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 31 毫秒
1.
Escherichia coli are capable of growing anaerobically on L-rhamnose as a sole source of carbon and energy and without any exogenous hydrogen acceptor. When grown under such condition, synthesis of a nicotinamide adenine dinucleotide-linked L-lactaldehydepropanediol oxidoreductase is induced. The functioning of this enzyme results in the regeneration of nicotinamide adenine dinucleotide. The enzyme was purified to electrophoretic homogeneity. It has a molecular weight of 76,000, with two subunits that are indistinguishable by electrophoretic mobility. The enzyme reduces L-lactaldehyde to L-1,2-propanediol with reduced nicotinamide adenine dinucleotide as a cofactor. The Km were 0.035 mM L-lactaldehyde and 1.25 mM L-1,2-propanediol, at pH 7.0 and 9.5, respectively. The enzyme acts only on the L-isomers. Strong substrate inhibition was observed with L-1,2-propanediol (above 25 mM) in the dehydrogenase reaction. The enzyme has a pH optimum of 6.5 for the reduction of L-lactaldehyde and of 9.5 for the dehydrogenation of L-1,2-propanediol. The enzyme is, according to the parameters presented in this report, indistinguishable from the propanediol oxidoreductase induced by anaerobic growth on fucose.  相似文献   

2.
NAD-dependent 1,2-propanediol dehydrogenase (EC 1.1.1.4) activity was detected in cell-free crude extracts of various propane-grown bacteria. The enzyme activity was much lower in 1-propanol-grown cells than in propane-grown cells of Pseudomonas fluorescens NRRL B-1244, indicating that the enzyme may be inducible by metabolites of propane subterminal oxidation. 1,2-Propanediol dehydrogenase was purified from propane-grown Ps. fluorescens NRRL B-1244. The purified enzyme fraction shows a single-protein band upon acrylamide gel electrophoresis and has a molecular weight of 760,000. It consists of 10 subunits of identical molecular weight (77,600). It oxidizes diols that possess either two adjacent hydroxy groups, or a hydroxy group with an adjacent carbonyl group. Primary and secondary alcohols are not oxidized. The pH and temperature optima for 1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propanediol and NAD are 2 X 10(-2) and 9 X 10(-5) M, respectively. The 1,2-propanediol dehydrogenase activity was inhibited by strong thiol reagents, but not by metal-chelating agents. The amino acid composition of the purified enzyme was determined. Antisera prepared against purified 1,2-propanediol dehydrogenase from propane-grown Ps. fluorescens NRRL B-1244 formed homologous precipitin bands with isofunctional enzymes derived from propane-grown Arthrobacter sp. NRRL B-11315, Nocardia paraffinica ATCC 21198, and Mycobacterium sp. P2y, but not from propane-grown Pseudomonas multivorans ATCC 17616 and Brevibacterium sp. ATCC 14649, or 1-propanol-grown Ps. fluorescens NRRL B-1244. Isofunctional enzymes derived from methane-grown methylotrophs also showed different immunological and catalytic properties.  相似文献   

3.
Methanol-utilizing bacteria, Klebsiella sp. No. 101 and Microcyclus eburneus could grow aerobically and statically on 1,2-propanediol. The authors examined the presence of enzyme activity of adenosyl-B12 dependent diol dehydratase as well as NAD dependent diol dehydroagenase. Adenosyl-B12 dependent diol dehydratase activity was not detected in these organisms, even if these grown statically.

The dehydrogenase activity was found in the extract from these methanol-utilizing bacteria cells grown on various carbon sources. The partially purified enzyme preparation from the cells of Mic. eburneus grown aerobically on 1,2-propanediol dehydrogenated 1,2-propanediol, 1,2-butanediol and 2,3-butanediol. The enzyme activity was separated into two fractions, namely enzyme I and II on DEAE-Sephadex A-25 column chromatography. The enzyme I was different from the enzyme II in the ratio of enzyme activity to 1,2-propanediol and 2,3-butanediol, heat stability, pH stability and pH optimum, and effect of 2-mercaptoethanol.  相似文献   

