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 46 毫秒
1.
When Corynebacterium glutamicum is grown with a sufficient nitrogen supply, urea crosses the cytoplasmic membrane by passive diffusion. A permeability coefficient for urea diffusion of 9 × 10–7 cm s–1 was determined. Under conditions of nitrogen starvation, an energy-dependent urea uptake system was synthesized. Carrier-mediated urea transport was catalyzed by a secondary transport system linked with proton motive force. With a K m for urea of 9 μM, the affinity of this uptake system was much higher than the affinity of urease towards its substrate (K m approximately 55 mM urea). The maximum uptake velocity depended on the expression level and was relatively low [2–3.5 nmol min–1 (mg dry wt.)–1]. Received: 11 August 1997 / Accepted: 2 December 1997  相似文献   

2.
Transmembrane threonine fluxes (i.e., uptake, diffusion, and carrier-mediated excretion) all contribut-ing to threonine production by a recombinant strain of Corynebacterium glutamicum, were analyzed and quantitated. A threonine-uptake carrier that transports threonine in symport with sodium ions was identified. Under production conditions (i.e., when internal threonine is high), this uptake system catalyzed predominantly threonine/threonine exchange. Threonine export via the uptake system was excluded. Threonine efflux from the cells was shown to comprise both carrier-mediated excretion and passive diffusion. The latter process was analyzed after inhibition of all carrier-mediated fluxes. Threonine diffusion was found to proceed with a first-order rate constant of 0.003 min–1 or 0.004 μl min–1 (mg dry wt.)–1, which corresponds to a permeability of 8 × 10–10 cm s–1. According to this permeability, less than 10% of the efflux observed under optimal conditions takes place via diffusion, and more than 90% must result from the activity of the excretion carrier. In addition, the excretion carrier was identified by (1) inhibition of its activity by amino acid modifying reagents and (2) its dependence on metabolic energy in the form of the membrane potential. Activity of the excretion system depended on the membrane potential, but not on the presence of sodium ions. Threonine export in antiport against protons is proposed. Received: 25 August 1995 / Accepted: 18 October 1995  相似文献   

3.
Corynebacterium glutamicum took up glutamine by a sodium-dependent secondary transport system. Both the membrane potential and the sodium gradient were driving forces. Glutamine uptake showed Michaelis-Menten kinetics, with aK m of 36 μM and aV max of 12.5 nmol min−1 (mg dry weight)−1 at pH 7. Despite a pH optimum in the alkaline range around pH 9, it was shown that uncharged glutamine is the transported species. The affinity for the cotransported sodium was relatively low; an apparentK m of 1.4 mM was determined. Among various substrates tested, only asparagine, when added in 50-fold excess, led to an inhibition of glutamine transport. It was concluded that glutamine uptake occurs via a specific transport system in symport with at least one sodium ion.  相似文献   

4.
Tetrahydrofolate was shown to function as a methyl acceptor in the anaerobic demethylation of dimethylsulfoniopropionate to methylthiopropionate in cell extracts of the sulfate-reducing bacterium strain WN. Dimethylsulfoniopropionate-dependent activities were 0.56 μmol methyltetrahydrofolate min–1 (mg protein)–1 and were higher than required to explain the growth rate of strain WN on dimethylsulfoniopropionate. The reaction did not require ATP or reductive activation by titanium(III)-nitrilotriacetic acid. Preincubation of the extract under air significantly decreased the activity (35% loss in 3 h). Three other dimethylsulfoniopropionate-demethylating sulfate reducers, Desulfobacterium niacini, Desulfobacterium vacuolatum, and Desulfobacterium strain PM4, had dimethylsulfoniopropionate:tetrahydrofolate methyltransferase activities of 0.16, 0.05, and 0.24 μmol min–1 (mg protein)–1, respectively. No methyltransferase activity to tetrahydrofolate was found with betaine as a substrate, not even in extracts of betaine-grown cells of these sulfate reducers. Dimethylsulfoniopropionate demethylation in cell extracts of strain WN was completely inhibited by 0.5 mM propyl iodide; in the light, the inhibition was far less strong, indicating involvement of a corrinoid-dependent methyltransferase. Received: 24 June 1997 / Accepted: 29 August 1997  相似文献   

