首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A bienzyme flow injection system is presented for the monitoring of α-ketoglutarate produced in a fermentation process, using glutamate dehydrogenase (GDH) and glutamate oxidase (GlOx) immobilised in two serially connected expanded bed reactors. The use of expanded bed resulted in unhindered passage of the bacterial cells through the columns, and thereby the need of a separate filtering step (e.g. microdialysis) was avoided. In the first reactor, α-ketoglutarate was converted to -glutamate by GDH in the presence of ammonia and NADH. In the following reactor, -glutamate was converted by GlOx to α-ketoglutarate, ammonia and hydrogen peroxide, which was detected in an electrochemical flow-through cell at +650 mV vs. Pt/(0.1 M KCl). The detection limit of α-ketoglutarate in the coupled packed bed reactors was 1 μM (defined as 3 S/N), the linear range 0–100 μM, and the sensitivity 0.80 nA/μM (R2 0.99). In the coupled expanded bed reactors, the detection limit of α-ketoglutarate was 7 μM (defined as 3 S/N), the linear range and the sensitivity being 0–500 μM and 0.11 nA/μM (R2 1.00), respectively. The response time (defined as the time between peak rise and return to baseline) was 5 min for coupled packed beds (injection of supernatant), and 12 min for coupled expanded beds (injection of sample containing cellular and particulate matter). Several other parameters, such as reactor stability, flow rate dependency, bed expansion, glutamate interference, etc. were investigated and characterised. When analysing real samples from a fermentation broth, the same results were obtained independent of the nature of the reactor system (packed or expanded bed). The hereby described system can easily be automatised and controlled from a personal computer.  相似文献   

2.
A 250- to 300-fold purification of a nicotinamide adenine denucleotide phosphate (NADP)-dependent glutamate dehydrogenase (GDH, E.C. 1.4.1.4) with a yield of 60% from a thermophilic bacillus is described. More than one NADP-specific GDH was detected by polyacrylamide gel electrophoresis. The enzyme is of high molecular weight (approximately 2 X 10-6), similar to that of the beef and frog liver GDH. The pI of the thermophilic GDH is at pH 5.24. The enzyme is highly thermostable at the pH range of 5.8 to 9.0. The purified GDH, unlike the crude enzyme, was very labile at subzero temperatures. An unidentified factor(s) from the crude cell-free extract prevented the inactivation of the purified GDH at -70 C. Various reactants of the GDH system and D-glutamate also protected, to some extent, the enzyme from inactivation at -70 C. From the Michaelis constants for glutamate (1.1 X 10-2M), NADP (3 X 10-4M), ammonia (2.1 X 10-2M), alpha-ketoglutarate (1.3 X 10-3M), and reduced NADP (5.3 X 10-5M), it is suggested that the enzyme catalyzes in vivo the formation of glutamate from ammonia and alpha-ketoglutarate. The amination of alpha-ketoglutarate and deamination of glutamate by the thermophilic GDH are optimal at the pH values of 7.2 and 8.4, respectively.  相似文献   

3.
Glutamate dehydrogenase (GDH) (L-glutamate:NADP+ oxidoreductase, deaminating, EC 1.4.1.4) from the cellulolytic ruminal bacterium Ruminococcus flavefaciens has been purified and characterized. The native enzyme and subunit are 280 and 48 kDa, respectively, suggesting that the native enzyme is a hexamer. The enzyme requires 0.5 M KCl for optimal activity and has a pH optimum of 6.9 to 7.0. The Kms for ammonia, alpha-ketoglutarate, and glutamate are 19, 0.41, and 62 mM, respectively. The sigmoidal NADPH saturation curve revealed positive cooperativity for the binding of this coenzyme. The first residue in the N-terminal amino acid sequence from R. flavefaciens GDH was alanine, suggesting that the protein may be modified posttranslationally. Comparison of the N-terminal sequence with those of Escherichia coli, Salmonella typhimurium, and Clostridium symbiosum revealed only 39% amino acid homologies. The GDH from R. flavefaciens was unique in that its specific activity was highest during ammonia-limited growth but was not affected by ammonia shock treatment (20 mM).  相似文献   

