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.
Rhodococcus sp. strain DTB (DSM 44534) grows on bis(1-chloro-2-propyl) ether (DDE) as sole source of carbon and energy. The non-chlorinated diisopropyl ether and bis(1-hydroxy-2-propyl) ether, however, did not serve as substrates. In ether degradation experiments with dense cell suspensions, 1-chloro-2-propanol and chloroacetone were formed, which indicated that scission of the ether bond is the first step while dehalogenation of the chlorinated C(3)-compounds occurs at a later stage of the degradation pathway. Inhibition of ether scission by methimazole suggested that the first step in degradation is catalyzed by a flavin-dependent enzyme activity. The non-chlorinated compounds 1,2-propanediol, hydroxyacetone, lactate, pyruvate, 1-propanol, propanal, and propionate also supported growth, which suggested that the intermediates 1,2-propanediol and hydroxyacetone are converted to pyruvate or to propionate, which can be channeled into the citric acid cycle by a number of routes. Total release of chloride and growth-yield experiments with DDE and non-chlorinated C(3)-compounds suggested complete biodegradation of the chlorinated ether.  相似文献   

2.
NAD-dependent Thermotoga maritima glycerol dehydrogenase (TmGlyDH) converts glycerol into dihydroxyacetone (DHA), a valuable synthetic precursor and sunless tanning agent. In this work, recombinant TmGlyDH was characterized to determine if it can be used to catalyze DHA production. The pH optima for glycerol oxidation and DHA reduction at 50 °C were 7.9 and 6.0, respectively. Under the conditions tested, TmGlyDH had a linear Arrhenius plot up to 80 °C. TmGlyDH was more thermostable than other glycerol dehydrogenases, remaining over 50 % active after 7 h at 50 °C. TmGlyDH was active on racemic 1,2-propanediol and produced (R)-1,2-propanediol from hydroxyacetone with an enantiomeric excess above 99 %, suggesting that TmGlyDH can also be used for chiral synthesis. (R)-1,2-propanediol production from hydroxyacetone was demonstrated for the first time in a one-enzyme cycling reaction using glycerol as the second substrate. Negative cooperativity was observed with glycerol and DHA, but not with the cofactor. Apparent kinetic parameters for glycerol, DHA, and NAD(H) were determined over a broad pH range. TmGlyDH showed little activity with N6-carboxymethyl-NAD+ (N6-CM-NAD), an NAD+ analog modified for easy immobilization to amino groups, but the double mutation V44A/K157G increased catalytic efficiency with N6-CM-NAD+ ten-fold. Finally, we showed for the first time that a GlyDH is active with immobilized N6-CM-NAD+, suggesting that N6-CM-NAD+ can be immobilized on an electrode to allow TmGlyDH activity in a system that reoxidizes the cofactor electrocatalytically.  相似文献   

3.
Escherichia coli were grown on 14.3% uniformly 13C-labeled glucose as the sole carbon source and challenged anaerobically with 90% 13C-labeled formaldehyde. The major multiply labeled metabolites were identified by 13C NMR spectroscopy to be glycerol and 1,2-propanediol, and a minor metabolite was shown to be 1,3-propanediol. In each case, formaldehyde is incorporated only into the C1 position. A novel form of 13C NMR isotope dilution analysis of the major products reveals that all the 1,2-diol C1 is formaldehyde derived but that about 40% of the glycerol C1 is derived from bacterial sources. Glycerokinase converted the metabolite [1-13C]glycerol to equal amounts of [3-13C]glycerol 3-phosphate and [1-13C]glycerol 3-phosphate, demonstrating that the metabolite is racemic. When [13C]formaldehyde incubation was carried out in H2O/D2O mixtures, deuterium incorporation was detected by beta- and gamma-isotope shifts. The 1,3-diol is deuterium labeled only at C2 and only once, while the 1,2-diol and glycerol are each labeled independently at both C2 and C3; C3 is multiply labeled. Deuterium incorporation levels are different for each metabolite, indicating that the biosynthetic pathways probably diverge early.  相似文献   