4.
Salmonella enterica forms polyhedral organelles involved in coenzyme B(12)-dependent 1,2-propanediol degradation. These organelles are thought to consist of a proteinaceous shell that encases coenzyme B(12)-dependent diol dehydratase and perhaps other enzymes involved in 1,2-propanediol degradation. The function of these organelles is unknown, and no detailed studies of their structure have been reported. Genes needed for organelle formation and for 1,2-propanediol degradation are located at the 1,2-propanediol utilization (pdu) locus, but the specific genes involved in organelle formation have not been identified. Here, we show that the pduA gene encodes a shell protein required for the formation of polyhedral organelles involved in coenzyme B(12)-dependent 1,2-propanediol degradation. A His(6)-PduA fusion protein was purified from a recombinant Escherichia coli strain and used for the preparation of polyclonal antibodies. The anti-PduA antibodies obtained were partially purified by a subtraction procedure and used to demonstrate that the PduA protein localized to the shell of the polyhedral organelles. In addition, electron microscopy studies established that strains with nonpolar pduA mutations were unable to form organelles. These results show that the pduA gene is essential for organelle formation and indicate that the PduA protein is a structural component of the shell of these organelles. Physiological studies of nonpolar pduA mutants were also conducted. Such mutants grew similarly to the wild-type strain at low concentrations of 1,2-propanediol but exhibited a period of interrupted growth in the presence of higher concentrations of this growth substrate. Growth tests also showed that a nonpolar pduA deletion mutant grew faster than the wild-type strain at low vitamin B(12) concentrations. These results suggest that the polyhedral organelles formed by S. enterica during growth on 1,2-propanediol are not involved in the concentration of 1,2-propanediol or coenzyme B(12), but are consistent with the hypothesis that these organelles moderate aldehyde production to minimize toxicity.  相似文献   

5.
A bacterium designated Nocardia A60 was isolated for its capacity to utilize propylene oxide (1,2-epoxypropame) aerobically as a carbon and energy source for growth. Extracts of cells grown on the epoxide catalyzed the conversion of propylene oxide to 1,2-propanediol This epoxidase activity was absent in cells grown on 1,2-propanediol or succinate. During growth of the organism on propylene oxide and 1,2-propanediol it contained high levels of diol dehydratase (EC 4.2.1.28). Enhanced levels of propionyl-CoA carboxylase during growth on propylene oxide and 1,2-propanediol suggest that these compounds are metabolized via propionate and succinate.  相似文献   

6.
Klebsiella pneumoniae (Aerobacter aerogenes) ATCC 8724 was able to grow anaerobically on 1,2-propanediol and 1,2-ethanediol as carbon and energy sources. Whole cells of the bacterium grown anaerobically on 1,2-propanediol or on glycerol catalyzed conversion of 1,2-diols and aldehydes to the corresponding acids and alcohols. Glucose-grown cells also converted aldehydes, but not 1,2-diols, to acids and alcohols. The presence of activities of coenzyme B(12)-dependent diol dehydratase, alcohol dehydrogenase, coenzyme-A-dependent aldehyde dehydrogenase, phosphotransacetylase, and acetate kinase was demonstrated with crude extracts of 1,2-propanediol-grown cells. The dependence of the levels of these enzymes on growth substrates, together with cofactor requirements in in vitro conversion of these substrates, indicates that 1,2-diols are fermented to the corresponding acids and alcohols via aldehydes, acyl-coenzyme A, and acyl phosphates. This metabolic pathway for 1,2-diol fermentation was also suggested in some other genera of Enterobacteriaceae which were able to grow anaerobically on 1,2-propanediol. When the bacteria were cultivated in a 1,2-propanediol medium not supplemented with cobalt ion, the coenzyme B(12)-dependent conversion of 1,2-diols to aldehydes was the rate-limiting step in this fermentation. This was because the intracellular concentration of coenzyme B(12) was very low in the cells grown in cobalt-deficient medium, since the apoprotein of diol dehydratase was markedly induced in the cells grown in the 1,2-propanediol medium. Better cell yields were obtained when the bacteria were grown anaerobically on 1,2-propanediol. Evidence is presented that aerobically grown cells have a different metabolic pathway for utilizing 1,2-propanediol.  相似文献   

7.
An enzyme exhibiting alpha-L-rhamnosidase activity was purified by fractionating a culture filtrate of Aspergillus nidulans grown on L-rhamnose as the sole carbon source. The alpha-L-rhamnosidase was shown to be N-glycosylated and had a molecular mass of 102 kDa, of which approximately 7% was contributed by carbohydrate. The enzyme, optimally active at pH 4.5-6 and 60 degrees C, had an isoelectric point of 5. With rho-nitrophenyl-alpha-L-rhamnopyranoside as the substrate it showed Km and Vmax values of 0.27 mmol l-1 and 64.6 U mg-1, respectively. The enzyme was competitively inhibited by L-rhamnose (Ki 0.3 mmol l-1). Ca2+ (2 mmol l-1) stimulated the activity of the enzyme by 14%, whereas Mg2+ (2 mmol l-1) inhibited it by 63%. Substrate specificity studies showed the alpha-L-rhamnosidase to be active both on alpha-1,2 and alpha-1,6 linkages to beta-D-glucosides.  相似文献   

