-dihydroxy-3-cyclobutene-1,2-dione). alpha-Naphtho could not be replaced in the test by beta-naphthol, 1,2,3,4,-tetrahydroxy-1-naphthol, or 5,6,7,8-tetrahydroxy-1-naphthol. Creatine could not be replaced by arginine, guanidine . HCl, or guanidinoacetic acid.  相似文献   

11.
The genes involved in the 2,3-butanediol pathway coding for alpha-acetolactate decarboxylase, alpha-acetolactate synthase (alpha-ALS), and acetoin (diacetyl) reductase were isolated from Klebsiella terrigena and shown to be located in one operon. This operon was also shown to exist in Enterobacter aerogenes. The budA gene, coding for alpha-acetolactate decarboxylase, gives in both organisms a protein of 259 amino acids. The amino acid similarity between these proteins is 87%. The K. terrigena genes budB and budC, coding for alpha-ALS and acetoin reductase, respectively, were sequenced. The 559-amino-acid-long alpha-ALS enzyme shows similarities to the large subunits of the Escherichia coli anabolic alpha-ALS enzymes encoded by the genes ilvB, ilvG, and ilvI. The K. terrigena alpha-ALS is also shown to complement an anabolic alpha-ALS-deficient E. coli strain for valine synthesis. The 243-amino-acid-long acetoin reductase has the consensus amino acid sequence for the insect-type alcohol dehydrogenase/ribitol dehydrogenase family and has extensive similarities with the N-terminal and internal regions of three known dehydrogenases and one oxidoreductase.  相似文献   

12.
The NADPH-linked diacetyl reductase system from the cytosolic fraction of Saccharomyces cerevisiae has been resolved into two oxidoreductases catalyzing irreversibly the enantioselective reduction of diacetyl (2,3-butanedione) to (S)- and (R)-acetoin (3-hydroxy-2-butanone) [so-called (S)- and (R)-diacetyl reductases] (EC 1.1.1.5) which have been isolated to apparent electrophoretical purity. The clean-up procedures comprising streptomycin sulfate treatment, Sephadex G-25 filtration, DEAE-Sepharose CL-6B column chromatography, affinity chromatography on Matrex Gel Red A and Superose 6 prep grade filtration led to 120-fold and 368-fold purifications, respectively. The relative molecular mass of the (R)-diacetyl reductase, estimated by means of HPLC filtration on Zorbax GF 250 and sodium dodecyl sulfate/polyacrylamide gel electrophoresis, was 36,000. The (R)-enzyme was most active at pH 6.4 and accepted in addition to diacetyl C5-, C6-2,3-diketones, 1,2-cyclohexanedione, 2-oxo aldehydes and short-chain 2- and 3-oxo esters as substrates. The enzyme was characterized by high enantioselectivity and regiospecificity. The Km values for diacetyl and 2,3-pentanedione were determined as 2.0 mM. The Mr of the (S)-diacetyl reductase was determined as 75,000 by means of HPLC filtration of Zorbax GF 250. The enzyme decomposed into subunits of Mr 48,000 and 24,000 on sodium dodecyl sulfate/polyacrylamide gel electrophoresis. The optimum pH was 6.9. The purified (S)-enzyme reduced stereospecifically a broad spectrum of substrates, comprising 2,3-, 2,4- and 2,5-diketones, 2-oxo aldehydes, 1,2-cyclohexanedione and methyl ketones as well as 3-, 4- and 5-oxo esters. The 2,3- and 2,4-diketones are transformed to the corresponding (S)-2-hydroxy ketones; 2,5-hexanedione, however, was reduced to (S,S)-2,5-hexanediol. The Km values for diacetyl and 2,3-pentanedione were estimated as 2.3 and 1.5 mM, respectively. Further characterization of the (S)-diacetyl reductase revealed that it is identical with the so-called '(S)-enzyme', involved in the enantioselective reduction of 3-, 4- and 5-oxo esters in baker's yeast.  相似文献   

