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.
The gene encoding D-lactate dehydrogenase (D-lactate: NAD+ oxidoreductase, EC 1.1.1.28) of Lactobacillus plantarum has been sequenced, and expressed in Escherichia coli cells with an inducible expression plasmid, in which the 5'-noncoding region of the gene was replaced with the tac promoter. Comparison of the sequence of D-lactate dehydrogenase with L-lactate dehydrogenases, including the L. plantarum L-lactate dehydrogenase, showed no significant homology. In contrast, the D-lactate dehydrogenase is homologous to E. coli D-3-phosphoglycerate dehydrogenase and Lactobacillus casei D-2-hydroxyisocaproate dehydrogenase. This indicates that D-lactate dehydrogenase is a member of a new family of 2-hydroxyacid dehydrogenases recently proposed, being distinct from L-lactate dehydrogenase and L-malate dehydrogenase, and strongly suggests that the new family consists of D-isomer-stereospecific enzymes. In the reductive reaction, the enzyme showed a broad substrate specificity, although pyruvate was the most favorable of all 2-ketocarboxylic acids tested. In particular, hydroxypyruvate is effectively reduced by the enzyme, the reaction rate, and Km value being comparable to those in the case of pyruvate, indicating that the enzyme has not only D-lactate dehydrogenase activity but also D-glycerate dehydrogenase activity. The conserved residues in this family appear to be the residues involved in the substrate binding and the catalytic reaction, and thus to be targets for site-directed mutagenesis.  相似文献   

2.
K Rush  R Sbragia  C Wills 《FEBS letters》1986,198(1):89-91
A mutant lacking L-lactate dehydrogenase (EC 1.1.2.3) of Saccharomyces cerevisiae was isolated by its inability to grow on minimal medium with L-lactate as a carbon source. A simple activity gel assay for visualization of this enzyme and the two D-lactate dehydrogenases in this organism (EC 1.1.2.4 and 1.1.1.28) was developed. This enabled us to screen spontaneous and ethylmethanesulfonate-induced back mutants for electrophoretic mobility. Two mutants with a mobility faster than that of the wild type were isolated, and proved to be allelic to the L-lactate dehydrogenase negative mutant.  相似文献   

3.
A comparison of the primary structures of NAD(+)-dependent D-lactate dehydrogenase with L-lactate dehydrogenase and L-malate dehydrogenase failed to show any sequence similarity. However, D-2-hydroxyisocaproate dehydrogenase from Lactobacillus casei, glycerate dehydrogenase from cucumber, D-3-phosphoglycerate dehydrogenase and erythronate 4-phosphate dehydrogenase from Escherichia coli showed 38%, 24%, 24% and 22% amino acid identity, respectively. The profile analysis of the aligned sequences confirmed their relatedness. The hydropathy profiles of the aligned dehydrogenases were almost identical between residues 100-300 indicating largely preserved folding patterns of their polypeptide chains. The data suggest that L- and D-specific 2-hydroxy acid dehydrogenase genes evolved from two different ancestors and thus represent two different sets of enzyme families.  相似文献   

4.
Incubation of an NAD+-dependent succinic semialdehyde dehydrogenase from bovine brain with 4-dimethylaminoazobenzene-4-iodoacetamide (DABIA) resulted in a time-dependent loss of enzymatic activity. This inactivation followed pseudo first-order kinetics with a second-order rate constant of 168 m(-1).min(-1). The spectrum of DABIA-labeled enzyme showed a characteristic peak of the DABIA alkylated sulfhydryl group chromophore at 436 nm, which was absent from the spectrum of the native enzyme. A linear relationship was observed between DABIA binding and the loss of enzyme activity, which extrapolates to a stoichiometry of 8.0 mol DABIA derivatives per mol enzyme tetramer. This inactivation was prevented by preincubating the enzyme with substrate, succinic semialdehyde, but not by preincubating with coenzyme NAD+. After tryptic digestion of the enzyme modified with DABIA, two peptides absorbing at 436 nm were isolated by reverse-phase HPLC. The amino acid sequences of the DABIA-labeled peptides were VCSNQFLVQR and EVGEAICTDPLVSK, respectively. These sites are identical to the putative active site sequences of other brain succinic semialdehyde dehydrogenases. These results suggest that the catalytic function of succinic semialdehyde dehydrogenase is inhibited by the specific binding of DABIA to a cysteine residue at or near its active site.  相似文献   