5.
Growth and magnetite formation in Magnetospirillum gryphiswaldense MSR-1 were found close to the maximum at an extracellular iron concentration of 15–20 μM. Ferrous iron was incorporated by a slow, diffusion-like process. Several iron chelators including various microbial siderophores were unable to promote transport of iron into the cells. In contrast, spent culture fluids stimulated the uptake of ferric iron in iron-depleted cells at a high rate, whereas fresh medium and transport buffer were unable to promote iron uptake. However, no siderophore-like compound could be detected in spent culture fluids by the Chrome Azurol S assay. Ferric iron uptake followed Michaelis-Menten kinetics with a K m of 3 μM and a V max of 0.86 nmol min–1 (mg dry weight)–1, suggesting a comparatively low-affinity, but high-velocity transport system. Iron incorporation was sensitive to 2,4-dinitrophenol and carbonylcyanide-m-chlorophenylhydrazone, indicating an energy-dependent transport process. Received: 21 May 1996 / Accepted: 7 August 1996  相似文献   

6.
Insertion of magnesium into protoporphyrin IX is a complex ATP-dependent reaction catalysed by the enzyme Mg-chelatase. Three separate proteins (Mg-chelatase subunits), designated as D, H and I, are involved in the chelation reaction. The genes encoding the Mg-chelatase subunits of the green sulfur bacterium Chlorobium vibrioforme and of the cyanobacterium Synechocystis strain PCC6803 were expressed in Escherichia coli. The recombinant proteins were purified, tested for ATPase and phosphate exchange activities, and compared with the activities of the corresponding subunits of Rhodobacter sphaeroides. The Synechocystis strain PCC6803 I subunit and the C. vibrioforme H and I subunits hydrolysed ATP at the rates of 2.0, 1.8 and 0.16 nmol (mg protein)–1 min–1, respectively. The ATPase activity of the C. vibrioforme H subunit was similar to that reported for the R. sphaeroides H subunit. The Synechocystis strain PCC6803 H subunit failed to hydrolyse ATP. The I subunit of Synechocystis strain PCC6803 and C. vibrioforme catalysed a transfer of PO4 from ATP to ADP (exchange activity) at the rate of 1.75 ± 0.15 nmol (mg protein)–1 min–1. This exchange rate was 300-fold lower than that reported for the R. sphaeroides I subunit. The PO4 exchange activities were correlated with the presence of the sequence GXRGTGKSTXVRALA in the primary structure of the three I subunits. Mg-chelatase activity was reconstituted by combining the three subunits of the same bacterium [rates of 41–89 pmol Mg-deuteroporphyrin (mg protein)–1 min–1]. Heterologous subunit combinations resulted in low or no Mg-chelatase activity. Received: 25 May 1998 / Revision received: 24 November 1998 / Accepted: 27 November 1998  相似文献   

7.
Hydrogenase was solubilized from the cytoplasmic membrane fraction of betaine-grown Sporomusa sphaeroides, and the enzyme was purified under oxic conditions. The oxygen-sensitive enzyme was partially reactivated under reducing conditions, resulting in a maximal activity of 19.8 μmol H2 oxidized min–1 (mg protein)–1 with benzyl viologen as electron acceptor and an apparent K m value for H2 of 341 μM. The molecular mass of the native protein estimated by native PAGE and gel filtration was 122 and 130 kDa, respectively. SDS-PAGE revealed two polypeptides with molecular masses of 65 and 37 kDa, present in a 1:1 ratio. The native protein contained 15.6 ± 1.7 mol Fe, 11.4 ± 1.4 mol S2–, and 0.6 mol Ni per mol enzyme. The hydrogenase coupled with viologen dyes, but not with other various artificial electron carriers, FAD, FMN, or NAD(P)+. The amino acid sequence of the N-termini of the subunits showed a high degree of similarity to eubacterial membrane-bound uptake hydrogenases. Washed membranes catalyzed a H2-dependent cytochrome b reduction at a rate of 0.18 nmol min–1 (mg protein)–1. Received: 7 September 1995 / Accepted: 4 December 1995  相似文献   