4.
Glutamate dehydrogenase (GDH) (L-glutamate:NADP+ oxidoreductase, deaminating, EC 1.4.1.4) from the cellulolytic ruminal bacterium Ruminococcus flavefaciens has been purified and characterized. The native enzyme and subunit are 280 and 48 kDa, respectively, suggesting that the native enzyme is a hexamer. The enzyme requires 0.5 M KCl for optimal activity and has a pH optimum of 6.9 to 7.0. The Kms for ammonia, alpha-ketoglutarate, and glutamate are 19, 0.41, and 62 mM, respectively. The sigmoidal NADPH saturation curve revealed positive cooperativity for the binding of this coenzyme. The first residue in the N-terminal amino acid sequence from R. flavefaciens GDH was alanine, suggesting that the protein may be modified posttranslationally. Comparison of the N-terminal sequence with those of Escherichia coli, Salmonella typhimurium, and Clostridium symbiosum revealed only 39% amino acid homologies. The GDH from R. flavefaciens was unique in that its specific activity was highest during ammonia-limited growth but was not affected by ammonia shock treatment (20 mM).  相似文献   

5.
The activities of the following enzymes were studied in connection with dinitrogen fixation in pea bacteroids: glutamine synthetase(L-glutamate: ammonia ligase (ADP-forming)(EC 6.3.1.2)(GS); glutamate dehydrogenase (NADP+)(L-glutamate: NADP+ oxidoreductase (deaminating)(EC 1.4.1.4)(GDH); glutamate synthase (L-glutamine: 2-exeglutarate aminotransferase (NADPH-oxidizing))(EC 2.6.1.53)(GOGAT). GS activity was high throughout the growth of the plant and GOGAT activity was always low. It is unlikely that GDH or the GS-GOGAT pathway can account for the incorporation of ammonia from dinitrogen fixation in the pea bacteroid,  相似文献   

6.
An NAD(+)-dependent glutamate dehydrogenase (GDH; EC 1.4.1.24) was cloned from the ruminal ciliate protozoan, Entodinium caudatum. The gene had high sequence similarity to GDH genes from the Bacteroides (class)--a class of bacteria which is highly represented in the rumen. When expressed in Escherichia coli the enzyme had a high affinity for ammonia and alpha-ketoglutarate (apparent K(m) of 2.33 and 0.71 mM, respectively) and a low affinity for glutamate (apparent K(m) of 98 mM). GDH activity and GDH mRNA concentration were increased by incubating washed E. caudatum cells with ammonia and antibiotics. These results suggest that the GDH is an anabolic enzyme catalysing the assimilation of ammonia by E. caudatum in the rumen and that the gene was probably acquired by lateral gene transfer from a ruminal bacterium.  相似文献   

7.
A Brown  A H Colen  H F Fisher 《Biochemistry》1979,18(26):5924-5928
Ammonia is known to inhibit the steady-state rate of oxidation of L-glutamate catalyzed by glutamate dehydrogenase. We reported previously [Brown, A., Colen, A. H., & Fisher, H. F. (1978) Biochemistry 17, 2031] kinetic evidence supporting the formation in the initial rapid phase of a complex which is composed of enzyme, reduced coenzyme, alpha-ketoglutarate, and ammonia. We show here that the effects of ammonia on the steady-state reaction can be correlated with transient-state kinetic effects related to the concentration of that ammonia-containing complex. These results indicate the existence of alternate reaction pathways which become important at high ammonia concentrations. These new pathways provide an additional route for the release of NADPH from the enzyme surface. The expanded mechanism shows that the noncompetitive product inhibition by ammonia can occur without the simultaneous presence of ammonia and L-glutamate on the enzyme. This mechanism also accommodates the observed substrate inhibition by L-glutamate.  相似文献   

8.
The ability of isolated pea-shoot mitochondria conditioned to incorporate ammonia into glutamate to reassimilate endogenously produced ammonia from glycine transformation was investigated. In the presence of 1 mM to 20 mM glycine less than 15% of the ammonia liberated was found to be incorporated into glutamate. Thus, a prominent role of mitochondrial glutamate dehydrogenase in the reassimilation of intramitochondrially produced ammonia can be excluded.Abbreviation GDH Glutamate dehydrogenase (L-glutamate: NAD+ oxidoreductase (deaminating), EC 1.4.1.2)  相似文献   

9.
Ammonium ions were incorporated into L-glutamate and alpha-ketoglutarate in epimastigote forms of Trypanosoma cruzi through the following enzymatic systems: NADPH and NADH-dependent glutamate dehydrogenase, NADPH-dependent glutamate synthase, L-glutamine synthetase and NADH-dependent glutamate synthase in order of decreasing specific activity (mumoles of product formed/min/mg protein). The pH optima and Km's for the glutamate dehydrogenase system were determined. Disc electrophoresis showed the presence of cathodic bands of GDH activity, which were highly dependent on NADP+.  相似文献   

