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 364 毫秒
1.
The sulfate-reducing bacterimDesulfovibrio strain HDv (DSM 6830) grew faster on (S)- and on (R, S)-1,2-propanediol (μmax 0.053 h) than on (R)-propanediol (0.017 h−1) and ethanol (0.027 h−1). From (R, S)-1,2-propanediol-grown cells, an alcohol dehydrogenase was purified. The enzyme was oxygen-labile, NAD-dependent, and decameric; the subunit mol. mass was 48 kDa. The N-terminal amino acid sequence indicated similarity to alcohol dehydrogenases belonging to family III of NAD-dependent alcohol dehydrogenases, the first 21 N-terminal amino acids being identical to those of theDesulfovibrio gigas alcohol dehydrogenase. Best substrates were ethanol and propanol (K m of 0.48 and 0.33 mM, respectively). (R, S)-1,2-Propanediol was a relatively poor substrate for the enzyme, but activities in cell extracts were high enough to account for the growth rate. The enzyme showed a preference for (S)-1,2-propanediol over (R)-1,2-propanediol. Antibodies raised against the alcohol dehydrogenase ofD. gigas showed cross-reactivity with the alcohol dehydrogenase ofDesulfovibrio strain HDv and with cell extracts of six other ethanol-grown sulfate-reducing bacteria.  相似文献   

2.
Acetobacterium woodii utilizes the Wood-Ljungdahl pathway for reductive synthesis of acetate from carbon dioxide. However, A. woodii can also perform non-acetogenic growth on 1,2-propanediol (1,2-PD) where instead of acetate, equal amounts of propionate and propanol are produced as metabolic end products. Metabolism of 1,2-PD occurs via encapsulated metabolic enzymes within large proteinaceous bodies called bacterial microcompartments. While the genome of A. woodii harbours 11 genes encoding putative alcohol dehydrogenases, the BMC-encapsulated propanol-generating alcohol dehydrogenase remains unidentified. Here, we show that Adh4 of A. woodii is the alcohol dehydrogenase required for propanol/ethanol formation within these microcompartments. It catalyses the NADH-dependent reduction of propionaldehyde or acetaldehyde to propanol or ethanol and primarily functions to recycle NADH within the BMC. Removal of adh4 gene from the A. woodii genome resulted in slow growth on 1,2-PD and the mutant displayed reduced propanol and enhanced propionate formation as a metabolic end product. In sum, the data suggest that Adh4 is responsible for propanol formation within the BMC and is involved in redox balancing in the acetogen, A. woodii.  相似文献   

3.
1,3-Propanediol oxidoreductase encoded by dhaT gene, a gene of 1,3-propanediol regulon, is important in converting glycerol to 1,3-propanediol in Klebsiella pneumoniae. DhaT gene was amplified from the genome of K. pneumoniae, sequenced and its amino acid sequence deduced. A predicted secondary structure and 3D-structural model was constructed by homology modelling. Based on these results, we infer that 1,3-propanediol oxidoreductase belongs to NAD(P)-dependent alcohol dehydrogenase group III of iron-activated dehydrogenases.  相似文献   

4.

Purpose of work

A pair of NAD+- and NADP+-dependent group III-alcohol dehydrogenases was characterized from the enterobacterium, Dickeya zeae, to expand our understanding of the distribution and biochemical properties of this interesting group of enzymes. Two putative group III-alcohol dehydrogenases (ADHs) were identified in the genome of Dickeya zeae. Amino acid alignments and phylogenetic analysis revealed that Adh3.1 and Adh3.2 are only distantly related (~25 % identity at the protein level). Both proteins were purified to homogeneity after heterologous expression in E. coli. A specific activity of 1.8 U/mg was measured for the NAD+-dependent enzyme Adh3.1 with ethanol used as substrate, while NADPH-dependent Adh3.2 preferred butanal (29.1 U/mg) as substrate. Maximum activity for Adh3.1 was at 50 °C and pH 10 and for Adh3.2 at 70 °C and pH 6. Cell viability assays were used to confirm activity towards butanal and glyoxals. Biochemical characterization and phylogenetic analyses led to the hypothesis that Adh3.1 and Adh3.2 are probably the result of an ancient gene duplication event followed by functional diversification.  相似文献   