4.
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   

5.
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   

6.
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   

7.
P Boutron 《Cryobiology》1992,29(3):347-358
A 2,3-butanediol containing 96.7% (w/w) racemic mixture of the levo and dextro isomers and only 3.1% (w/w) of the meso isomer (called 2,3-butanediol 97% dl) has been used for the cryoprotection of red blood cells. The erythrocytes were cooled to -196 degrees C at rates between 2 and 3500 degrees C/min, followed by slow or rapid warming. Up to 20% (w/w) of this polyalcohol, only the classical peak of survival is observed, as with up to 20% (w/w) 1,2-propanediol or 1,3-butanediol. Twenty percent 2,3-butanediol 97% dl can protect red blood cells very efficiently. The maximum survival, of 90%, as with 20% glycerol, is a little lower than with 20% 1,2-propanediol and higher than with 20% 1,3-butanediol. Fifteen percent 2,3-butanediol protects fewer red blood cells than 15% glycerol or 1,2-propanediol, with a maximum survival of about 80%. The best cryoprotection by 30% 2,3-butanediol 97% dl is obtained at the slowest cooling and warming rates, where survival approaches 90%. After a minimum, an increase of survival is observed at the fastest cooling rates, which would correspond to complete vitrification. These rates are lower than with 30%, 1,2-propanediol or 1,3-butanediol, in agreement with the higher glass-forming tendency of 2,3-butanediol 97% dl solutions. In agreement with the remarkable physical properties of its aqueous solutions, the present experiments also suggest that 2,3-butanediol containing mainly the levo and dextro isomers could be a very useful cryoprotectant for organ cryopreservation. However, it would perhaps be better to use it in combination with other cryoprotectants, since it is a little more toxic than glycerol or 1,2-propanediol at high concentrations.  相似文献   

8.

Background

Conversion of industrial processes to more nature-friendly modes is a crucial subject for achieving sustainable development. Utilization of hydrogen-oxidation reactions by hydrogenase as a driving force of bioprocess reaction can be an environmentally ideal method because the reaction creates no pollutants. We expressed NAD-dependent alcohol dehydrogenase from Kluyveromyces lactis in a hydrogen-oxidizing bacterium: Ralstonia eutropha. This is the first report of hydrogen-driven in vivo coupling reaction of the alcohol dehydrogenase and indigenous soluble NAD-reducing hydrogenase. Asymmetric reduction of hydroxyacetone to (R)-1,2-propanediol, which is a commercial building block for antibacterial agents, was performed using the transformant as the microbial cell catalyst.

Results

The two enzymes coupled in vitro in vials without a marked decrease of reactivity during the 20 hr reaction because of the hydrogenase reaction, which generates no by-product that affects enzymes. Alcohol dehydrogenase was expressed functionally in R. eutropha in an activity level equivalent to that of indigenous NAD-reducing hydrogenase under the hydrogenase promoter. The hydrogen-driven in vivo coupling reaction proceeded only by the transformant cell without exogenous addition of a cofactor. The decrease of reaction velocity at higher concentration of hydroxyacetone was markedly reduced by application of an in vivo coupling system. Production of (R)-1,2-propanediol (99.8% e.e.) reached 67.7 g/l in 76 hr with almost a constant rate using a jar fermenter. The reaction velocity under 10% PH2 was almost equivalent to that under 100% hydrogen, indicating the availability of crude hydrogen gas from various sources. The in vivo coupling system enabled cell-recycling as catalysts.

Conclusions

Asymmetric reduction of hydroxyacetone by a coupling reaction of the two enzymes continued in both in vitro and in vivo systems in the presence of hydrogen. The in vivo reaction system using R. eutropha transformant expressing heterologous alcohol dehydrogenase showed advantages for practical usage relative to the in vitro coupling system. The results suggest a hopeful perspective of the hydrogen-driven bioprocess as an environmentally outstanding method to achieve industrial green innovation. Hydrogen-oxidizing bacteria can be useful hosts for the development of hydrogen-driven microbial cell factories.  相似文献   

9.
Spectrophotometric techniques for determining the activities of lipases, lysophospholipases, and phospholipases are reviewed. These methods involve the use of thioester substrate analogs as well as omega-nitrophenyl derivatives of the corresponding lipids. The most promising results are obtained with the thioester substrate analogs. Mono- and diacylglycerol lipases are assayed by using rac-1-S-decanoyl-1-mercapto-2,3-propanediol and rac-1,2-S,O-didecanoyl-1-mercapto-2,3-propanediol, respectively. Phospholipases A1 and A2 are determined by using rac-1,2-S,O-didecanoyl-3-phosphocholine-1-mercapto-2,3-propanediol and 2-hexadecanoylthio-1-ethyl-phosphocholine, respectively. Lysophospholipases are measured by using 2-hexadecanoylthio-1-ethyl-phosphocholine. Phospholipase C is assayed with rac-1-S-phosphocholine-2,3-O-didecanoyl-1-mercapto-2,3-propanediol. Thioester substrate analog assay procedures are more rapid, sensitive, convenient, continuous, and less expensive than the classical radiochemical techniques.  相似文献   