5.
NAD(+)-dependent D-lactate dehydrogenase from Lactobacillus helveticus was purified to apparent homogeneity, and the sequence of the first 36 amino acid residues determined. Using forward and reverse oligonucleotide primers, based on the N-terminal sequence and amino acid residues 220-215 of the Lactobacillus bulgaricus enzyme [Kochhar, S., Hunziker, P. E., Leong-Morgenthaler, P. & Hottinger, H. (1992) J. Biol. Chem. 267, 8499-8513], a 0.6-kbp DNA fragment was amplified from L. helveticus genomic DNA by the polymerase chain reaction. This amplified DNA fragment was used as a probe to identify two recombinant clones containing the D-lactate dehydrogenase gene. Both plasmids overexpressed D-lactate dehydrogenase (greater than 60% total soluble cell protein) and were stable in Escherichia coli, compared to plasmids carrying the L. bulgaricus and Lactobacillus plantarum genes. The entire nucleotide sequence of the L. helveticus D-lactate dehydrogenase gene was determined. The deduced amino acid sequence indicated a polypeptide consisting of 336 amino acid residues, which showed significant amino acid sequence similarity to the recently identified family of D-2-hydroxy-acid dehydrogenases [Kochhar, S., Hunziker, P. E., Leong-Morgenthaler, P. & Hottinger, H. (1992) Biochem. Biophys. Res. Commun. 184, 60-66]. The physicochemical and catalytic properties of recombinant D-lactate dehydrogenase were identical to those of the wild-type enzyme, e.g. alpha 2 dimeric subunit structure, isoelectric pH, Km and Kcat for pyruvate and other 2-oxo-acid substrates. The kinetic profiles of 2-oxo-acid substrates showed some marked differences from that of L-lactate dehydrogenase, suggesting different mechanisms for substrate binding and specificity.  相似文献   

6.
Heat-stable and fructose-1,6-bisphosphate-activated L-lactate dehydrogenase (EC 1.1.1.27) has been purified from an extremely thermophilic bacterium, Thermus caldophilus GK24 [Taguchi, H., Yamashita, M., Matsuzawa, H. and Ohta, T. (1982) J. Biochem. (Tokyo) 91, 1343-1348]. N-terminal sequence analysis of the first 34 amino acids of the enzyme indicates that the N-terminal arm region (first 1-20 residues) known for the vertebrate L-lactate dehydrogenases is completely missing in the T. caldophilus enzyme, while there is a high homology of sequence between the regions which are considered to be part of the NAD-binding domain. The C-terminal amino acid of the enzyme was phenylalanine. Analysis of the amino acid composition showed that T. caldophilus enzyme contained much more arginine and fewer lysine than other bacterial and vertebrate L-lactate dehydrogenases. On modification reaction with 2,3-butanedione in the presence of NADH and oxamate, an enhanced activity of the T. caldophilus L-lactate dehydrogenase was obtained independently of fructose 1,6-bisphosphate, and the modified enzyme was desensitized to fructose 1,6-bisphosphate. Amino acid analysis indicated that such a desensitization in the active state was caused by the modification of only one arginine residue per the enzyme subunit. Desensitization of the enzyme was inhibited in the presence of fructose 1,6-bisphosphate. A similar desensitization was observed using 1,2-cyclohexanedione instead of 2,3-butanedione. The enzyme was irreversibly modified with 2,3-butanedione and characterized. The irreversibly modified enzyme also showed an enhanced activity independently of fructose 1,6-bisphosphate, and its pyruvate saturation curve was similar to that of the native enzyme measured in the presence of fructose 1,6-bisphosphate. Fructose 1,6-bisphosphate, which increases the thermostability of the native enzyme, did not affect that of the modified enzyme, while thermostability of the modified enzyme slightly decreased. Amino acid analysis indicated that only the arginine content was decreased by the modification. These results show that arginine residue(s) exist in the binding site for fructose 1,6-bisphosphate on the enzyme, and that the arginine residue(s) play some important role in the allosteric regulation of the enzyme activity.  相似文献   