5.
Oh BR  Seo JW  Heo SY  Hong WK  Luo LH  Joe MH  Park DH  Kim CH 《Bioresource technology》2011,102(4):3918-3922
A mutant strain of Klebsiella pneumoniae, termed GEM167, was obtained by γ irradiation, in which glycerol metabolism was dramatically affected on exposure to γ rays. Levels of metabolites of the glycerol reductive pathway, 1,3-propanediol (1,3-PD) and 3-hydroxypropionic acid (3-HP), were decreased in the GEM167 strain compared to a control strain, whereas the levels of metabolites derived from the oxidative pathway, 2,3-butanediol (2,3-BD), ethanol, lactate, and succinate, were increased. Notably, ethanol production from glycerol was greatly enhanced upon fermentation by the mutant strain, to a maximum production level of 21.5 g/l, with a productivity of 0.93 g/l/h. Ethanol production level was further improved to 25.0 g/l upon overexpression of Zymomonas mobilispdc and adhII genes encoding pyruvate decarboxylase (Pdc) and aldehyde dehydrogenase (Adh), respectively in the mutant strain GEM167.  相似文献   

6.
Plant gene products that have been described as `alcohol dehydrogenases' are surveyed and related to their CPGN nomenclature. Most are Zn-dependent medium chain dehydrogenases, including `classical' alcohol dehydrogenase (Adh1), glutathione-dependent formaldehyde dehydrogenase (Fdh1), cinnamyl alcohol dehydrogenase (Cad2), and benzyl alcohol dehydrogenase (Bad1). Plant gene products belonging to the short-chain dehydrogenase class should not be called alcohol dehydrogenases unless such activity is shown.  相似文献   

7.
Klebsiella pneumoniae is a nosocomial pathogen frequently isolated from opportunistic infections, especially in clinical environments. In spite of its potential pathogenicity, this microorganism has several metabolic potentials that could be used in biotechnology applications. K. pneumoniae is able to metabolize glycerol as a sole source of carbon and energy. 1,3-Propanediol dehydrogenase is the core of the metabolic pathway for the use of glycerol. We have determined the crystallographic structure of 1,3-propanediol dehydrogenase, a type III Fe-NAD-dependent alcohol dehydrogenase, at 2.7-Å resolution. The structure of the enzyme monomer is closely related to that of other alcohol dehydrogenases. The overall arrangement of the enzyme showed a decameric structure, formed by a pentamer of dimers, which is the catalytic form of the enzyme. Dimers are associated by strong ionic interactions that are responsible for the highly stable in vivo packing of the enzyme. Kinetic properties of the enzyme as determined in the article would suggest that this decameric arrangement is related to the cooperativity between monomers.  相似文献   

8.
In the cofermentation of glycerol with a sugar by Lactobacillus brevis and Lactobacillus buchneri, a 1,3-propanediol:NAD+ oxidoreductase provides an additional method of NADH disposal. The enzyme has been purified from both L. brevis B22 and L. buchneri B190 and found to have properties very similar to those reported for the enzyme from Klebsiella pneumoniae. The enzymes required Mn2+ and are probably octamers with a molecular mass of 350 kDa. Although not absolutely specific for 1,3-propanediol when tested as dehydrogenases, the enzymes have less than 10% activity with glycerol, ethanol, and 1,2-propanediol. These properties contrast sharply with those of a protein isolated from another Lactobacillus species (L. reuteri) that ferments glycerol with glucose and previously designated a 1,3-propanediol dehydrogenase.  相似文献   

9.
In the cofermentation of glycerol with a sugar by Lactobacillus brevis and Lactobacillus buchneri, a 1,3-propanediol:NAD+ oxidoreductase provides an additional method of NADH disposal. The enzyme has been purified from both L. brevis B22 and L. buchneri B190 and found to have properties very similar to those reported for the enzyme from Klebsiella pneumoniae. The enzymes required Mn2+ and are probably octamers with a molecular mass of 350 kDa. Although not absolutely specific for 1,3-propanediol when tested as dehydrogenases, the enzymes have less than 10% activity with glycerol, ethanol, and 1,2-propanediol. These properties contrast sharply with those of a protein isolated from another Lactobacillus species (L. reuteri) that ferments glycerol with glucose and previously designated a 1,3-propanediol dehydrogenase.  相似文献   