8.
Glutamate excretion due to amino acid starvation was investigated in “stringent” and “relaxed” strains ofEscherichia coli. The observed excretion process isrelA-dependent, carrier-mediated, and glutamate-specific. After induction, excretion was detected within less than 2 min and continued for more than 5h with a rate of 7–10 nmol (mg dry weight)−1 min−1. Using carbonyl cyanidem-chlorophenylhydrazone or polymyxin B nonapeptide, together with valinomycin, it was shown that glutamate excretion is driven by the membrane potential.  相似文献   

9.
Acetate is effectively taken up by whole cells of Corynebacterium glutamicum via a specific carrier with a pH optimum of 8. The K m of acetate uptake was 50 μM and the V max 25–35 nmol/mg dw min. The activation energy was determined to be 70 kJ/mol. Acetate uptake was competitively inhibited by propionate with a K i of about 30 μM and blocked by addition of sulfhydryl reagents. The transport activity was clearly dependent on the membrane potential, but independent of the presence of Na+-ions. It is concluded that uptake of acetate proceeds by a secondary, proton coupled mechanism.  相似文献   

10.
A Corynebacterium glutamicum strain with inactivated pyruvate dehydrogenase complex and a deletion of the gene encoding the pyruvate:quinone oxidoreductase produces about 19 mM l-valine, 28 mM l-alanine and about 55 mM pyruvate from 150 mM glucose. Based on this double mutant C. glutamicumaceEpqo, we engineered C. glutamicum for efficient production of pyruvate from glucose by additional deletion of the ldhA gene encoding NAD+-dependent l-lactate dehydrogenase (LdhA) and introduction of a attenuated variant of the acetohydroxyacid synthase (△C–T IlvN). The latter modification abolished overflow metabolism towards l-valine and shifted the product spectrum to pyruvate production. In shake flasks, the resulting strain C. glutamicumaceEpqoldhA △C–T ilvN produced about 190 mM pyruvate with a Y P/S of 1.36 mol per mol of glucose; however, it still secreted significant amounts of l-alanine. Additional deletion of genes encoding the transaminases AlaT and AvtA reduced l-alanine formation by about 50%. In fed-batch fermentations at high cell densities with adjusted oxygen supply during growth and production (0–5% dissolved oxygen), the newly constructed strain C. glutamicumaceEpqoldhA △C–T ilvNalaTavtA produced more than 500 mM pyruvate with a maximum yield of 0.97 mol per mole of glucose and a productivity of 0.92 mmol g(CDW)−1 h−1 (i.e., 0.08 g g(CDW) −1 h−1) in the production phase.  相似文献   

11.
This is the first report describing the complete oxidation of dimethyl sulfide (DMS) to sulfate by an anoxygenic, phototrophic purple sulfur bacterium. Complete DMS oxidation was observed in cultures of Thiocapsa roseopersicina M11 incubated under oxic/light conditions, resulting in a yield of 30.1 mg protein mmol–1. No oxidation of DMS occurred under anoxic/light conditions. Chloroform, methyl butyl ether, and 3-amino-1,2,4-triazole, which are specific inhibitors of aerobic DMS oxidation in thiobacilli and hyphomicrobia, did not affect DMS oxidation in strain M11. This could be due to limited transport of the inhibitors through the cell membrane. The growth yield on sulfide as sole electron donor was 22.2 mg protein mmol–1 under anoxic/light conditions. Since aerobic respiration of sulfide would have resulted in yields lower than 22 mg protein mmol–1, the higher yield on DMS under oxic/light conditions suggests that the methyl groups of DMS have served as an additional carbon source or as an electron donor in addition to the sulfide moiety. The kinetic parameters V max and K m for DMS oxidation under oxic/light conditions were 12.4 ± 1.3 nmol (mg protein)–1 min–1 and 2 μM, respectively. T. roseopersicina M11 also produced DMS by cleavage of dimethylsulfoniopropionate (DMSP). Specific DMSP cleavage rates increased with increasing initial substrate concentrations, suggesting that DMSP lyase was only partly induced at lower initial DMSP concentrations. A comparison of T. roseopersicina strains revealed that only strain M11 was able to oxidize DMS and cleave DMSP. Both strain M11 and strain 5811 accumulated DMSP intracellularly during growth, while strain 1711 showed neither of these characteristics. Phylogenetic comparison based on 16S rRNA gene sequence revealed a similarity of 99.0% between strain M11 and strain 5811, and 97.6% between strain M11 and strain 1711. DMS and DMSP utilization thus appear to be strain-specific. Received: 26 March 1999 / Accepted: 18 June 1999  相似文献   