7.
The activities of NAD-independent D- and L-lactate dehydrogenases (D-LDH, L-LDH) were detected in Rhodopseudomonas palustris No. 7 grown photoanaerobically on lactate. One of these enzymes, D-LDH, was purified as an electrophoretically homogeneous protein (M(r), about 235,000; subunit M(r) about 57,000). The pI was 5.0. The optimum pH and temperature of the enzyme were pH 8.5 and 50 degrees C, respectively. The Km of the enzyme for D-lactate was 0.8 mM. The enzyme had narrow substrate specificity (D-lactate and DL-2-hydroxybutyrate). The enzymatic activity was competitively inhibited by oxalate (Ki, 0.12 mM). The enzyme contained a FAD cofactor. Cytochrome c(2) was purified from strain No. 7 as an electrophoretically homogeneous protein. Its pI was 9.4. Cytochrome c(2) was reduced by incubating with D-LDH and D-lactate.  相似文献   

8.
The activity of the pyridine nucleotide-independent lactate dehydrogenase (iLDH) was characterized in mitochondria isolated from the protist Euglena gracilis. The dissociation constants for L- and D-lactate were similar, but the V(max) was higher with the d isomer. A ping-pong kinetic mechanism was displayed with 2,4-dichlorophenol-indolphenol (DCPIP), or coenzyme Q(1), reacting as the second substrate with the modified, reduced enzyme. Oxamate was a competitive inhibitor against both L- and D-lactate. Oxalate exerted a mixed-type inhibition regarding L- or D-lactate and also against DCPIP. The rate of L-lactate uptake was partially inhibited by mersalyl and lower than the rate of dehydrogenation, which was mersalyl-insensitive. These data suggested that the active site of L-iLDH was orientated toward the intermembrane space. The following observations indicated the existence of two stereo-specific iLDH enzymes in the inner membrane of Euglena mitochondria: a greater affinity of the D-iLDH for both inhibitors, D-iLDH thermo-stability at 70 degrees C and denaturation of L-iLDH, opposite signs in the enthalpy change for the association reaction of the isomers to the enzyme, differential solubilization of both activities with detergents, and different molecular mass.  相似文献   

9.
L-Mandelate dehydrogenase was purified from Acinetobacter calcoaceticus by Triton X-100 extraction from a 'wall + membrane' fraction, ion-exchange chromatography on DEAE-Sephacel, (NH4)2SO4 fractionation and gel filtration followed by further ion-exchange chromatography. The purified enzyme was partially characterized with respect to its subunit Mr (44,000), pH optimum (7.5), pI value (4.2), substrate specificity and susceptibility to various potential inhibitors including thiol-blocking reagents. FMN was identified as the non-covalently bound cofactor. The properties of L-mandelate dehydrogenase are compared with those of D-mandelate dehydrogenase, D-lactate dehydrogenase and L-lactate dehydrogenase from A. calcoaceticus.  相似文献   

10.
Heat-stable fructose 1,6-bisphosphate-dependent L-lactate dehydrogenase [EC 1.1.1.27] was purified from an extremely thermophilic bacterium, Thermus aquaticus YT-1. The amino acid composition and NH2-terminal 34 amino acid sequence of the enzyme were determined. Its NH2-terminal sequence shows high homology with those of Thermus caldophilus GK24 (82% identity) and some other bacterial L-lactate dehydrogenases (44-53% identity), indicating the close phylogenic relationship of the two Thermus species. At the same time, the two Thermus L-lactate dehydrogenases were found not to be identical not only chemically but also kinetically and immunologically. Citrate activated the T. aquaticus enzyme in the weak acidic pH region, while fructose 1,6-bisphosphate did in both acidic and neutral pH regions. The maximum activity obtained with citrate at pH 5.0 was about 2.5 times higher than that in the presence of fructose 1,6-bisphosphate at pH 6.7. The enzymes modified with 2,3-butanedione, acetic anhydride and diethyl pyrocarbonate in the presence of both NADH and oxamate were desensitized to fructose 1,6-bisphosphate, and the modified enzymes were active even in the absence of fructose 1,6-bisphosphate. All of the modified enzymes examined were still activated by citrate similarly to the native enzyme. These results suggest that the mechanism of activation by citrate is different from that by fructose 1,6-bisphosphate, and that the citrate-binding site is different from the fructose 1,6-bisphosphate-binding site.  相似文献   

