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 NADP(+)-preferring glucose dehydrogenase from thermoacidophilic archaeon Thermoplasma acidophilum has been characterized, and its crystal structure has been determined (Structure, 2:385-393, 1994). Its sequence and structure are not homologous to bacterial NAD(P)(+)-dependent glucose dehydrogenases, and its molecular weight is also quite defferent. On the other hand, three functionally unknown genes with homologies to bacterial NAD(P)(+)-dependent glucose dehydrogenases have been sequenced as part of the T. acidophilum genome project (gene names: Ta0191, Ta0747, and Ta0754 respectively). We expressed two genes of three, Ta0191 and Ta0754, in Escherichia coli, and purified the gene products to homogeneity. Dehydrogenase activities were thereby detected from the purified proteins. The Ta0754 gene product exhibited aldohexose dehydrogenase activity, and the Ta0191 gene product exhibited weak 2-deoxyglucose dehydrogenase activity. No aldohexose dehydrogenase gene has been isolated, while the enzyme was reported in 1968. This is the first report of the gene and primary structure. The purified Ta0754 gene product, designated AldT, was characterized. The enzyme AldT effectively catalyzed the oxidation of various aldohexoses, especially D-mannose. Lower activities on D-2-deoxyglucose, D-xylose, D-glucose, and D-fucose were detected although no activities were shown on other aldohexoses or additional sugars. As a cofactor, NAD(+) was much more suitable for the activity than NADP(+). The NAD(+)-preferring dehydrogenase most effectively reacting to D-mannose is for the first time. AldT was most active at pH 10 and above 70 degrees C, and completely stable up to 60 degrees C after incubation for 15 min. Other enzymatic properties were also investigated.  相似文献   

2.
The escalation in genome sequencing has presented a mass of potentially useful new alcohol dehydrogenases (ADHs) in the form of putative open reading frame (ORF). To take advantage of such available resources, PCR primers based on the genome sequence of Bacillus thuringiensis serovar israelensis were used to clone a gene encoding a hypothetical alcohol dehydrogenase (named as BtADH). Activity studies of the translation product revealed that the alcohol dehydrogenases catalyse the inter-conversion of aliphatic aldehydes and corresponding primary alcohol with chain length of two to ten carbons. The required co-factor for such inter-conversion was found to be NAD(H). The ADH gene was engineered for heterologous expression in Escherichia coli, and the enzyme was produced in a soluble form. The recombinant enzyme was purified to homogeneity and physical, spectral and catalytical properties were determined.The findings lead us to propose that BtADH represents a novel primary–secondary alcohol dehydrogenase that acts on primary alcohols of medium chain lengths and simple ketones. Besides, BtADH shares high sequence similarity with well known ADHs from thermophilic origins. Such biochemical characterisation of BtADH provides valuable information for the study of sequence–function relationship including source of thermal stability, cofactor and substrate preferences.  相似文献   

3.
The aldehyde dehydrogenases (ALDHs) are a superfamily of multimeric enzymes which catalyse the oxidation of a broad range of aldehydes into their corresponding carboxylic acids with the reduction of their cofactor, NAD or NADP, into NADH or NADPH. At present, the only known structures concern NAD-dependent ALDHs. Three structures are available in the Protein Data Bank: two are tetrameric and the other is a dimer. We solved by molecular replacement the first structure of an NADP-dependent ALDH isolated from Streptococcus mutans, in its apo form and holo form in complex with NADP, at 1.8 and 2.6 A resolution, respectively. Although the protein sequence shares only approximately 30 % identity with the other solved tetrameric ALDHs, the structures are very similar. However, a large local conformational change in the region surrounding the 2' phosphate group of the adenosine moiety is observed when the enzyme binds NADP, in contrast to the NAD-dependent ALDHs.Structure and sequence analyses reveal several properties. A small number of residues seem to determine the oligomeric state. Likewise, the nature (charge and volume) of the residue at position 180 (Thr in ALDH from S. mutans) determines the cofactor specificity in comparison with the structures of NAD-dependent ALDHs. The presence of a hydrogen bond network around the cofactor not only allows it to bind to the enzyme but also directs the side-chains in a correct orientation for the catalytic reaction to take place. Moreover, a specific part of this network appears to be important in substrate binding. Since the enzyme oxidises the same substrate, glyceraldehyde-3-phosphate (G3P), as NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenases (GAPDH), the active site of GAPDH was compared with that of the S. mutans ALDH. It was found that Arg103, Arg283 and Asp440 might be key residues for substrate binding.  相似文献   