8.
Glycerol has become an attractive carbon source in the biotechnology industry owing to its low price and reduced state. However, glycerol is rarely used as a carbon source in Saccharomyces cerevisiae because of its low utilization rate. In this study, we used glycerol as a main carbon source in S. cerevisiae to produce 1,2-propanediol. Metabolically engineered S. cerevisiae strains with overexpression of glycerol dissimilation pathway genes, including glycerol kinase (GUT1), glycerol 3-phosphate dehydrogenase (GUT2), glycerol dehydrogenase (gdh), and a glycerol transporter gene (GUP1), showed increased glycerol utilization and growth rate. More significant improvement of glycerol utilization and growth rate was accomplished by introducing 1,2-propanediol pathway genes, mgs (methylglyoxal synthase) and gldA (glycerol dehydrogenase) from Escherichia coli. By engineering both glycerol dissimilation and 1,2-propanediol pathways, the glycerol utilization and growth rate were improved 141% and 77%, respectively, and a 2.19 g 1,2- propanediol/l titer was achieved in 1% (v/v) glycerolcontaining YEPD medium in engineered S. cerevisiae.  相似文献   

9.
Batch and continuous cultures of a newly isolated Clostridium butyricum strain were carried out on industrial glycerol, the major by-product of the bio-diesel production process. For both types of cultures, the conversion yield obtained was around 0.55 g of 1,3-propanediol formed per 1 g of glycerol consumed whereas the highest 1,3-propanediol concentration, achieved during the single-stage continuous cultures was 35-48 g l-1. Moreover, the strain presented a strong tolerance at the inhibitory effect of the 1,3-propanediol, even at high concentrations of this substance at the chemostat (e.g. 80 g l-1). 1,3-Propanediol was associated with cell growth whereas acetate and butyrate seemed non growth-associated products. At low and medium dilution rates (until 0.1 h-1), butyrate production was favoured, whereas at higher rates acetate production increased. The maximum 1,3-propanediol volumetric productivity obtained was 5.5 g l-1 h-1. A two-stage continuous fermentation was also carried out. The first stage presented high 1,3-propanediol volumetric productivity, whereas the second stage (with a lower dilution rate) served to further increase the final product concentration. High 1,3-propanediol concentrations were achieved (41-46 g l-1), with a maximum volumetric productivity of 3.4 g l-1 h-1. A cell concentration decrease was reported between the second and the first fermentor.  相似文献   

10.
Synechocystis sp. PCC 6803 PG is a cyanobacterial strain capable of synthesizing 1,2-propanediol from carbon dioxide (CO2) via a heterologous three-step pathway and a methylglyoxal synthase (MGS) originating from Escherichia coli as an initial enzyme. The production window is restricted to the late growth and stationary phase and is apparently coupled to glycogen turnover. To understand the underlying principle of the carbon partitioning between the Calvin-Benson-Bassham (CBB) cycle and glycogen in the context of 1,2-propanediol production, experiments utilizing 13C labeled CO2 have been conducted. Carbon fluxes and partitioning between biomass, storage compounds, and product have been monitored under permanent illumination as well as under dark conditions. About one-quarter of the carbon incorporated into 1,2-propanediol originated from glycogen, while the rest was derived from CO2 fixed in the CBB cycle during product formation. Furthermore, 1,2-propanediol synthesis was depending on the availability of photosynthetic active radiation and glycogen catabolism. We postulate that the regulation of the MGS from E. coli conflicts with the heterologous reactions leading to 1,2-propanediol in Synechocystis sp. PCC 6803 PG. Additionally, homology comparison of the genomic sequence to genes encoding for the methylglyoxal bypass in E. coli suggested the existence of such a pathway also in Synechocystis sp. PCC 6803. These findings are critical for all heterologous pathways coupled to the CBB cycle intermediate dihydroxyacetone phosphate via a MGS and reveal possible engineering targets for rational strain optimization.  相似文献   