11.
Bomati EK  Noel JP 《The Plant cell》2005,17(5):1598-1611
We describe the three-dimensional structure of sinapyl alcohol dehydrogenase (SAD) from Populus tremuloides (aspen), a member of the NADP(H)-dependent dehydrogenase family that catalyzes the last reductive step in the formation of monolignols. The active site topology revealed by the crystal structure substantiates kinetic results indicating that SAD maintains highest specificity for the substrate sinapaldehyde. We also report substantial substrate inhibition kinetics for the SAD-catalyzed reduction of hydroxycinnamaldehydes. Although SAD and classical cinnamyl alcohol dehydrogenases (CADs) catalyze the same reaction and share some sequence identity, the active site topology of SAD is strikingly different from that predicted for classical CADs. Kinetic analyses of wild-type SAD and several active site mutants demonstrate the complexity of defining determinants of substrate specificity in these enzymes. These results, along with a phylogenetic analysis, support the inclusion of SAD in a plant alcohol dehydrogenase subfamily that includes cinnamaldehyde and benzaldehyde dehydrogenases. We used the SAD three-dimensional structure to model several of these SAD-like enzymes, and although their active site topologies largely mirror that of SAD, we describe a correlation between substrate specificity and amino acid substitution patterns in their active sites. The SAD structure thus provides a framework for understanding substrate specificity in this family of enzymes and for engineering new enzyme specificities.  相似文献   

12.
The structural framework of cod liver alcohol dehydrogenase is similar to that of horse and human alcohol dehydrogenases. In contrast, the substrate pocket differs significantly, and main differences are located in three loops. Nevertheless, the substrate pocket is hydrophobic like that of the mammalian class I enzymes and has a similar topography in spite of many main-chain and side-chain differences. The structural framework of alcohol dehydrogenase is also present in a number of related enzymes like glucose dehydrogenase and quinone oxidoreductase. These enzymes have completely different substrate specificity, but also for these enzymes, the corresponding loops of the substrate pocket have significantly different structures. The domains of the two subunits in the crystals of the cod enzyme further differ by a rotation of the catalytic domains by about 6 degrees. In one subunit, they close around the coenzyme similarly as in coenzyme complexes of the horse enzyme, but form a more open cleft in the other subunit, similar to the situation in coenzyme-free structures of the horse enzyme. The proton relay system differs from the mammalian class I alcohol dehydrogenases. His 51, which has been implicated in mammalian enzymes to be important for proton transfer from the buried active site to the surface is not present in the cod enzyme. A tyrosine in the corresponding position is turned into the substrate pocket and a water molecule occupies the same position in space as the His side chain, forming a shorter proton relay system.  相似文献   

13.
Formaldehyde dehydrogenase (EC 1.2.1.1) is a widely occurring enzyme which catalyzes the oxidation of S-hydroxymethylglutathione, formed from formaldehyde and glutathione, into S-formyglutathione in the presence of NAD. We determined the amino acid sequences for 5 tryptic peptides (containing altogether 57 amino acids) from electrophoretically homogeneous rat liver formaldehyde dehydrogenase and found that they all were exactly homologous to the sequence of rat liver class III alcohol dehydrogenase (ADH-2). Formaldehyde dehydrogenase was found to be able at high pH values to catalyze the NAD-dependent oxidation of long-chain aliphatic alcohols like n-octanol and 12-hydroxydodecanoate but ethanol was used only at very high substrate concentrations and pyrazole was not inhibitory. The amino acid sequence homology and identical structural and kinetic properties indicate that formaldehyde dehydrogenase and the mammalian class III alcohol dehydrogenases are identical enzymes.  相似文献   