4.
We found N-methyl-L-amino acid dehydrogenase activity in various bacterial strains, such as Pseudomonas putida and Bacillus alvei, and cloned the gene from P. putida ATCC12633 into Escherichia coli. The enzyme purified to homogeneity from recombinant E. coli catalyzed the NADPH-dependent formation of N-alkyl-L-amino acids from the corresponding alpha-oxo acids (e.g. pyruvate, phenylpyruvate, and hydroxypyruvate) and alkylamines (e.g. methylamine, ethylamine, and propylamine). Ammonia was inert as a substrate, and the enzyme was clearly distinct from conventional NAD(P)-dependent amino acid dehydrogenases, such as alanine dehydrogenase (EC 1.4.1.1). NADPH was more than 300 times more efficient than NADH as a hydrogen donor in the enzymatic reductive amination. Primary structure analysis revealed that the enzyme belongs to a new NAD(P)-dependent oxidoreductase superfamily, the members of which show no sequence homology to conventional NAD(P)-dependent amino acid dehydrogenases and opine dehydrogenases.  相似文献   

5.
We purified to homogeneity an enzyme from Citrobacter sp. strain KCTC 18061P capable of decolorizing triphenylmethane dyes. The native form of the enzyme was identified as a homodimer with a subunit molecular mass of about 31 kDa. It catalyzes the NADH-dependent reduction of triphenylmethane dyes, with remarkable substrate specificity related to dye structure. Maximal enzyme activity occurred at pH 9.0 and 60°C. The enzymatic reaction product of the triphenylmethane dye crystal violet was identified as its leuco form by UV-visible spectral changes and thin-layer chromatography. A gene encoding this enzyme was isolated based on its N-terminal and internal amino acid sequences. The nucleotide sequence of the gene has a single open reading frame encoding 287 amino acids with a predicted molecular mass of 30,954 Da. Although the deduced amino acid sequence displays 99% identity to the hypothetical protein from Listeria monocytogenes strain 4b H7858, it shows no overall functional similarity to any known protein in the public databases. At the N terminus, the amino acid sequence has high homology to sequences of NAD(P)H-dependent enzymes containing the dinucleotide-binding motif GXXGXXG. The enzyme was heterologously expressed in Escherichia coli, and the purified recombinant enzyme showed characteristics similar to those of the native enzyme. This is the first report of a triphenylmethane reductase characterized from any organism.  相似文献   

6.
The locking-on strategy uses soluble analogues of the enzymes specific substrate to produce biospecific adsorption of individual NAD(P)(+)-dependent dehydrogenases on immobilized NAD(P)(+) derivatives, which is so selective that a single enzyme activity can be purified from crude cellular extracts in a single chromatographic step with yields approaching 100%. However, attempts to further develop and apply this strategy to the biospecific chromatographic purification of a range of NAD(P)(+)-dependent dehydrogenases revealed some anomalous chromatographic behavior and certain unexplained phenomenon. Much of this can be attributed to nonbiospecific interference effects. Identification and elimination of this interference is discussed in the present study focusing on bovine liver glutamate dehydrogenase (GDH; EC 1.4.1.3) as the "test" enzyme. Results further confirm the potential of the locking-on strategy for the rapid purification of NAD(P)(+)-dependent dehydrogenases and provide further insight into the parameters which should be considered during the development of a truly biospecific affinity chromatographic system based on the locking-on strategy. The kinetic mechanism of bovine liver GDH has been the topic of much controversy with some reports advocating a sequential ordered mechanism of substrate binding and others reporting a sequential random mechanism. Since the kinetic locking-on strategy is dependent on the target NAD(P)(+)-dependent dehydrogenase having an ordered sequential mechanism of substrate binding, the bioaffinity chromatographic behavior of bovine liver GDH using the locking-on tactic suggests that this enzyme has an ordered sequential mechanism of substrate binding under a variety of experimental conditions when NAD(+) is used as cofactor.  相似文献   

