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 9 毫秒
1.
NAD-Dependent formate dehydrogenase (FDH) has been isolated from methylotrophyc strain Bacterium sp 1 by (NH4)2SO4 fractionation of cell extract, ion-exchange chromatography and preparative isotachophoresis. Preparation of FDH is homogeneous in analytical polyacrylamide gel electrophoresis and under ultracentrifugation. Sedimentation coefficient of FDH is 4.9S. Mikhaelis constants are 1.1-10(-4) M for NAD and 1.5-10(-2) M for formate. In the absence of sulfhydril compounds FDH is unstable, but it is stable in the presence of mercaptoethanol or ditiotreitol.  相似文献   

2.
A panel of 4 monoclonal antibodies and 7 polyclonal antisera against NAD-dependent formate dehydrogenase from methylotrophic bacterium Pseudomonas sp. 101 has been obtained. The reactivity of the 37 overlapping proteolytic peptides with the monoclonal antibodies and polyclonal antisera has been studied with ELISA test. The data obtained were interpreted residing on the structural model of the formate dehydrogenase at 3 Å resolution. The immunodominant regions in the formate dehydrogenase molecule and the epitopes for the monoclonal antibodies were elucidated.  相似文献   

3.
For better understanding of the coenzyme specificity in NAD-dependent MDH (tMDH) from Thermus flavus AT-62, we determined the crystal structures of tMDH-NADP(H) complex at maximally 1.65 A resolution. The overall structure is almost the same as that of the tMDH-NADH complex. However, NADP(H) binds to tMDH in the reverse orientation, where adenine occupies the position near the catalytic center and nicotinamide is positioned at the adenine binding site of the tMDH-NADH complex. Consistent with this, kinetic analysis of the malate-oxidizing reaction revealed that NADP(+) inhibited tMDH at high concentrations. This has provided the first evidence for the alternative binding mode of the nicotinamide coenzyme, that has pseudo-symmetry in its structure, in a single enzyme.  相似文献   

4.
The comparative analysis of the primary and tertiary structures of NAD-dependent bacterial formate dehydrogenase (FDH) from methylotrophic bacterium Pseudomonas sp. 101 and a number of structurally characterized NAD-dependent dehydrogenases were performed. FDH has a highly conservative fold of the coenzyme binding domain. Position of the symmetry axis in the FDH molecule relative to the beta-sheets of its coenzyme binding domain with the respective sequences of the other NAD-dependent enzymes was performed on the basis of the spatial homology between these structures. Only one of the three amino acid residues previously thought to be conserved in the coenzyme binding domains of NAD-dependent dehydrogenases is preserved in the FDH molecule (Asp-221). Two glycine residues found in all previously studied dehydrogenases are substituted in FDH by Ala-198 and Pro-256, respectively. Position of the essential thiol of FDH (Cys-255) in the protein structure was established. It is suggested that Cys-255 is situated on or near polypeptide locus taking part in the conformational changes of the protein in the course of the catalysis.  相似文献   

5.
1. NAD-dependent formate dehydrogenase was isolated from gram-negative methylotrophic bacteria, strain 1, grown on methanol. The purification procedure involved ammonium sulfate fractionation, ion-exchange chromatography and preparative isotachophoresis or gel filtration; it resulted in a yield of 40%. 2. The final enzyme preparations were homogeneous as judged by sedimentation in an ultracentrifuge. Formate dehydrogenase purified in the presence of EDTA reveals two bands on electrophoresis in polyacrylamide gel both after protein and activity staining. Two components are transformed into a single one after prolonged storage in the presence of 2-mercaptoethanol. 3. Formate dehydrogenase is a dimer composed of identical or very similar subunits. The molecular weight of the enzyme is about 80 000. 4. Amino acid composition and some other physico-chemical properties of the enzyme were studied. 5. Formate dehydrogenase is specific for formate and NAD as electron acceptor. The Michaelis constant was 0.11 mM for NAD and 15 mM for formate (pH 7.0, 37 degrees C). 6. Formate dehydrogenase was rapidly inactivated in the absence of -SH compounds. The enzyme retained full activity upon storage at ambient temperature in solution for half a year in the presence of 2-mercaptoethanol or EDTA.  相似文献   