10.
The degradation of 1,2-dichloroethane and 2-chloroethanol by Xanthobacter autotrophicus GJ10 proceeds via chloroacetaldehyde, a reactive and potentially toxic intermediate. The organism produced at least three different aldehyde dehydrogenases, of which one is plasmid encoded. Two mutants of strain GJ10, designated GJ10M30 and GJ10M41, could no longer grow on 2-chloroethanol and were found to lack the NAD-dependent aldehyde dehydrogenase that is the predominant protein in wild-type cells growing on 2-chloroethanol. Mutant GJ10M30, selected on the basis of its resistance to 1,2-dibromoethane, also had lost haloalkane dehalogenase activity and Hg2+ resistance, indicating plasmid loss. From a gene bank of strain GJ10, different clones that complemented one of these mutants were isolated. In both transconjugants, the aldehyde dehydrogenase that was absent in the mutants was overexpressed. The enzyme was purified and was a tetrameric protein of 55-kDa subunits. The substrate range was rather broad, with the highest activity measured for acetaldehyde. The Km value for chloroacetaldehyde was 160 μM, higher than those for other aldehydes tested. It is concluded that the ability of GJ10 to grow with 2-chloroethanol is due to the high expression level of an aldehyde dehydrogenase with a rather low activity for chloroacetaldehyde.  相似文献   

11.
Because of its key role in the metabolism of glycerol during fermentation, 1,3-propanediol dehydrogenase (EC 1.1.1.202) of Enterobacter agglomerans CNCM 1210 was purified to homogeneity and studied with respect to its sensitivity to pH and to nucleotide and 1,3-propanediol concentrations. Enzyme activity was optimal at pH 7.8. The enzyme was competitively inhibited by NAD+ (Ki of 0.29 mM), and 1,3-propanediol exerted a strong inhibitory effect according to a mixed-type inhibition with a Ki of 13.7 mM and an a-factor of 9.0. It is proposed that these dehydrogenase properties be extended to the dehydrogenases of Citrobacter freundii and Klebsiella pneumoniae, which exhibited numerous similar physical properties. Received: 4 December 1996 / Accepted: 24 March 1997  相似文献   

12.
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   

13.
Esterases are widely used in food processing industry, but there is little information concerning enzymes involved in decompositions of esters contributing to pollution of environment. Vinyl acetate (an ester of vinyl alcohol and acetic acid) is a representative of volatile organic compounds (VOCs) in decomposition, of which hydrolyses and oxidoreductases are mainly involved. Their activities under periodically changing conditions of environment are essential for the removal of dangerous VOCs. Esterase and alcohol/aldehyde dehydrogenase activities were determined in crude cell extract from Pseudomonas fluorescens PMC 2123 after vinyl acetate induction. All examined enzymes exhibit their highest activity at 30–35 °C and pH 7.0–7.5. Esterase preferably hydrolyzed ester bonds with short fatty chains without plain differences for C2 or C4. Comparison of Km values for alcohol and aldehyde dehydrogenases for acetaldehyde suggested that this metabolite was preferentially oxidized than reduced. Activity of alcohol dehydrogenase reducing acetaldehyde to ethanol suggested that one mechanism of defense against the elevated concentration of toxic acetaldehyde could be its temporary reduction to ethanol. Esterase activity was inhibited by phenylmethanesulfonyl fluoride, while β-mercaptoethanol, dithiothreitol, and ethylenediaminetetraacetic acid had no inhibitor effect. From among metal ions, only Mg2+ and Fe2+ stimulated the cleavage of ester bond.  相似文献   

14.
Two groups of metabolically related enzymes, the Group III family of Fe2+-dependent alcohol dehydrogenases (ADHs) and the separate subfamily of nucleoside diphosphates linked to x (nudix) hydrolases that activate Group III ADHs are under-characterized. Here we report the steady-state initial-velocity forward direction (alcohol → aldehyde) reaction of a Group III ADH, namely gamma-hydroxybutyrate dehydrogenase (GHBDH, UniProt: Q59104), cloned from Cupriavidus necator as a fusion protein. We also report the effects of nudix hydrolases on the GHBDH reaction. At optimal pH 9.0, the GHBDH reaction is activated ~2-fold by two different saturating purified nudix hydrolases, namely Bacillus methanolicus activator (ACT, UniProt: I3EA59) and Escherichia coli NudF (UniProt Q93K97) proteins. At physiological pH values of ~7.0, ACT activates by >3.5-fold. Initial-rate characterization at pH 9.0 of the forward direction un-activated and ACT-activated reactions show for both cases competitive inhibition by the product succinic semialdehyde versus GHB, and noncompetitive inhibitions by the three other substrate-product combinations. This pattern is consistent with NAD+ binding first in Mono-Iso Theorell-Chance kinetics. Mutants of some possibly important residues in GHBDH also were characterized. H265, conserved among all Group III ADHs and previously proposed to be a critical general base, is only ~4-fold helpful for GHBDH activity relevant to H265A. The four previously proposed conserved Fe2+ chelators (D193, H197, H261 and H280) each are essential for GHBDH activity. A 2-step explanation for cross-species stimulation by sub-stoichiometric ACT in the forward direction and confirmed lack of ACT stimulation in the reverse direction reaction is proposed.  相似文献   