12.
Seasonal activities of the digestive enzyme trypsin were measured between August 1998 and May 1999 to study different nutritional strategies of the two copepods Pseudocalanus minutus and Oithona similis in the Arctic Kongsfjorden (Svalbard) using a highly sensitive fluorescence technique. Stage-, depth- and season-specific characteristics of digestive activity were reflected in the trypsin activity. P. minutus females and stage V copepodids (C) had highest trypsin activities in spring during reproduction (197.5 and 145.7 nmol min−1 ng C−1, respectively). In summer stages CIII–V and in autumn stages CIV and V had high activities (80–116 nmol min−1 ng C−1) in the shallow layer (< 100 m) presumably as a consequence of prolonged feeding before descending to overwintering depth. Trypsin activities at depth (> 100 m) in summer and autumn were low in stages CIII and CIV (29–60 nmol min−1 ng C−1) and in winter in all stages in both layers (20–43 nmol min−1 ng C−1). Based on low trypsin activity, males most likely did not feed. In O. similis, the spring phytoplankton bloom did not significantly affect trypsin activity as compared to the other seasons. O. similis CV and females had high trypsin activities in summer in the deep stratum (304.5 nmol min−1 ng C−1), which was concomitant with reproductive processes and energy storage for overwintering. In autumn, stage CV and female O. similis had significantly higher activities than stage CIV (130–152 versus 78 nmol min−1 ng C−1), which is in accordance with still ongoing developmental and reproductive processes in CVs and females. Comparisons of both species revealed different depth-related responses emphasizing different nutritional preferences: the mainly herbivorous P. minutus is more actively feeding in the shallow layer, where primary production occurs, whereas the omnivorous O. similis is not as much restricted to a certain depth layer, when searching for food. P. minutus had lower levels of trypsin activity during all seasons. In contrast to P. minutus, higher enzyme activities in males of O. similis suggest that they continue to feed and survive after fertilization of females.  相似文献   

13.
14.
The anaerobic fungus Piromyces sp. strain E2 metabolizes xylose via xylose isomerase and d-xylulokinase as was shown by enzymatic and molecular analyses. This resembles the situation in bacteria. The clones encoding the two enzymes were obtained from a cDNA library. The xylose isomerase gene sequence is the first gene of this type reported for a fungus. Northern blot analysis revealed a correlation between mRNA and enzyme activity levels on different growth substrates. Furthermore, the molecular mass calculated from the gene sequence was confirmed by gel permeation chromatography of crude extracts followed by activity measurements. Deduced amino acid sequences of both genes were used for phylogenetic analysis. The xylose isomerases can be divided into two distinct clusters. The Piromyces sp. strain E2 enzyme falls into the cluster comprising plant enzymes and enzymes from bacteria with a low G+C content in their DNA. The d-xylulokinase of Piromyces sp. strain E2 clusters with the bacterial d-xylulokinases. The xylose isomerase gene was expressed in the yeast Saccharomyces cerevisiae, resulting in a low activity (25±13 nmol min–1mg protein-1). These two fungal genes may be applicable to metabolic engineering of Saccharomyces cerevisiae for the alcoholic fermentation of hemicellulosic materials.  相似文献   