6.
The primary structure of NAD-dependent formate dehydrogenase from methylotrophic bacterium Pseudomonas sp. 101 is determined. The enzyme is composed of two identical subunits, each comprising 393 amino acid residues, and has a molecular weight of 43.1 kD. To elucidate the protein's amino acid sequence, four types of digestion were used: cyanogen bromide cleavage at methionine residues, endoproteinase Lys-C digestion at lysine residues, endoproteinase Glu-C cleavage at glutamic acid residues, and tryptic digestion. The peptides obtained were purified to homogeneity and characterized.  相似文献   

7.
The gene of the NAD-dependent formate dehydrogenase (FDH) from the yeast Candida boidinii was cloned by PCR using genomic DNA as a template. Expression of the gene in Escherichia coli yielded functional FDH with about 20% of the soluble cell protein. To confirm the hypothesis of a thiol-coupled inactivation process, both cysteine residues in the primary structure of the enzyme have been exchanged by site-directed mutagenesis using a homology model based on the 3D structure of FDH from Pseudomonas sp. 101 and from related dehydrogenases. Compared to the wt enzyme, most of the mutants were significantly more stable towards oxidative stress in the presence of Cu(II) ions, whereas the temperature optima and kinetic constants of the enzymatic reaction are not significantly altered by the mutations. Determination of the Tm values revealed that the stability at temperatures above 50 degrees C is optimal for the native and the recombinant wt enzyme (Tm 57 degrees C), whereas the Tm values of the mutant enzymes vary in the range 44-52 degrees C. Best results in initial tests concerning the application of the enzyme for regeneration of NADH in biotransformation of trimethyl pyruvate to Ltert leucine were obtained with two mutants, FDHC23S and FDHC23S/C262A, which are significantly more stable than the wt enzyme.  相似文献   

8.
In many plant species, one of the most abundant soluble proteins (as judged by two-dimensional polyacrylamide gel electrophoresis) in mitochondria from nongreen tissues is a 40-kD polypeptide that is relatively scarce in mitochondria from photosynthetic tissues. cDNA sequences encoding this polypeptide were isolated from a lambda gt11 cDNA expression library from potato (Solanum tuberosum L.) by screening with a specific antibody raised against the 40-kD polypeptide. The cDNA sequence contains an open reading frame of 1137 nucleotides whose predicted amino acid sequence shows strong homology to an NAD-dependent formate dehydrogenase (EC 1.2.1.2) from Pseudomonas sp. 101. Comparison of the cDNA sequence with the N-terminal amino acid sequence of the mature 40-kD polypeptide suggests that the polypeptide is made as a precursor with a 23-amino acid presequence that shows characteristics typical of mitochondrial targeting signals. The identity of the polypeptide was confirmed by assaying the formate dehydrogenase activity in plant mitochondria from various tissues and by activity staining of mitochondrial proteins run on native gels combined with antibody recognition. The abundance and distribution of this protein suggest that higher plant mitochondria from various nonphotosynthetic plant tissues (tubers, storage roots, seeds, dark-grown shoots, cauliflower heads, and tissues grown in vitro) might contain a formate-producing fermentation pathway similar to those described in bacteria and algae.  相似文献   

9.
Three heavy atom isomorphous derivatives were used for the X-ray analysis of the holo form of NAD-dependent bacterial formate dehydrogenase (ternary complex enzyme-NAD-azide) at 3.0 A resolution. The enzyme subunit contains a catalytic and a coenzyme binding domain, with the active centre and the coenzyme binding site in the cleft between the domains. The polypeptide chain's fold and the position of 393 C alpha-atoms were determined. The secondary structure of the formate dehydrogenase was resolved. The structure of the NAD-binding domain is shown to be similar to that of other NAD-dependent enzymes.  相似文献   