13.
An enzyme capable of reducing acetoin in the presence of NADH was purified from Mycobacterium sp. B-009, a non-clinical bacterial strain of soil origin. The enzyme is a homotetramer and can be classified as a medium-chain alcohol dehydrogenase/reductase based on the molecular weight of the monomer. Identification of the structural gene revealed a limited distribution of homologous genes only among actinomycetes. In addition to its activity as a reductase specific for (S)-acetoin (EC 1.1.1.76), the enzyme showed both diacetyl reductase (EC 1.1.1.304) and NAD+-dependent alcohol dehydrogenase (EC 1.1.1.1) activities. (S)-Acetoin and diacetyl reductases belong to a group of short-chain alcohol dehydrogenase/reductases but do not have superior abilities to dehydrogenate monoalcohols. Thus, the purified enzyme can be readily distinguished from other enzymes. We used the dual functionality of the enzyme to effectively reduce diacetyl to (S)-acetoin, coupled with the oxidation of 1-butanol.  相似文献   

14.
The presently best known and largest group of quinoproteins consists of enzymes using the cofactor 2,7,9-tricarboxy-1H-pyrrolo[2,3-f]quinoline- 4,5-dione (PQQ), a compound having a pyrrole ring fused to a quinoline ring with an o-quinone group in it. Representatives of this group are found among the bacterial, NAD(P)-independent, periplasmic dehydrogenases. Despite their high midpoint redox potential, the overall behaviour of quinoprotein dehydrogenases is similar to that of their counterparts, those using a flavin cofactor or a nicotinamide coenzyme. Apart from an exceptional Gram-positive one, the sole organisms where the presence of PQQ has really been established are Gram-negative bacteria. Evidence for the occurrence of covalently bound PQQ is lacking since it has now been shown that several enzymes previously considered to contain this prosthetic group do not in fact do so. Another group of quinoproteins, consisting of amine oxidoreductases, has a protein chain containing one of the following quinonoid aromatic amino acids: 6-hydroxy-phenylalanine-3,4-dione (TPQ) or 4-(2'-tryptophyl)-tryptophan-6,7-dione (TTQ). There is no doubt that these o-quinones play a role as cofactor, in the case of TPQ in prokaryotic as well as eukaryotic amine oxidases. It appears, therefore, that a novel class of amino-acid-derived cofactors is emerging, ranging from the free radical form of tyrosine and tryptophan to those containing a dicarbonyl group (like the already known pyryvoyl group and the o-quinones here described.  相似文献   

15.
Bacterial 2,3-butanediol dehydrogenases   总被引:3,自引:0,他引:3  
Enterobacter aerogenes, Aeromonas hydrophila, Serratia marcescens and Staphylococcus aureus possessing L(+)-butanediol dehydrogenase produced mainly meso-butanediol and small amounts of optically active butanediol; Acetobacter suboxydans, Bacillus polymyxa and Erwinia carotovora containing D(-)-butanediol dehydrogenase produced more optically active butanediol than meso-butanediol. Resting and growing cells of these organisms oxidized only one enantiomer of racemic butanediol. The D(-)-butanediol dehydrogenase from Bacillus polymyxa was partially purified (30-fold) with a specific activity of 24.5. Except NAD and NADH no other cofactors were required. Optimum pH-values for oxidation and reduction were pH 9 and pH 7, respectively. The optimum temperature was about 60°C. The molecular weight was 100000 to 107000. The K m-values were 3.3 mM for D(-)-butanediol, 6.25 mM for meso-butanediol, 0.53 mM for acetoin, 0.2 mM for NAD, 0.1 mM for NADH, 87 mM for diacetyl, 38 mM for 1,2-propanediol; 2,3-pentanedion was not a substrate for this enzyme. The L(+)-butanediol dehydrogenase from Serratia marcescens was purified 57-fold (specific activity 22.3). Besides NAD or NADH no cofactors were required. The optimum value for oxidation was about pH 9 and for reduction pH 4.5. The optimum temperature was 32–36°C. The molecular weight was 100000 to 107000. The K m-values were 5 mM for meso-butanediol, 10 mM for racemic butanediol, 6.45 for acetoin, 1 mM for NAD, 0.25 mM for NADH, 2.08 mM for diacetyl, 16.7 mM for 2,3-pentanedion and 11.8 mM for 1,2-propanediol.Abbreviations Bud 2,3-butanediol - DH dehydrogenase  相似文献   