10.
Fermentation of glycerol to 1,3-propanediol: use of cosubstrates   总被引:16,自引:0,他引:16  
Three fermentable substances, glucose, 1,2-ethanediol and 1,2-propanediol were checked as cosubstrates for the fermentation of glycerol by Clostridium butyricum and Citrobacter freundii with the aim of achieving a complete conversion of glycerol to 1,3-propanediol. Glucose was fermented by C. butyricum mainly to acetate, CO2 and reducing equivalents in the presence of glycerol and contributed markedly to the 1,3-propanediol yield. However, because of relatively slow growth on glucose, complete conversion was not achieved. If the two glycols were used as cosubstrates for glycerol fermentation, the 1,3-propanediol yield did not increase but dimished considerably, as they were converted to more reduced products, i.e. alcohols instead of acids. From 1,2-propanediol 2-propanol was formed in addition to 1-propanol. The ratio of the propanols was dependent on the culture conditions.  相似文献   

11.
In previous studies, we demonstrated that 1,2-propanediol induces shortening and bundling of actin filaments, both in vitro and in vivo, and that it enhances actin/alpha-actinin interaction, especially at low temperature. 1,2-Propanediol also promotes homogeneous microporous networks which can be vitrified by rapid cooling. In the present study, dynamical rheological measurements were performed under various sets of experimental conditions including temperature (4 or 20 degrees C), protein concentrations (actin and alpha-actinin), and 1,2-propanediol presence or absence. Gelation kinetics were monitored, and the resulting actin mechanical properties investigated, in order to untangle the respective effects of the experimental parameters. Whether in the presence or absence of solvent, low temperature brings about a rigidification of the sample, as does high protein concentration, as expected. However, 1,2-propanediol itself involves either softening of the sample (at high temperature and low protein concentration or at low temperature and high protein concentration) or rigidification in the case of low temperature and low protein concentration. These effects result from the competition between actin/alpha-actinin affinity (enhanced by both low temperature and 1,2-propanediol), bundling of filaments (fostered by alpha-actinin for alpha-actinin/actin ratios used), rate of actin polymerization (higher at high temperature), shortening effect of 1,2-propanediol on actin filaments, and chain mobility (lower at high protein concentration). As discussed, only the combination of low temperature and low protein concentration induces full crosslinking of the system into a viscoelastic solid under the influence of 1,2-propanediol.  相似文献   

12.
Yamanishi M  Ide H  Murakami Y  Toraya T 《Biochemistry》2005,44(6):2113-2118
The reaction catalyzed by adenosylcobalamin-dependent diol dehydratase proceeds by a radical mechanism. A radical pair consisting of the Co(II) of cob(II)alamin and an organic radical intermediate formed during catalysis gives EPR spectra. The high-field doublet and the low-field broad signals arise from the weak interaction of an organic radical with the low-spin Co(II) of cob(II)alamin. To characterize the organic radical intermediate in the diol dehydratase reaction, several deuterated and (13)C-labeled 1,2-propanediols were synthesized, and the EPR spectra observed in the catalysis were measured using them as substrate. The EPR spectra with the substrates deuterated on C1 showed significant line width narrowing of the doublet signal. A distinct change in the hyperfine coupling was seen with [1-(13)C]-1,2-propanediol, but not with the [2-(13)C]-counterpart. Thus, the organic radical intermediate observed by EPR spectroscopy was identified as the 1,2-propanediol-1-yl radical, a C1-centered substrate-derived radical.  相似文献   

13.
This short review covers metabolic pathways, genetics and metabolic engineering of 1,2-propanediol formation in microbes. 1,2-Propanediol production by bacteria and yeasts has been known for many years and two general pathways are recognized. One involves the metabolism of deoxyhexoses, where lactaldehyde is formed during the glycolytic reactions and is then reduced to 1,2-propanediol. The second pathway derives from the formation of methylglyoxal from dihydroxyacetonephosphate and its subsequent reduction to 1,2-propanediol. The enzymes involved in the reduction of methylglyoxal can generate isomers of lactaldehyde or acetol, which can be further reduced by specific reductases, giving chiral 1,2-propanediol as the product. The stereospecificity of the enzymes catalyzing the two reduction steps is important in deriving a complete pathway. Through genetic engineering, appropriate combinations of enzymes have been brought together in Escherichia coli and yeast to generate 1,2-propanediol from glucose. The optimization of these strains may yield microbial processes for the production of this widely used chemical. Received: 25 May 2000 / Received revision: 24 July 2000 / Accepted: 25 July 2000  相似文献   