14.
NAD-linked lactate dehydrogenases specific for the D- and L-lactate have been demonstrated in a number of strains of unicellular cyanobacteria. The D-lactate dehydrogenase of one strain (Synechococcus 6716) was partially purified and its properties were studied. The enzyme has a molecular weight of ca. 115000-120000, is highly specific, autooxidizable, and susceptible to inhibition by iodoacetamide, oxamate and ATP. The possible physiological functions of the enzyme in the metabolism of the organism were investigated. D-lactate carbon was incorporated in cell material during photosynthetic growth with CO2, but lactate was not used as sole source for carbon for photosynthetic or chemosynthetic development. D-lactate and pyruvate were oxidized aerobically in the dark by resting cell suspensions with the assimilation mainly of the C2 and the C3 carbon atoms. In the oxidation of lactate, acetate was excreted into the medium. No fermentation of glucose was found, but a small amount of D-lactate was detected as a product of endogenous dark metabolism of the cell. All enzymes required for the production of lactate from glucose and from glycogen were found in exponentially growing cells, but the activity of some key enzymes was low or undetectable in old cultures.  相似文献   

15.
Aerobically grown yeast cells express mitochondrial lactate dehydrogenases that localize to the mitochondrial inner membrane. The D-lactate dehydrogenase is a zinc-flavoprotein with high acceptor specificity for cytochrome c, that catalyzes the oxidation of D-lactate into pyruvate. In this paper, we show that mitochondrial respiratory rate in phosphorylating or non-phosphorylating conditions with D-lactate as substrate is stimulated by carboxylic acids. This stimulation does not affect the yield of oxidative phosphorylation. Furthermore, this stimulation lies at the level of the D-lactate dehydrogenase. It is non-competitive, hyperbolic and its dimension is directly related to the number of carboxylic groups on the activator. The physiological meaning of such a regulation is discussed.  相似文献   

16.
L-Malate dehydrogenase from the extremely thermophilic mathanogen Methanothermus fervidus was isolated and its phenotypic properties were characterized. The primary structure of the protein was deducted from the coding gene. The enzyme is a homomeric dimer with a molecular mass of 70 kDa, possesses low specificity for NAD+ or NADP+ and catalyzes preferentially the reduction of oxalacetate. The temperature dependence of the activity as depicted in the Arrhenius and van't Hoff plots shows discontinuities near 52 degrees C, as was found for glyceraldehyde-3-phosphate dehydrogenase from the same organism. With respect to the primary structure, the archaebacterial L-malate dehydrogenase deviates strikingly from the eubacterial and eukaryotic enzymes. The sequence similarity is even lower than that between the L-malate dehydrogenases and L-lactate dehydrogenases of eubacteria and eukaryotes. The phylogenetic meaning of this relationship is discussed.  相似文献   

17.
Initial velocity, product inhibition, and substrate inhibition studies suggest that the endogenous lactate dehydrogenase activity of duck epsilon-crystallin follows an order Bi-Bi sequential mechanism. In the forward reaction (pyruvate reduction), substrate inhibition by pyruvate was uncompetitive with inhibition constant of 6.7 +/- 1.7 mM. In the reverse reaction (lactate oxidation), substrate inhibition by L-lactate was uncompetitive with inhibition constant of 158 +/- 25 mM. The cause of these inhibitions may be due to epsilon-crystallin-NAD(+)-pyruvate and epsilon-crystallin-NADH-L-lactate abortive ternary complex formation as suggested by the multiple inhibition studies. Pyruvate binds to free enzyme very poorly, with a very large dissociation constant. Bromopyruvate, fluoropyruvate, pyruvate methyl ester, and pyruvate ethyl ester are alternative substrates for pyruvate. 3-Acetylpyridine adenine dinucleotide, nicotinamide 1,N6-ethenoadenine dinucleotide, and nicotinamide hypoxanthine dinucleotide serve as alternative coenzymes for epsilon-crystallin. All the above alternative substrates or coenzymes showed an intersecting initial-velocity pattern conforming to the order Bi--Bi kinetic mechanism. Nicotinic acid adenine dinucleotide, thionicotinamide adenine dinucleotide, and 3-aminopyridine adenine dinucleotide acted as inhibitors for this enzymatic crystallin. The inhibitors were competitive versus NAD+ and noncompetitive versus L-lactate. alpha-NAD+ was a noncompetitive inhibitor with respect to the usual beta-NAD+. D-Lactate, tartronate, and oxamate were strong dead-end inhibitors for the lactate dehydrogenase activity of epsilon-crystallin. Both D-lactate and tartronate were competitive inhibitors versus L-lactate while oxamate was a competitive inhibitor versus pyruvate. We conclude that the structural requirements for the substrate and coenzyme of epsilon-crystallin are similar to those of other dehydrogenases and that the carboxamide carbonyl group of the nicotinamide moiety is important for the coenzyme activity.  相似文献   