10.
The three-dimensional structures of NAD-dependent D-lactate dehydrogenase (D-LDH) and formate dehydrogenase (FDH), which resemble each other, imply that the two enzymes commonly employ certain main chain atoms, which are located on corresponding loop structures in the active sites of the two enzymes, for their respective catalytic functions. These active site loops adopt different conformations in the two enzymes, a difference likely attributable to hydrogen bonds with Asn97 and Glu141, which are also located at equivalent positions in D-LDH and FDH, respectively. X-ray crystallography at 2.4-A resolution revealed that replacement of Asn97 with Asp did not markedly change the overall protein structure but markedly perturbed the conformation of the active site loop in Lactobacillus pentosus D-LDH. The Asn97-->Asp mutant D-LDH exhibited virtually the same k(cat), but about 70-fold higher K(M) value for pyruvate than the wild-type enzyme. For Paracoccus sp. 12-A FDH, in contrast, replacement of Glu141 with Gln and Asn induced only 5.5- and 4.3-fold increases in the K(M) value, but 110 and 590-fold decreases in the k(cat) values for formate, respectively. Furthermore, these mutant FDHs, particularly the Glu141-->Asn enzyme, exhibited markedly enhanced catalytic activity for glyoxylate reduction, indicating that FDH is converted to a 2-hydroxy-acid dehydrogenase on the replacement of Glu141. These results indicate that the active site loops play different roles in the catalytic reactions of D-LDH and FDH, stabilization of substrate binding and promotion of hydrogen transfer, respectively, and that Asn97 and Glu141, which stabilize suitable loop conformations, are essential elements for proper loop functioning.  相似文献   

11.
The NAD-dependent glutamate dehydrogenase from Candida utilis was isolated from 32P-labeled cells following enzyme inactivation promoted by glutamate starvation and found to exist in a phosphorylated form. Analysis of purified, fully active NAD-dependent glutamate dehydrogenase (a form) and inactive NAD-dependent glutamate dehydrogenase (b form) for alkalilabile phosphate revealed that the a form contained 0.09 +/- 0.06 mol of phosphate/mol of enzyme subunit and b form 1.25 +/- 0.06 mol of phosphate/mol of enzyme subunit. Phosphorylation caused a 10-fold reduction in enzyme specific activity. Dephosphorylation (release of 32P) and enzyme reactivation occurred on incubation with cell-free yeast extracts, indicating the presence of a phosphoprotein phosphatase in such preparations.  相似文献   

12.
Sorbitol dehydrogenase (SDH) is a distant relative to the alcohol dehydrogenases (ADHs) with sequence identities around 20%. SDH is a tetramer with one zinc ion per subunit. We have crystallized rat SDH and determined the structure by molecular replacement using a tetrameric bacterial ADH as search object. The conformation of the bound coenzyme is extended and similar to NADH bound to mammalian ADH but the interactions with the NMN-part have several differences with those of ADH. The active site zinc coordination in SDH is significantly different than in mammalian ADH but similar to the one found in the bacterial tetrameric NADP(H)-dependent ADH of Clostridiim beijerinckii. The substrate cleft is significantly more polar than for mammalian ADH and a number of residues are ideally located to position the sorbitol molecule in the active site. The SDH molecule can be considered to be a dimer of dimers, with subunits A–B and C–D, where the dimer interactions are similar to those in mammalian ADH. The tetramers are composed of two of these dimers, which interact with their surfaces opposite the active site clefts, which are accessible on the opposite side. In contrast to the dimer interactions, the tetramer-forming interactions are small with only few hydrogen bonds between side-chains.  相似文献   