11.
Salmonella enterica degrades 1,2-propanediol by a pathway dependent on coenzyme B12 (adenosylcobalamin [AdoCb1]). Previous studies showed that 1,2-propanediol utilization (pdu) genes include those for the conversion of inactive cobalamins, such as vitamin B12, to AdoCbl. However, the specific genes involved were not identified. Here we show that the pduO gene encodes a protein with ATP:cob(I)alamin adenosyltransferase activity. The main role of this protein is apparently the conversion of inactive cobalamins to AdoCbl for 1,2-propanediol degradation. Genetic tests showed that the function of the pduO gene was partially replaced by the cobA gene (a known ATP:corrinoid adenosyltransferase) but that optimal growth of S. enterica on 1,2-propanediol required a functional pduO gene. Growth studies showed that cobA pduO double mutants were unable to grow on 1,2-propanediol minimal medium supplemented with vitamin B(12) but were capable of growth on similar medium supplemented with AdoCbl. The pduO gene was cloned into a T7 expression vector. The PduO protein was overexpressed, partially purified, and, using an improved assay procedure, shown to have cob(I)alamin adenosyltransferase activity. Analysis of the genomic context of genes encoding PduO and related proteins indicated that particular adenosyltransferases tend to be specialized for particular AdoCbl-dependent enzymes or for the de novo synthesis of AdoCbl. Such analyses also indicated that PduO is a bifunctional enzyme. The possibility that genes of unknown function proximal to adenosyltransferase homologues represent previously unidentified AdoCbl-dependent enzymes is discussed.  相似文献   

12.
1,2-Propanediol (propylene glycol) is an existing commodity chemical and can be produced from renewable resources using microbes. By virtue of being a natural product, relevant biochemical pathways can be harnessed into fermentation processes to produce 1,2-propanediol. In the present review, the chemical process and different biological strategies for the production of 1,2-propanediol are reviewed and compared with the potentials and limitations of all processes. For the successful commercial production of this diol, it is necessary to establish the metabolic pathways and production hosts (microorganisms), which are capable of delivering final product with high yields and volumetric productivity. Three pathways which have been recognized for 1,2-propanediol production are discussed here. In the first, de-oxy sugars like fucose and rhamnose are used as the carbon sources, while in the other route, the glycolytic intermediate-dihydroxyacetonephosphate (DHAP) is used to produce 1,2-propanediol via the formation of methylglyoxal. A new pathway of 1,2-propanediol production by lactic acid degradation under anoxic conditions and the enzymes involved is also discussed. The production of this diol has gained attention because of their newer applications in industries such as polymers, food, pharmaceuticals, textiles, etc. Furthermore, improvement in fermentation technology will permit its uses in other applications. Future prospect in the light of the current research and its potential as a major bulk chemical are discussed.  相似文献   

13.
Wild-type strains of Escherichia coli are unable to use L-1,2-propanediol as a carbon and energy source. A series of mutants, able to grow on this compound at progressively faster rates, had been isolated by repeated transfers to a medium containing 20 mM L-1,2-propanediol. These strains synthesize at high constitutive levels a propanediolmicotinamide adenine dinucleotide oxidoreductase, an enzyme serving as a lactaldehyde during L-fucose fermentation by wild type cells. In this study, a mutant that can grow rapidly on the novel carbon source was subjected to further selection in a medium containing L-1,2-propanediol never exceeding 0.5 mM to obtain a derivative that has an increased power to extract the substrate from the medium. The emerging mutant exhibited four changes at the enzymatic level: (i) fuculose 1-phosphate aldolase activity is lost; (ii) the constitutive propanediol oxidoreductase activity is increased in its level; (iii) lactaldehyde dehydrogenase becomes constitutive and shows an elevated specific activity in crude extracts; and (iv) at low concentrations of propanediol, the facilitated diffusion across the cell membrane is enhanced. Changes two to four seem to act in concert in the trapping of propanediol by hastening its rate of entry and conversion to an ionized metabolite, lactate.  相似文献   

14.
Y M Chen  Y Zhu    E C Lin 《Journal of bacteriology》1987,169(7):3289-3294
Mutant analysis revealed that complete utilization of L-fucose and L-rhamnose by Escherichia coli requires the activity of a common NAD-linked aldehyde dehydrogenase which converts L-lactaldehyde to L-lactate. Mutations affecting this activity mapped to the ald locus at min 31, well apart from the fuc genes (min 60) encoding the trunk pathway for L-fucose dissimilation (as well as L-1,2-propanediol oxidoreductase) and the rha genes (min 88) encoding the trunk pathway for L-rhamnose dissimilation. Mutants that grow on L-1,2-propanediol as a carbon and energy source also depend on the ald gene product for the conversion of L-lactaldehyde to L-lactate.  相似文献   