15.
Summary The gene braB, encoding the Na-coupled carrier for branched-chain amino acids in Pseudomonas aeruginosa PAO, was cloned on cosmid pMMB34. The cosmid clones carrying the braB gene were identified as those that restored growth at low leucine concentration and Na-dependent leucine transport activity to P. aeruginosa PAO3536 defective in the transport of branched-chain amino acids. Determination of the nucleotide sequence of the DNA fragment shows that the braB gene comprises 1311 bp and encodes a hydrophobic protein of 437 amino acids with a calculated Mr of 45279. The hydropathy profile suggests that there exist in the carrier protein 12 hydrophobic segments long enough to traverse the membrane. The amino acid sequence shows a high degree of homology with thebrnQ product, a branched-chain amino acid carrier of Salmonella typhimurium, while no homology in the nucleotide sequences is found in the braB and brnQ genes.  相似文献   

16.
Xanthobacter tagetidis grew as a chemolithotrophic autotroph on thiosulfate and other inorganic sulfur compounds, as a heterotroph on thiophene-2-carboxylic acid, acetic acid and α-ketoglutaric acid, and as a mixotroph on thiosulfate in combination with thiophene-2-carboxylic acid and/or acetic acid. Autotrophic growth on one-carbon organosulfur compounds, and intermediates in their oxidation are also reported. Thiosulfate enhanced the growth yields in mixotrophic cultures, presumably by acting as a supplementary energy source, since ribulose bisphosphate carboxylase was only active in thiosulfate-grown cells and was not detected in mixotrophic cultures using thiosulfate with thiophene-2-carboxylic acid. Bacteria grown on thiophene-2-carboxylic acid also oxidized sulfide, thiosulfate and tetrathionate, indicating these as possible sulfur intermediates in thiophene-2-carboxylic acid degradation. Thiosulfate and tetrathionate were oxidized completely to sulfate and, consequently, did not accumulate as products of thiophene-2-carboxylic acid oxidation in growing cultures. K m and V max values for the oxidation of thiosulfate, tetrathionate or sulfide were 13 μM and 83 nmol O2 min–1 (mg dry wt.)–1, respectively; thiosulfate and tetrathionate became autoinhibitory at concentrations above 100 μM. The true growth yield (Ymax) on thiophene-2-carboxylic acid was estimated from chemostat cultures (at dilution rates of 0.034–0.094 h–1) to be 112.2 g mol–1, with a maintenance coefficient (m) of 0.3 mmol thiophene-2-carboxylic acid (g dry wt.)–1 h–1, and the maximum specific growth rate (μmax) was 0.116 h–1. Growth in chemostat culture at a dilution rate of 0.041 h–1 indicated growth yields [g dry wt. (mol substrate)–1] of 8.1 g (mol thiosulfate)–1, 60.9 g (mol thiophene-2-carboxylic acid)–1, and 17.5 g (mol acetic acid)–1, with additive yields for growth on mixtures of these substrates. At a dilution rate of 0.034 h–1, yields of 57.8 g (mol α-ketoglutaric acid)–1 and 60.7 g (mol thiophene-2-carboxylic acid)–1 indicated some additional energy conservation from oxidation of the thiophene-sulfur. SDS-PAGE of cell-free preparations indicated a polypeptide (M r, 21.0 kDa) specific to growth on thiophene-2-carboxylic acid for which no function can yet be ascribed: no metabolism of thiophene-2-carboxylic acid by cell-free extracts was detected. It was shown that X. tagetidis exhibits a remarkable degree of metabolic versatility and is representative of facultatively methylotrophic and chemolithotrophic autotrophs that contribute significantly to the turnover of simple inorganic and organic sulfur compounds (including substituted thiophenes) in the natural environment. Received: 1 July 1997 / Accepted: 3 November 1997  相似文献   