16.
The three-dimensional structure of one of the three lipoamide dehydrogenases occurring in Pseudomonas putida, LipDH Val, has been determined at 2.45 A resolution. The orthorhombic crystals, grown in the presence of 20 mM NAD+, contain 458 residues per asymmetric unit. A crystallographic 2-fold axis generates the dimer which is observed in solution. The final crystallographic R-factor is 21.8% for 18,216 unique reflections and a model consisting of 3,452 protein atoms, 189 solvent molecules and 44 NAD+ atoms, while the overall B-factor is unusually high: 47 A2. The structure of LipDH Val reveals the conformation of the C-terminal residues which fold "back" into the putative lipoamide binding region. The C-terminus has been proven to be important for activity by site-directed mutagenesis. However, the distance of the C-terminus to the catalytically essential residues is surprisingly large, over 6 A, and the precise role of the C-terminus still needs to be elucidated. In this crystal form LipDH Val contains one NAD+ molecule per subunit. Its adenine-ribose moiety occupies an analogous position as in the structure of glutathione reductase. However, the nicotinamide-ribose moiety is far removed from its expected position near the isoalloxazine ring and points into solution. Comparison of LipDH Val with Azotobacter vinelandii lipoamide dehydrogenase yields an rms difference of 1.6 A for 440 well defined C alpha atoms per subunit. Comparing LipDH Val with glutathione reductase shows large differences in the tertiary and quaternary structure of the two enzymes. For instance, the two subunits in the dimer are shifted by 6 A with respect to each other. So, LipDH Val confirms the surprising differences in molecular architecture between glutathione reductase and lipoamide dehydrogenase, which were already observed in Azotobacter vinelandii LipDH. This is the more remarkable since the active sites are located at the subunit interface and are virtually identical in all three enzymes.  相似文献   

17.
Diacetyl removal from beer was studied with whole cells and crude enzyme extracts of yeasts and bacteria. Cells of Streptococcus diacetilactis 18-16 destroyed diacetyl in solutions at a rate almost equal to that achieved by the addition of whole yeast cells. Yeast cells impregnated in a diatomaceous earth filter bed removed all diacetyl from solutions percolated through the bed. Undialyzed crude enzyme extracts from yeast cells removed diacetyl very slowly from beer at its normal pH (4.1); at a pH of 5.0 or higher, rapid diacetyl removal was achieved. Dialyzed crude enzyme extracts from yeast cells were found to destroy diacetyl in a manner quite similar to that of diacetyl reductase from Aerobacter aerogenes, and both the bacterial and the yeast extracts were stimulated significantly by the addition of reduced nicotinamide adenine dinucleotide (NADH). Diacetyl reductase activity of four strains of A. aerogenes was compared; three of the strains produced enzyme with approximately twice the specific activity of the other strain (8724). Gel electrophoresis results indicated that at least three different NADH-oxidizing enzymes were present in crude extracts of diacetyl reductase. Sephadex-gel chromotography separated NADH oxidase from diacetyl reductase. It was also noted that ethyl alcohol concentrations approximately equivalent to those found in beer were quite inhibitory to diacetyl reductase.  相似文献   