15.
Biochemical studies indicate that alcohol dehydrogenase (ADH) metabolizes retinol to retinal, and that aldehyde dehydrogenase (ALDH) metabolizes retinal to retinoic acid, a molecule essential for growth and development. Summarized herein are several genetic studies supporting in vivo functions for ADH and ALDH in retinoic acid synthesis. Gene targeting was used to create knockout mice for either Adh1 or Adh4. Both knockout mice were viable and fertile without obvious defects. However, when wild-type and Adh4 knockout mice were subjected to vitamin A deficiency during gestation, the survival rate at birth was 3.3-fold lower for Adh4 knockout mice. When adult mice were examined for production of retinoic acid following retinol administration, Adh1 knockout mice exhibited 10-fold lower retinoic acid levels in liver compared with wild-type, whereas Adh4 knockout mice differed from wild-type by less than 2-fold. Thus, Adh1 plays a major role in the metabolism of a large dose of retinol to retinoic acid in adults, whereas Adh4 plays a role in maintaining sufficient retinol metabolism for development during retinol deficiency. ALDHs were examined by overexpression studies in frog embryos. Injection of mRNAs for either mouse Raldh1 or Raldh2 stimulated retinoic acid synthesis in frog embryos at the blastula stage when retinoic acid is normally undetectable. Overexpression of human ALDH2, human ALDH3, and mouse Aldh-pb did not stimulate retinoic acid production. In addition, Raldh2 knockout mice exhibit embryonic lethality with defects in retinoid-dependent tissues. Overall, these studies provide genetic evidence that Adh1, Adh4, Raldh1, and Raldh2 encode retinoid dehydrogenases involved in retinoic acid synthesis in vivo.  相似文献   

16.
The alcohol dehydrogenase from horse liver is able to catalyze the oxidation of a number of 1,2-diols and α-aminoalcohols enantioselectively to l-α-hydroxyaldehydes and l-α-amino aldehydes. A decrease of enantioselectivity was found in reactions with 1,3-diols and substrates with hydrophobic substituent at position 3. α-Aminoalcohols are not substrates for yeast alcohol dehydrogenase, but the enzyme can catalyze the oxidation of most of the diols to l-hydroxyaldehydes. New methods for determination of the optical purity of α-hydroxy-and α-aminoaldehydes via converting them in situ to the corresponding acids, catalyzed by the aldehyde dehydrogenase from yeast, have been developed. The coupled alcohol dehydrogenase/aldehyde dehydrogenase has been extended to preparatory scale synthesis of optically pure l-α-hydroxyacids in the presence of a cofactor regeneration system. The active-site cubic-space section model has been shown not to be applicable to all substrates.  相似文献   

17.
Evidence is presented for the existence in Pseudomonas putida of two NAD-linked dehydrogenases that function sequentially to oxidize benzyl alcohol. Induction of muconate lactonizing enzyme, a 3-oxoadipate pathway enzyme, indicated that P. putida oxidized benzyl alcohol to benzoate. Polyacrylamide gel electrophoresis with activity staining and enzymatic assays for an NAD-dependent dehydrogenase both showed that cells contained a single, constitutive alcohol dehydrogenase capable of oxidizing benzyl alcohol. This enzyme was shown to have the same specificity in extracts of glucose-grown as in benzy alcoholgrown cells. An NAD-aldehyde dehydrogenase oxidized benzaldehyde but was most active with normal alkyl aldehydes. This aldehyde dehydrogenase was shown to be induced, by enzymatic assays and by activity staining of polyacrylamide gel electropherograms, not only in cells grown on benzyl alcohol, but also in cells grown on ethanol. These experiments suggested that the aldehyde dehydrogenase was induced by the alcohol being oxidized rather than the substrate aldehyde.In sum, the evidence from enzyme assays and polyacrylamide gel electrophoresis of extracts indicates that Pseudomonas putida catabolizes benzyl alcohol slowly when it is the sole carbon and energy source, by the action of a constitutive, nonspecific, alcohol dehydrogenase and an alcohol-induced, nonspecific aldehyde dehydrogenase to yield benzoate, which is further metabolized via the 3-oxoadipate (beta-ketoadipate) pathway.In memory of R. Y. Stanier  相似文献   