17.
Nitrite-driven anaerobic ATP synthesis in barley and rice root mitochondria   总被引:4,自引:0,他引:4  
Mitochondria isolated from the roots of barley (Hordeum vulgare L.) and rice (Oryza sativa L.) seedlings were capable of oxidizing external NADH and NADPH anaerobically in the presence of nitrite. The reaction was linked to ATP synthesis and nitric oxide (NO) was a measurable product. The rates of NADH and NADPH oxidation were in the range of 12–16 nmol min−1 mg−1 protein for both species. The anaerobic ATP synthesis rate was 7–9 nmol min−1 mg−1 protein for barley and 15–17 nmol min−1 mg−1 protein for rice. The rates are of the same order of magnitude as glycolytic ATP production during anoxia and about 3–5% of the aerobic mitochondrial ATP synthesis rate. NADH/NADPH oxidation and ATP synthesis were sensitive to the mitochondrial inhibitors myxothiazol, oligomycin, diphenyleneiodonium and insensitive to rotenone and antimycin A. The uncoupler FCCP completely eliminated ATP production. Succinate was also capable of driving ATP synthesis. We conclude that plant mitochondria, under anaerobic conditions, have a capacity to use nitrite as an electron acceptor to oxidize cytosolic NADH/NADPH and generate ATP.  相似文献   

18.
19.
Methyl mercury uptake in free cells and different immobilizates of the cyanobacteriumNostoc calcicola has been examined. The general growth of the immobilized cyanobacterial cells could be negatively correlated with methyl mercury uptake. Alginate spheres proved most efficient in terms of uptake rate (0.48 nmol mg protein–1 min–1, 10 min) and total bioaccumulation (10.71 nmol mg protein–1, 1 h) with a bioconcentration factor of 3.3×103. Alginate biofilms showed a faster methyl mercury accumulation rate (0.83 nmol mg protein–1 min–1, 10 min) with a saturation of 10.28 nmol mg protein–1 reached within only 30 min (bioconcentration factor, 3.1×103). Foam preparations with a slow initial uptake approximated biofilms but were characterized by a lower bioconcentration factor (2.8×103). Free cells, in comparison, maintained the initial slow rate of uptake (0.62 nmol mg protein–1 min–1, 10 min), saturating at 30 min (8.81 nmol mg protein–1), and the resultant lowest bioconcentration factor (2.7×103). Cell ageing (30 days) brought a drastic reduction (3-fold) in organomercury uptake by free cells while alginate spheres maintained the same potential. Foam preparations of the same age showed a significant improvement in methyl mercury uptake followed by only a marginal decline in alginate biofilms. Data are discussed in the light of the physiological efficiency and longevity of immobilized cells.  相似文献   

20.
The uptake of soluble phosphate by the green sulfur bacterium Chlorobium limicola UdG6040 was studied in batch culture and in continuous cultures operating at dilution rates of 0.042 or 0.064 h–1. At higher dilution rates, washout occurred at phosphate concentrations below 7.1 μM. This concentration was reduced to 5.1 μM when lower dilution rates were used. The saturation constant for growth on phosphate (K μ) was between 2.8 and 3.7 μM. The specific rates of phosphate uptake in continuous culture were fitted to a hyperbolic saturation model and yielded a maximum rate (Va max) of 66 nmol P (mg protein)–1 h–1 and a saturation constant for transport (K t) of 1.6 μM. In batch cultures specific rates of phosphate uptake up to 144 nmol P (mg protein)–1 h–1 were measured. This indicates a difference between the potential transport of cells and the utilization of soluble phosphate for growth, which results in a significant change in the specific phosphorus content. The phosphorus accumulated within the cells ranged from 0.4 to 1.1 μmol P (mg protein)–1 depending on the growth conditions and the availability of external phosphate. Transport rates of phosphate increased in response to sudden increases in soluble phosphate, even in exponentially growing cultures. This is interpreted as an advantage that enables Chl. limicola to thrive in changing environments. Received: 9 February 1998 / Accepted: June 1998  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