18.
1. [(14)C]Acetoin was enzymically synthesized from [(14)C]pyruvate with a pyruvate decarboxylase preparation. Its optical activity was [alpha](20) (d)-78 degrees . 2. Large amounts (1000-fold higher than physiological concentrations) of acetoin were incubated with rat liver mince. Acetoin disappeared but very little (14)CO(2) was evolved. A compound accumulated, which was purified and identified as butane-2,3-diol. Chromatography on borate-impregnated paper indicated the presence of both the erythro and threo forms. 3. Liver extracts capable of interconverting biacetyl, acetoin and butane-2,3-diol were obtained. These interconversions were catalysed by two different enzymes: acetoin dehydrogenase (EC 1.1.1.5) and butane-2,3-diol dehydrogenase (EC 1.1.1.4), previously identified in bacteria. Both required NAD(+) or NADP(+) as cofactors and were different from alcohol dehydrogenase. The equilibrium in both cases favoured the more reduced compound. 4. The activity of butane-2,3-diol dehydrogenase was decreased by dialysis against EDTA: the addition of Co(2+), Cu(2+), Zn(2+) and other bivalent metal ions restored activity. 5. Biacetyl reductase was resolved into multiple forms by CM-Sephadex chromatography and electrophoresis.  相似文献   

19.
In this report, we first cloned a cDNA for a protein that is highly expressed in mouse kidney and then isolated its counterparts in human, rat hamster, and guinea pig by polymerase chain reaction-based cloning. The cDNAs of the five species encoded polypeptides of 244 amino acids, which shared more than 85% identity with each other and showed high identity with a human sperm 34-kDa protein, P34H, as well as a murine lung-specific carbonyl reductase of the short-chain dehydrogenase/reductase superfamily. In particular, the human protein is identical to P34H, except for one amino acid substitution. The purified recombinant proteins of the five species were about 100-kDa homotetramers with NADPH-linked reductase activity for alpha-dicarbonyl compounds, catalyzed the oxidoreduction between xylitol and l-xylulose, and were inhibited competitively by n-butyric acid. Therefore, the proteins are designated as dicarbonyl/l-xylulose reductases (DCXRs). The substrate specificity and kinetic constants of DCXRs for dicarbonyl compounds and sugars are similar to those of mammalian diacetyl reductase and l-xylulose reductase, respectively, and the identity of the DCXRs with these two enzymes was demonstrated by their co-purification from hamster and guinea pig livers and by protein sequencing of the hepatic enzymes. Both DCXR and its mRNA are highly expressed in kidney and liver of human and rodent tissues, and the protein was localized primarily to the inner membranes of the proximal renal tubules in murine kidneys. The results imply that P34H and diacetyl reductase (EC ) are identical to l-xylulose reductase (EC ), which is involved in the uronate cycle of glucose metabolism, and the unique localization of the enzyme in kidney suggests that it has a role other than in general carbohydrate metabolism.  相似文献   

20.
Expression of the tripeptide permease gene tppB is anaerobically induced. This induction is independent of the fnr (oxrA) gene product, which is known to be required for the anaerobic induction of several respiratory enzymes. We isolated, characterized, and mapped mutations in two genes, oxrC and tppR, which prevent the anaerobic induction of tppB expression. Mutations in oxrC were highly pleiotropic, preventing the anaerobic expression of the formate dehydrogenase component of formate hydrogen lyase (fhl), a tripeptidase (pepT), and two of the three known hydrogenase isoenzymes (hydrogenases 1 and 3). On the other hand, expression of nitrate reductase, fumarate reductase, and a number of other fnr (oxrA)-dependent enzymes was not affected by mutations in oxrC. Thus, there appeared to be at least two distinct classes of anaerobically induced genes, those which required fnr for their expression and those which required oxrC. It seems that fnr-dependent enzymes perform primarily respiratory functions, whereas oxrC-dependent enzymes served fermentative or biosynthetic roles. We found the primary defect of oxrC mutants to be a deficiency in phosphoglucose isomerase activity, implying that a product of glycolysis functions as an anaerobic regulatory signal. Mutations in tppR were specific for tppB and did not affect expression of other oxrC-dependent genes. However, tppR did exhibit phenotypes other than the regulation of tppB. Both oxrC and tppR mutants were hypersensitive to the toxic NAD analog 6-aminonicotinic acid. This suggests that oxrC and tppR may play a role in the regulation of NAD biosynthesis or, alternatively, that NAD or a related nucleotide serves as the anaerobic signal for oxrC-dependent enzymes.  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