18.
Rapid immobilization with the one-pot purification of galactitol dehydrogenase (GatDH) and formate dehydrogenase (FDH) is achieved by using iminodiacetic acid (IDA) with chelated Co2+ modified magnetic nanoparticles as a carrier. Lactate dehydrogenase (LDH) from recombinant Escherichia coli and FDH commencing Candida methylica were used as an auxiliary enzyme for the regeneration of NADH/NAD+ with a representative synthesis of (S)-1,2-propanediol and l-tagatose starting from hydroxyacetone and galactitol. The affinity magnetic nanoparticles were characterized by scanning electron microscopy (SEM) and Fourier transform infrared spectroscopy (FTIR), while the purity of GatDH and FDH was assayed by SDS-PAGE analysis. The immobilized two-enzyme system, reflecting the pH dependence of its constituent enzymes, showed optimal activity at pH 7 and 8 for (S)-1,2-propanediol and l-tagatose production, respectively. The immobilized enzyme system retained up to 70% of its activity after one week of repeated use. The use of affinity magnetic nanoparticles offers the advantage of a one-pot purification of His(6)-tagged GatDH and FDH followed by the production of rare sugar and chiral diol.  相似文献   

19.
Electrophoretic and activity variants have been observed for stomach and testis aldehyde dehydrogenases, respectively, among inbred strains of the house mouse (Mus musculus). Genetic evidence was obtained for two new loci encoding these isozymes (designated Ahd-4 and Ahd-6, respectively, for the stomach and testis isozymes) which segregated independently of a number of mouse gene markers, including Ahd-1 (encoding mitochondrial aldehyde dehydrogenase) on chromosome 4, ep (pale ears), a marker for chromosome 19, on which Ahd-2 (encoding liver cytosolic aldehyde dehydrogenase) has been previously localized, and Adh-3 (encoding the stomach-specific isozyme of alcohol dehydrogenase) on chromosome 3. Recombination studies have indicated, however, that Ahd-4 and Ahd-6 are distinct but closely linked loci on the mouse genome. An extensive survey of the distribution of Ahd-1, Ahd-2, Ahd-4, and Ahd-6 alleles among 56 strains of mice is reported. No variants have been observed, so far, for the microsomal (AHD-3) and mitochondrial/cytosolic (AHD-5) isozymes previously described. This study, in combination with previous investigations on mouse aldehyde dehydrogenases, provides evidence for six genetic loci for this enzyme.  相似文献   

20.
Clostridium acetobutylicum is not able to grow on glycerol as the sole carbon source since it cannot reoxidize the excess of NADH generated by glycerol catabolism. Nevertheless, when the pSPD5 plasmid, carrying the NADH-consuming 1,3-propanediol pathway from C. butyricum VPI 3266, was introduced into C. acetobutylicum DG1, growth on glycerol was achieved, and 1,3-propanediol was produced. In order to compare the physiological behavior of the recombinant C. acetobutylicum DG1(pSPD5) strain with that of the natural 1,3-propanediol producer C. butyricum VPI 3266, both strains were grown in chemostat cultures with glycerol as the sole carbon source. The same “global behavior” was observed for both strains: 1,3-propanediol was the main fermentation product, and the qH2 flux was very low. However, when looking at key intracellular enzyme levels, significant differences were observed. Firstly, the pathway for glycerol oxidation was different: C. butyricum uses a glycerol dehydrogenase and a dihydroxyacetone kinase, while C. acetobutylicum uses a glycerol kinase and a glycerol-3-phosphate dehydrogenase. Secondly, the electron flow is differentially regulated: (i) in C. butyricum VPI 3266, the in vitro hydrogenase activity is 10-fold lower than that in C. acetobutylicum DG1(pSPD5), and (ii) while the ferredoxin-NAD+ reductase activity is high and the NADH-ferredoxin reductase activity is low in C. acetobutylicum DG1(pSPD5), the reverse is observed for C. butyricum VPI 3266. Thirdly, lactate dehydrogenase activity is only detected in the C. acetobutylicum DG1(pSPD5) culture, explaining why this microorganism produces lactate.  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