15.
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   

16.
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   

17.
Combining ion-exchange (AG MP-1) and reversed-phase (C-18) partition chromatography accomplishes a higher degree of purification of NADH than either method can provide alone. Final elution in 95% ethanol, dehydration with anhydrous sodium sulfate, and storage in dry 1,2-propanediol over molecular sieves prevents hydrolysis of the purified dinucleotide.  相似文献   

18.
C zerkawski J.W., P iatkova M. & B reckenridge , G. 1984. Microbial metabolism of 1,2-propanediol studied by the Rumen. Simulation Technique (Rusitec). Journal of Applied Bacteriology , 56 , 81–94.
A series of experiments with the Rumen Simulation Technique (Rusitec) showed that 1,2-propanediol was metabolized efficiently by rumen micro-organisms and that the main end-products of fermentation were propionic and 2-methylbutyric acids. 'Propionaldehyde and n-propanol were also formed as intermediate compounds.' The effect of the diol on digestion of the basal diet appeared to be small with concentrate, or when the roughage was supplemented with additional nitrogen (urea). The decrease in the output of acetic and butyric acids was consistent with utilization of C2 units for synthesis of 2-methylbutyric acid. The fermentation of 1,2-propanediol resulted in little or no increase in the output of additional microbial matter. The distribution of radioactivity from [1-14C] 1,2-propanediol confirmed that propionaldehyde and n -propanol were the primary products of metabolism of the diol and that the end-products were propionic and 2-methylbutyric acids, with very little labelling of microbial matter. Between 2% and 3% of radioactivity was found in gases and surprisingly the specific radioactivity of methane was higher than that of carbon dioxide, particularly during the initial stages of incubation. Possible pathways in the degradation of 1,2-propanediol by rumen micro-organisms are suggested and discussed in relation to similar reactions established in other systems.  相似文献   

19.
The properties of D-1-amino-2-propanol oxidoreductase from wild-type Escherichia coli have been compared with those of a glycerol dehydrogenase from mutant E. coli 424 and of a 1,2-propanediol oxidoreductase from Neisseria gonorrhoeae. Several independent lines of evidence indicate that the former two enzymes are identical. (i) Both enzymatic activities purified to virtual homogeneity in an identical manner, and the ratio of specific activities (glycerol/aminopropanol) remained constant at all stages. (ii) When electrophoresed, both purified enzymes showed a major as well as a minor band of protein coincident with activity, and these two bands from each enzyme had the same mobility. (iii) The subunit molecular weights and isoelectric points were identical for each enzyme, and (iv) kinetic constants (Km and Vmax values) determined with three different substrates were the same. The somewhat greater stability of the glycerol dehydrogenase to controlled heat denaturation at 74 degrees C was the only difference observed between these two enzymes. In contrast, D-1-amino-2-propanol oxidoreductase was found to be immunochemically and kinetically distinct from the 1,2-propanediol oxidoreductase from N. gonorrhoeae.  相似文献   

20.
An enzyme capable of dehalogenating vicinal haloalcohols to their corresponding epoxides was purified from the 3-chloro-1,2-propanediol-utilizing bacterium Arthrobacter sp. strain AD2. The inducible haloalcohol dehalogenase converted 1,3-dichloro-2-propanol, 3-chloro-1,2-propanediol, 1-chloro-2-propanol, and their brominated analogs, 2-bromoethanol, as well as chloroacetone and 1,3-dichloroacetone. The enzyme possessed no activity for epichlorohydrin (3-chloro-1,2-epoxypropane) or 2,3-dichloro-1-propanol. The dehalogenase had a broad pH optimum at about 8.5 and a temperature optimum of 50 degrees C. The enzyme followed Michaelis-Menten kinetics, and the Km values for 1,3-dichloro-2-propanol and 3-chloro-1,2-propanediol were 8.5 and 48 mM, respectively. Chloroacetic acid was a competitive inhibitor, with a Ki of 0.50 mM. A subunit molecular mass of 29 kDa was determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. With gel filtration, a molecular mass of 69 kDa was found, indicating that the native protein is a dimer. The amino acid composition and N-terminal amino acid sequence are given.  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