7.
Lactate dehydrogenase (D-lactate:NAD+ oxidoreductase, EC 1.1.1.28) from the horseshoe crab, Limulus polyphemus, a dimeric enzyme stereospecific for D-lactate, has been purified by affinity chromatography. Maleyl tryptic peptides containing arginine residues isolated from the Limulus enzyme have been characterized and sequenced. The small peptides obtained from similarly treated L-lactate-specific enzyme homologs define major portions of the substrate and coenzyme binding regions and are virtually identical among L-lactate-specific enzymes. Although the six small peptides and free arginine isolated from the Limulus enzyme indicate that the small number of arginine tryptic peptides are located in a few discrete consecutive clusters similarly to the L-lactate dehydrogenases, the peptides nevertheless show no obvious sequence homology to the corresponding peptides from L-lactate dehydrogenases. These results indicate that this lactate dehydrogenase of altered substrate specificity either evolved with major rearrangements of the active site if it evolved from an L-lactate dehydrogenase, or that D-lactate dehydrogenases have evolved from a different protein. The results contradict proposed models which suggest that minor changes in the spatial orientation of pyruvate resulting from minimal rearrangement of the active site could accommodate the change in substrate specificity.  相似文献   

8.

Genome sequence of the hyperthermophilic archaeon Pyrobaculum calidifontis contains an open reading frame, Pcal_0632, annotated as glyceraldehyde-3-phosphate dehydrogenase, which is partially overlapped with phosphoglycerate kinase. In the phylogenetic tree, Pcal_0632 clustered with phosphorylating glyceraldehyde-3-phosphate dehydrogenases characterized from hyperthermophilic archaea and exhibited highest identity of 54% with glyceraldehyde-3-phosphate dehydrogenase from Sulfolobus tokodaii. To examine biochemical function of the protein, Pcal_0632 gene was expressed in Escherichia coli and the gene product was purified. The recombinant enzyme catalyzed the conversion of glyceraldehyde 3-phosphate and inorganic phosphate into 1,3-bisphosphoglycerate utilizing both NAD and NADP as cofactor with a marked preference for NADP. The enzyme was highly stable against temperature and denaturants. Half-life of the enzyme was 60 min at 100 °C. It retained more than 60% of its activity even after an incubation of 72 h at room temperature in the presence of 6 M urea. High thermostability and resistance against denaturants make Pcal_0632 a novel glyceraldehyde-3-phosphate dehydrogenase.

  相似文献   

9.
Five different immobilized NAD+ derivatives were employed to compare the behavior of four amino acid dehydrogenases chromatographed using kinetic-based enzyme capture strategies (KBECS): S6-, N6-, N1-, 8'-azo-, and pyrophosphate-linked immobilized NAD+. The amino acid dehydrogenases were NAD+-dependent phenylalanine (EC 1.4.1.20), alanine (EC 1.4.1.1), and leucine (EC 1.4.1.9) dehydrogenases from various microbial species and NAD(P)+-dependent glutamate dehydrogenase from bovine liver (GDH; EC 1.4.1.3). KBECS for bovine heart L-lactate dehydrogenase (EC 1.1.1.27) and yeast alcohol dehydrogenase (EC 1.1.1.1) were also applied to assist in a preliminary assessment of the immobilized cofactor derivatives. Results confirm that the majority of the enzymes studied retained affinity for NAD+ immobilized through an N6 linkage, as opposed to an N1 linkage, replacement of the nitrogen with sulfur to produce an S6 linkage, or attachment of the cofactor through the C8 position or the pyrophosphate group of the cofactor. The one exception to this was the dual-cofactor-specific GDH from bovine liver, which showed no affinity for N6-linked NAD+ but was biospecifically adsorbed to S6-linked NAD+ derivatives in the presence of its soluble KBEC ligand. The molecular basis for this is discussed together with the implications for future development and application of KBECS.  相似文献   

10.
NAD(P)+-dependent alcohol dehydrogenases (ADH) are widely distributed in all phyla. These proteins can be assigned to three nonhomologous groups of isozymes, with group III being highly diverse with regards to catalytic activity and primary structure. Members of group III ADHs share a conserved stretch of amino acid residues important for cofactor binding and metal ion coordination, while sequence identities for complete proteins are highly diverse (<20 to >90 %). A putative group III ADH PaYqhD has been identified in BLAST analysis from the plant pathogenic enterobacterium Pectobacterium atrosepticum. The PaYqhD gene was expressed in the heterologous host Escherichia coli, and the recombinant protein was purified in a two-step purification procedure to homogeneity indicating an obligate dimerization of monomers. Four conserved amino acid residues involved in metal ion coordination were substituted with alanine, and their importance for catalytic activity was confirmed by circular dichroism spectrum determination, in vitro, and growth experiments. PaYqhD exhibits optimal activity at 40 °C with short carbon chain aldehyde compounds and NADPH as cofactor indicating the enzyme to be an aldehyde reductase. No oxidative activities towards alcoholic compounds were detectable. EDTA completely inhibited catalytic activity and was fully restored by the addition of Co2+. Activity measurements together with sequence alignments and structure analysis confirmed that PaYqhD belongs to the butanol dehydrogenase-like enzymes within group III of ADHs.  相似文献   