14.
The substrate specificity of adenosylcobalamin-dependent diol dehydrase was further studied in detail using an enzyme preparation that appears homogeneous by ultracentrifugal and gel electrophoretical criteria. Besides 1,2-propanediol and 1,2-ethanediol, glycerol, 1,2- and 2,3-butanediol were found to serve as substrate for the enzyme, whereas 1,3-propanediol was not. Of the substrate analogs tested, glycerol displayed some striking features: it was dehydrated to β-hydroxypropionaldehyde with concomitant inactivation of the enzyme. Although the initial velocity with glycerol was comparable to that with 1,2-propanediol, the dehydration reaction ceased almost completely within 3 min accompanying rapid, irreversible inactivation of the holoenzyme. 1,2- and 2,3-Butanediol were converted to butyraldehyde and methyl ethyl ketone, respectively, at a rate much lower than that with 1,2-propanediol. 2,3-Butanediol is the only compound, other than 1,2-diols, known at present to show a considerable substrate activity.  相似文献   

15.
All the aqueous solutions of linear saturated polyalcohols with four carbons have been investigated at low temperature. Only ice has been observed in the solutions of 1,3-butanediol and 1,2,3- and 1,2,4-butanetriol. For same solute concentration, the glass-forming tendency on cooling is highest with 2,3-butanediol, where it is comparable to that with 1,2-propanediol, the best solute reported to date. However, the quantity of ice and hydrate crystallized is particularly high on slow cooling or on subsequent rewarming. The highest stability of the amorphous state is observed on rewarming the 1,2-butanediol and 1,3-butanediol solutions. With respect to this property, these compounds come just after 1,2-propanediol and before all the other compounds studied so far. They are followed by dimethylsulfoxide and 1,2,3-butanetriol. The glass-forming tendency of the 1,3-butanediol solutions is also very high; it is third only to that of 1,2-propanediol and 2,3-butanediol. The glass-forming tendency is a little smaller with 1,2-butanediol, but it is cubic instead of ordinary hexagonal ice which crystallizes on cooling rapidly with 35% 1,2-butanediol. Cubic ice is thought to be innocuous. A gigantic glass transition is observed with 45% of this strange solute. 1,4-Butanediol, 45% also favors cubic ice greatly. Therefore, 1,2- and 1,3-butanediol with comparable physical properties are perhaps as interesting as 1,2-propanediol for cryopreservation of cells or organs by complete vitrification. Together with 1,2-propanediol, 1,2- and 1,3-butanetriol, 1,2,3-butanetriol, and perhaps 2,3-butanediol provide an interesting battery of solutions for cryopreservation by vitrification.  相似文献   

16.
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   

17.
Abstract In a mineral medium containing sulfate as terminal electron acceptor, the sulfate-reducing bacterium Desulfovibrio alcoholovorans oxidized stoichiometrically 1 mol glycerol to 1 mol acetate and 1 mol 1,3-propanediol to 1 mol acetate with the concomitant reduction of 0.75 and 1 mol sulfate, respectively; 1 mol 1,2-propanediol was degraded to 0.8 mol acetate and 0.1 mol proprionate, with the reduction of approximately 1 mol sulfate. The maximum specific growth rates (μmax in h−1) were 0.22, 0.086 and 0.09 with glycerol, 1,3-propanediol and 1,2-propanediol, respectively. The growth yields were 12.7 g, 11.1 g and 7.2 g dry weight/mol 1,3-propanediol, glycerol and 1,2-propanediol degraded, respectively. The growth yields and maximum specific growth rates of the H2-transferring associations were also calculated. In the absense of sulfate, all these reduced substrates were degraded to acids and methane when D. alcoholovorans was cocultured with Methanospirillum hungatei . Changes in the metabolic pathway were observed in the degradation of 1,2- and 1,3-propanediol. The metabolic efficiency of D. alcoholovorans to degrade glycerol, 1.2- and 1,3-propanediol is discussed.  相似文献   

18.
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   

19.
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   

20.
Cell physiology and metabolic flux distribution of Klebsiella pneumoniae under anaerobic, micro-aerobic and sufficient aerobic conditions were compared. Comparing with the anaerobic condition, the carbon flux flowed from glycerol to biomass increased 10.1% and 389.9%, while the flux flowed to 1,3-propanediol decreased 10.3% and 92.9% under micro-aerobic and sufficient aerobic conditions, respectively. Furthermore, the carbon flux flowed to TCA cycle increased 5.9% and 31.0% under such two conditions. The energy analysis results revealed that the oxygen was favorable for the NADH2 synthesis, but excessive oxygen was disadvantage for the NADH2 utilization in 1,3-propanediol synthesis process. So, the aeration control is significant for the aerobic 1,3-propanediol fermentation. This work is considered helpful for the further understanding of the glycerol metabolism by Klebsiella pneumoniae under aerobic condition and to establish a rational aeration control strategy for 1,3-propanediol aerobic fermentation in a large-scale bioreactor.  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