13.
Sorbitol dehydrogenase (SDH) is a distant relative to the alcohol dehydrogenases (ADHs) with sequence identities around 20%. SDH is a tetramer with one zinc ion per subunit. We have crystallized rat SDH and determined the structure by molecular replacement using a tetrameric bacterial ADH as search object. The conformation of the bound coenzyme is extended and similar to NADH bound to mammalian ADH but the interactions with the NMN-part have several differences with those of ADH. The active site zinc coordination in SDH is significantly different than in mammalian ADH but similar to the one found in the bacterial tetrameric NADP(H)-dependent ADH of Clostridiim beijerinckii. The substrate cleft is significantly more polar than for mammalian ADH and a number of residues are ideally located to position the sorbitol molecule in the active site. The SDH molecule can be considered to be a dimer of dimers, with subunits A-B and C-D, where the dimer interactions are similar to those in mammalian ADH. The tetramers are composed of two of these dimers, which interact with their surfaces opposite the active site clefts, which are accessible on the opposite side. In contrast to the dimer interactions, the tetramer-forming interactions are small with only few hydrogen bonds between side-chains.  相似文献   

14.
15.
Eucaryotes possess one or more NADP-dependent methylene-THF dehydrogenases as part of multifunctional enzymes. In addition, yeast expresses an unusual monofunctional NAD-dependent enzyme, yMTD. We report X-ray structures for the apoenzyme and its complex with NAD+ at 2.8 and 3.0 A resolution, respectively. The protein fold resembles that seen for the human and Escherichia coli dehydrogenase/cyclohydrolase bifunctional enzymes. The enzyme has two prominent domains, with the active site cleft between them. yMTD has a noncanonical NAD-binding domain that has two inserted strands compared with the NADP-binding domains of the bifunctional enzymes. This insert precludes yMTD from dimerizing in the same way as the bifunctional enzymes. yMTD functions as a dimer, but the mode of dimerization is novel. It does not appear that the difference in dimerization accounts for the difference in cofactor specificity or for the loss of cyclohydrolase activity. These functional differences are probably accounted for by minor differences within the tertiary structure of the active site of the monomeric protein.  相似文献   

16.
The extremely thermostable NAD-dependent glutamate dehydrogenase (NAD-GluDH) from Pyrobaculum islandicum, a member of the Crenarchaeota, was crystallized, and its 3D structure has been determined by X-ray diffraction methods. The homohexameric structure of Pb. islandicum glutamate dehydrogenase (Pis-GluDH) was solved and refined at a resolution of 2.9A with a crystallographic R-factor of 19.9% (Rfree 26.0%). The structure indicates that each subunit consists of two domains separated by a deep cleft containing an active site. The secondary structural elements and catalytically important residues of the enzyme were highly conserved among the NAD(P)-dependent GluDHs from other sources. A structural comparison of Pis-GluDH with other NAD(P)-dependent GluDHs suggests that a significant difference in the alpha8-loop-alpha9 region of this enzyme is associated with its coenzyme specificity. From the analysis of the 3D structure, hydrophobic interactions between intersubunits were found to be important features for the enzyme oligomerization. It has been reported that Pis-GluDH is highly thermostable, like the GluDH of the hyperthermophilic archaeum Pyrococcus furiosus, and the increase in the intersubunit ion pair networks is responsible for the extreme thermostability of the Pc. furiosus enzyme. However, the number of intersubunit ion pairs in the Pis-GluDH molecules is much smaller than those of the Pc. furiosus GluDH. The number of hydrophobic interactions at the intersubunit interfaces were increased and responsible for the extremely high thermostability. This indicates that the major molecular strategy for high thermostability of the GluDHs may be different for each hyperthermophile.  相似文献   