11.
The involvement of two primary alcohol dehydrogenases, BDH and BOH, in butane utilization in Pseudomonas butanovora (ATCC 43655) was demonstrated. The genes coding for BOH and BDH were isolated and characterized. The deduced amino acid sequence of BOH suggests a 67-kDa alcohol dehydrogenase containing pyrroloquinoline quinone (PQQ) as cofactor and in the periplasm (29-residue leader sequence). The deduced amino acid sequence of BDH is consistent with a 70.9-kDa, soluble, periplasmic (37-residue leader sequence) alcohol dehydrogenase containing PQQ and heme c as cofactors. BOH and BDH mRNAs were induced whenever the cell's 1-butanol oxidation activity was induced. When induced with butane, the gene for BOH was expressed earlier than the gene for BDH. Insertional disruption of bdh or boh affected adversely, but did not eliminate, butane utilization by P. butanovora. The P. butanovora mutant with both genes boh and bdh inactivated was unable to grow on butane or 1-butanol. These cells, when grown in citrate and incubated in butane, developed butane oxidation capability and accumulated 1-butanol. The enzyme activity of BOH was characterized in cell extracts of the P. butanovora strain with bdh disrupted. Unlike BDH, BOH oxidized 2-butanol. The results support the involvement of two distinct NAD(+)-independent, PQQ-containing alcohol dehydrogenases, BOH (a quinoprotein) and BDH (a quinohemoprotein), in the butane oxidation pathway of P. butanovora.  相似文献   

12.
Kho R  Baker BL  Newman JV  Jack RM  Sem DS  Villar HO  Hansen MR 《Proteins》2003,50(4):589-599
A novel method to organize protein structural information based solely on sequence is presented. The method clusters proteins into families that correlate with the three-dimensional protein structure and the conformation of the bound ligands. This procedure was applied to nicotinamide adenine dinucleotide [NAD(P)]-utilizing enzymes to identify a total of 94 sequence families, 53 of which are structurally characterized. Each of the structurally characterized proteins within a sequence family correlates to a single protein fold and to a common bound conformation of NAD(P). A wide range of structural folds is identified that recognize NAD(P), including Rossmann folds and beta/alpha barrels. The defined sequence families can be used to identify the type and prevalence of NAD(P)-utilizing enzymes in the proteomes of sequenced organisms. The proteome of Mycobacterium tuberculosis was mined to generate a proteome-wide profile of NAD(P)-utilizing enzymes coded by this organism. This enzyme family comprises approximately 6% of the open reading frames, with the largest subgroup being the Rossmann fold, short-chain dehydrogenases. The preponderance of short-chain dehydrogenases correlates strongly with the phenotype of M. tuberculosis, which is characterized as having one of the most complex prokaryotic cell walls.  相似文献   

13.
A new enzyme, L-tartrate decarboxylase, was found in cells of Pseudomonas sp. group Ve-2. The enzyme was purified to homogeneity and characterized. The enzyme requires K+, Mg2+, and NAD+ for L-tartrate decarboxylation. The dependence of the enzymatic decarboxylation on NAD+ suggests that the decarboxylation involves redox reactions of the substrate. The enzyme catalyzes NAD(+)-linked oxidative decarboxylation of D-malate as well. The enzyme is composed of four subunits with identical molecular weight (Mr 40,000). The apparent Michaelis constants for L-tartrate and NAD+ are 1.1 mM, respectively. The cofactor requirements and the physical properties of the enzyme were similar to those of L-tartrate dehydrogenase-D-malate dehydrogenase from Rhodopseudomonas sphaeroides, and tartrate dehydrogenase from P. putida.  相似文献   