18.
In Escherichia coli, the lct locus at min 80 on the chromosome map is associated with ability to grow on L-lactate and to synthesize a substrate-inducible flavin-linked dehydrogenase. Similar to that of the glpD-encoded aerobic glycerol-3-phosphate dehydrogenase, the level of induced enzyme activity is elevated by aerobiosis. Both of these controls are mediated by the two-component signal transduction system ArcB/ArcA, although sensitivity to the control is much more striking for L-lactate dehydrogenase. This study disclosed that the lct locus contained three overlapping genes in the clockwise order of lctD (encoding a flavin mononucleotide-dependent dehydrogenase), lctR (encoding a putative regulator), and lctP (encoding a permease) on the chromosomal map. These genes, however, are transcribed in the counterclockwise direction. No homology in amino acid sequence was found between aerobic glycerol-3-phosphate dehydrogenase and L-lactate dehydrogenase. A phi (lctD-lac) mutant was inducible by L-lactate but not D-lactate. Although the mutant lost the ability to grow on L-lactate, growth on D-lactate, known to depend on a different enzyme, remained normal.  相似文献   

19.
The function of the amino acid Thr246 in L-lactate dehydrogenase from Bacillus stearothermophilus has been investigated by site-directed replacement with glycine. Kinetic experiments with a number of 2-oxo acids showed strongly reduced activity for the mutated enzyme. However, the mutant enzyme shows a relative preference for the large hydrophobic sidechains of alpha-keto acids and an even higher specific activity than the wild-type lactate dehydrogenase for the polar oxaloacetate substrate. Graphic analyses indicate that the loss of one hydrogen bond, or intrusion of water into the active site, might be responsible for the reduced activity. The kinetic results suggest that the binding modes of bulky hydrophobic or polar substrates compensate to some degree for the partially disrupted active site.  相似文献   

20.
We investigated metabolic engineering of fermentation pathways in Escherichia coli for production of optically pure D- or L-lactate. Several pta mutant strains were examined, and a pta mutant of E. coli RR1 which was deficient in the phosphotransacetylase of the Pta-AckA pathway was found to metabolize glucose to D-lactate and to produce a small amount of succinate by-product under anaerobic conditions. An additional mutation in ppc made the mutant produce D-lactate like a homofermentative lactic acid bacterium. When the pta ppc double mutant was grown to higher biomass concentrations under aerobic conditions before it shifted to the anaerobic phase of D-lactate production, more than 62.2 g of D-lactate per liter was produced in 60 h, and the volumetric productivity was 1.04 g/liter/h. To examine whether the blocked acetate flux could be reoriented to a nonindigenous L-lactate pathway, an L-lactate dehydrogenase gene from Lactobacillus casei was introduced into a pta ldhA strain which lacked phosphotransacetylase and D-lactate dehydrogenase. This recombinant strain was able to metabolize glucose to L-lactate as the major fermentation product, and up to 45 g of L-lactate per liter was produced in 67 h. These results demonstrate that the central fermentation metabolism of E. coli can be reoriented to the production of D-lactate, an indigenous fermentation product, or to the production of L-lactate, a nonindigenous fermentation product.  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