17.
Formate dehydrogenase N (FDH-N) of Escherichia coli is a membrane-bound enzyme comprising FdnG, FdnH, and FdnI subunits organized in an (alphabetagamma)3 configuration. The FdnG subunit carries a Tat-dependent signal peptide, which localizes the protein complex to the periplasmic side of the membrane. We noted that substitution of the first arginine (R5) in the twin arginine signal sequence of FdnG for a variety of other amino acids resulted in a dramatic (up to 60-fold) increase in the levels of protein synthesized. Bioinformatic analysis suggested that the mRNA specifying the first 17 codons of fdnG forms a stable stem-loop structure. A detailed mutational analysis has demonstrated the importance of this mRNA stem-loop in modulating FDH-N translation.  相似文献   

18.
In the l-Serine biosynthesis, D-3-phosphoglycerate dehydrogenase (PGDH) catalyzes the inter-conversion of D-3-phosphoglycerate to phosphohydroxypyruvate. PGDH belongs to 2-hydroxyacid dehydrogenases family. We have determined the crystal structures of PGDH from Sulfolobus tokodaii (StPGDH) and Pyrococcus horikoshii (PhPGDH) using X-ray diffraction to resolution of 1.77 Å and 1.95 Å, respectively. The PGDH protomer from both species exhibits identical structures, consisting of substrate binding domain and nucleotide binding domain. The residues and water molecules interacting with the NAD are identified. The catalytic triad residues Glu-His-Arg are highly conserved. The residues involved in the dimer interface and the structural features responsible for thermostability are evaluated. Overall, structures of PGDHs with two domains and histidine at the active site are categorized as type IIIH and such PGDHs structures having this type are reported for the first time.  相似文献   

19.
Mannitol dehydrogenase, a mannitol:mannose 1-oxidoreductase, constitutes the first enzymatic step in the catabolism of mannitol in nonphotosynthetic tissues of celery (Apium graveolens L.). Endogenous regulation on the enzyme activity in response to environmental cues is critical in modulating tissue concentration of mannitol, which, importantly, contribute to stress tolerance of celery. The enzyme was purified to homogeneity from celery suspension cultures grown on D-mannitol as the carbon source. Mannitol dehydrogenase was purified 589-fold to a specific activity of 365 mumol h-1 mg-1 protein with a 37% yield of enzyme activity present in the crude extract. A highly efficient and simple purification protocol was developed involving polyethylene glycol fractionation, diethylaminoethyl-anion-exchange chromatography, and NAD-agarose affinity chromatography using NAD gradient elution. Sodium dodecylsulfate gel electrophoresis of the final preparation revealed a single 40-kD protein. The molecular mass of the native protein was determined to be approximately 43 kD, indicating that the enzyme is a monomer. Polyclonal antibodies raised against the enzyme inhibited enzymatic activity of purified mannitol dehydrogenase. Immunoblots of crude protein extracts from mannitol-grown celery cells and sink tissues of celery, celeriac, and parsley subjected to sodium dodecyl sulfate gel electrophoresis showed a single major immuno-reactive 40-kD protein.  相似文献   

20.
Kinetic studies of formate dehydrogenase   总被引:4,自引:1,他引:3       下载免费PDF全文
1. The kinetic mechanism of formate dehydrogenase is a sequential pathway. 2. The binding of the substrates proceeds in an obligatory order, NAD(+) binding first, followed by formate. 3. It seems most likely that the interconversion of the central ternary complex is extremely rapid, and that the rate-limiting step is the formation or possible isomerization of the enzyme-coenzyme complexes. 4. The secondary plots of the inhibitions with HCO(3) (-) and NO(3) (-) are non-linear, which suggests that more than one molecule of each species is able to bind to the same enzyme form. 5. The rate of the reverse reaction with carbon dioxide at pH6.0 is 20 times that with bicarbonate at pH8.0, although no product inhibition could be detected with carbon dioxide. The low rate of the reverse reaction precluded any steady-state analysis as the enzyme concentrations needed to obtain a measurable rate are of the same order as the K(m) values for NAD(+) and NADH.  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