14.
D C Crans  S M Schelble 《Biochemistry》1990,29(28):6698-6706
Vanadate dimer and tetramer inhibit glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides. The inhibition by a vanadate mixture containing vanadate monomer, dimer, tetramer, and pentamer was determined by measuring the rates of glucose 6-phosphate oxidation and reduction of NAD (or NADP) catalyzed by glucose-6-phosphate dehydrogenase. The inhibition by vanadate is competitive with respect to NAD or NADP and noncompetitive (a mixed type) with respect to glucose 6-phosphate (G6P) when NAD or NADP are cofactors. This inhibition pattern varies from that observed with phosphate and thus suggests vanadate interacts differently than a phosphate analogue with the enzyme. 51V NMR spectroscopy was used to directly correlate the inhibition of vanadate solutions to the vanadate dimer and/or tetramer, respectively. The activity of the vanadate oligomer varied depending on the cofactor and which substrate was being varied. The vanadate dimer was the major inhibiting species with respect to NADP. This is in contrast to the vanadate tetramer, which was the major inhibiting species with respect to G6P and with respect to NAD. The inhibition by vanadate when G6P was varied was weak. The competitive inhibition pattern with respect to NAD and NADP suggests the possibility that vanadate oligomers may also inhibit catalysis of other NAD- or NADP-requiring dehydrogenases. Significant concentrations of vanadate dimer and tetramer are only found at fairly high vanadate concentrations, so these species are not likely to represent vanadium species present under normal physiological conditions. It is however possible the vanadate dimer and/or tetramer represent toxic vanadate species.  相似文献   

15.
《BBA》1987,893(3):386-397
Three NAD(P)H dehydrogenases were found and purified from a soluble fraction of cells of the purple non-sulfur bacterium Rhodobacter capsulatus, strain B10. Molecular mass of NAD(P)H, NADPH and NADH dehydrogenases are 67 000 (4 · 18 000), 35 000 and 39 000, and the isoelectric points are 4.6, 4.3 and 4.5, respectively. NAD(P)H dehydrogenase is characterized by a higher sensitivity to quinacrine, NADPH dehydrogenase by its sensitivity to p-chloromercuribenzoate and NADH dehydrogenase by its sensitivity to sodium arsenite. In contrast to the other two enzymes, NAD(P)H dehydrogenase is capable of oxidizing NADPH as well as NADH, but the ratio of their oxidation rates depends on the pH. All NAD(P)H dehydrogenases reacted with ferricyanide, 2,6-dichlorophenolindophenol, benzoquinone and naphthoquinone, but did not exhibit transhydrogenase, reductase or oxidase activity. Moreover, NADH dehydrogenase was also capable of reducing FAD and FMN. NAD(P)H and NADH dehydrogenases possessed cytochrome-c reductase activity, which was stimulated by menadione and ubiquinone Q1. The activity of NAD(P)H and NADH dehydrogenases depended on culture-growth conditions. The activity of NAD(P)H dehydrogenase from cells grown under chemoheterotrophic aerobic conditions was the lowest and it increased notably under photoheterotrophic anaerobic conditions upon lactate or malate growth limitation. The activity of NADH dehydrogenase was higher from the cells grown under photoheterotrophic anaerobic conditions upon nitrate growth limitation and under chemoheterotrophic aerobic conditions. NADPH dehydrogenase synthesis dependence on R. capsulatus growth conditions was insignificant.  相似文献   

16.
A PCR protocol was used to identify and sequence a gene encoding a DNA ligase from Thermococcus fumicolans (Tfu). The recombinant enzyme, expressed in Escherichia coli BL21(DE3) pLysS, was purified to homogeneity and characterized. The optimum temperature and pH of Tfu DNA ligase were 65 degrees C and 7.0, respectively. The optimum concentration of MgCl2, which is indispensable for the enzyme activity, was 2 mM. We showed that Tfu DNA ligase displayed nick joining and blunt-end ligation activity using either ATP or NAD+, as a cofactor. In addition, our results would suggest that Tfu DNA ligase is likely to use the same catalytic residues with the two cofactors. The ability for DNA ligases, to use either ATP or NAD+, as a cofactor, appears to be specific of DNA ligases from Thermococcales, an order of hyperthermophilic microorganisms that belongs to the euryarchaeotal branch of the archaea domain.  相似文献   

17.
One of the most critical events in the origins of cellular life was the development of lipid membranes. Archaea use isoprenoid chains linked via ether bonds to sn-glycerol 1-phosphate (G1P), whereas bacteria and eukaryotes use fatty acids attached via ester bonds to enantiomeric sn-glycerol 3-phosphate. NAD(P)H-dependent G1P dehydrogenase (G1PDH) forms G1P and has been proposed to have played a crucial role in the speciation of the Archaea. We present here, to our knowledge, the first structures of archaeal G1PDH from the hyperthermophilic methanogen Methanocaldococcus jannaschii with bound substrate dihydroxyacetone phosphate, product G1P, NADPH, and Zn2+ cofactor. We also biochemically characterized the enzyme with respect to pH optimum, cation specificity, and kinetic parameters for dihydroxyacetone phosphate and NAD(P)H. The structures provide key evidence for the reaction mechanism in the stereospecific addition for the NAD(P)H-based pro-R hydrogen transfer and the coordination of the Zn2+ cofactor during catalysis. Structure-based phylogenetic analyses also provide insight into the origins of G1PDH.  相似文献   

18.
Reher M  Schönheit P 《FEBS letters》2006,580(5):1198-1204
Cells of Picrophilus torridus, grown on glucose, contained all enzyme activities of a non-phosphorylative Entner-Doudoroff pathway, including glucose dehydrogenase, gluconate dehydratase, 2-keto-3-deoxygluconate aldolase, glyceraldehyde dehydrogenase (GADH), glycerate kinase (2-phosphoglycerate forming), enolase and pyruvate kinase. GADH was purified to homogeneity. The 115-kDa homodimeric protein catalyzed the oxidation of glyceraldehyde with NADP+ at highest catalytic efficiency. NAD+ was not used. By MALDI-TOF analysis, open reading frame (ORF) Pto0332 was identified in the genome of P. torridus as the encoding gene, designated gadh, and the recombinant GADH was characterized. In Thermoplasma acidophilum ORF Ta0809 represents a gadh homolog with highest sequence identity; the gene was expressed and the recombinant protein was characterized as functional GADH with properties very similar to the P. torridus enzyme. Sequence comparison and phylogenetic analysis define both GADHs as members of novel enzyme family within the aldehyde dehydrogenase superfamily.  相似文献   

19.

Background

The high costs of pyridine nucleotide cofactors have limited the applications of NAD(P)-dependent oxidoreductases on an industrial scale. Although NAD(P)H regeneration systems have been widely studied, NAD(P)+ regeneration, which is required in reactions where the oxidized form of the cofactor is used, has been less well explored, particularly in whole-cell biocatalytic processes.

Methodology/Principal Findings

Simultaneous overexpression of an NAD+ dependent enzyme and an NAD+ regenerating enzyme (H2O producing NADH oxidase from Lactobacillus brevis) in a whole-cell biocatalyst was studied for application in the NAD+-dependent oxidation system. The whole-cell biocatalyst with (2R,3R)-2,3-butanediol dehydrogenase as the catalyzing enzyme was used to produce (3R)-acetoin, (3S)-acetoin and (2S,3S)-2,3-butanediol.

Conclusions/Significance

A recombinant strain, in which an NAD+ regeneration enzyme was coexpressed, displayed significantly higher biocatalytic efficiency in terms of the production of chiral acetoin and (2S,3S)-2,3-butanediol. The application of this coexpression system to the production of other chiral chemicals could be extended by using different NAD(P)-dependent dehydrogenases that require NAD(P)+ for catalysis.  相似文献   

20.
Aldehyde dehydrogenase from Pseudomonas testosteroni was purified to homogeneity. The enzyme has a pH optimum of 8.2, uses a wide range of aldehydes as substrates and cationic dyes (Wurster's blue, phenazine methosulphate and thionine), but not anionic dyes (ferricyanide and 2.6-dichloroindophenol), NAD(P)+ or O2, as electron acceptors. Haem c and pyrroloquinoline quinone appeared to be absent but the common cofactors of molybdenum hydroxylases were present. Xanthine was not a substrate and allopurinol was not an inhibitor. Alcohols were inhibitors only when turnover of the enzyme occurred in aldehyde conversion. The enzyme has a relative molecular mass of 186,000, consists of two subunits of equal size (Mr 92,000), and 1 enzyme molecule contains 1 FAD, 1 molybdopterin cofactor, 4 Fe and 4 S. It is a novel type of NAD(P)+-independent aldehyde dehydrogenase since its catalytic and physicochemical properties are quite different from those reported for already known aldehyde-converting enzymes like haemoprotein aldehyde dehydrogenase (EC 1.2.99.3), quino-protein alcohol dehydrogenases (EC 1.1.99.8) and molybdenum hydroxylases.  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