10.
In the yeast Saccharomyces cerevisiae, two NADP(+)-dependent glutamate dehydrogenases (NADP-GDHs) encoded by GDH1 and GDH3 catalyze the synthesis of glutamate from ammonium and alpha-ketoglutarate. The GDH2-encoded NAD(+)-dependent glutamate dehydrogenase degrades glutamate producing ammonium and alpha-ketoglutarate. Until very recently, it was considered that only one biosynthetic NADP-GDH was present in S. cerevisiae. This fact hindered understanding the physiological role of each isoenzyme and the mechanisms involved in alpha-ketoglutarate channeling for glutamate biosynthesis. In this study, we purified and characterized the GDH1- and GDH3-encoded NADP-GDHs; they showed different allosteric properties and rates of alpha-ketoglutarate utilization. Analysis of the relative levels of these proteins revealed that the expression of GDH1 and GDH3 is differentially regulated and depends on the nature of the carbon source. Moreover, the physiological study of mutants lacking or overexpressing GDH1 or GDH3 suggested that these genes play nonredundant physiological roles. Our results indicate that the coordinated regulation of GDH1-, GDH3-, and GDH2-encoded enzymes results in glutamate biosynthesis and balanced utilization of alpha-ketoglutarate under fermentative and respiratory conditions. The possible relevance of the duplicated NADP-GDH pathway in the adaptation to facultative metabolism is discussed.  相似文献   

11.
The NAD-dependent glutamate dehydrogenase (GDH) from Dictyostelium discoideum was purified 1101-fold with a yield of 23.4%. The enzyme has an apparent Mr of 356 kDa, determined using Sephacryl S400, and a subunit molecular weight of 54 kDa on SDS-polyacrylamide gel electrophoresis. The Kms for alpha-ketoglutarate, NADH, and NH4+ are 0.36 +/- 0.03 mM, 16.0 +/- 0.1 microM, and 34.5 +/- 2.7 mM, respectively. The purified enzyme has a pH optimum of pH 7.25-7.5. At 0.1 mM, ADP and AMP stimulate GDH activity 25 and 102%, respectively. Half-maximal activity in the presence of 0.1 mM AMP for alpha-ketoglutarate, NADH, and NH4+ is reached at 2.3 +/- 0.1 mM, 71.4 +/- 5.5 microM, and 27.9 +/- 3.6 mM, respectively.  相似文献   

12.
Cells of Clostridium acetobutylicum were immobilized by adsorption onto bonechar. and used for the production of solvents (ABE fermentation) from whey permeate. When the process was performed in packed bed reactors operated in a vertical or inclined mode, solvent productivities approximating 6 kg/(m3h) were obtained. However, the systems suffered from blockage due to excess biomass production and gas hold-up. These problems were less apparent when a partially-packed bed reactor was operated in the horizontal mode. A fluidized bed reactor proved to be the most stable of the systems investigated, and a productivity of 4.8 kg/(m3h) was maintained over a period of 2000 h of operation. The results demonstrate that this type of reactor may have a useful future role in the ABE fermentation.  相似文献   

13.
pH dependence of the Coxiella burnetii glutamate transport system.   总被引:10,自引:2,他引:8  
The transport of glutamate, apparently a primary energy source for Coxiella burnetii, has been examined. C. burnetii is shown to possess a pH-dependent active transport system for L-glutamate with an apparent Kt of 61.1 microM and Vmax of 8.33 pmol/s per mg at pH 3.5. Both L-glutamine and L-asparagine competitively inhibited transport of glutamate, but D-glutamate, L-aspartate, L-glutamate-gamma-methyl ester, methionine sulfoximine, or alpha-ketoglutarate did not compete. This transport system is both temperature and energy dependent. Uptake of glutamate is highly sensitive to uncouplers of oxidative phosphorylation such as 2,4-dinitrophenol and carbonyl cyanide-m-chlorophenyl hydrazone that decrease the proton motive force across the cytoplasmic membrane. ATPase inhibitors such as dicyclohexylcarbodiimide or metabolic poisons such as KCN, NaF, or arsenite were much less effective as inhibitors of glutamate transport. Uptake of glutamate did not appear to be coupled to Na+ symport as in Escherichia coli since no monovalent cation requirement could be demonstrated. Instead, the Vmax of glutamate transport showed good correlation with the transmembrane pH gradient (delta pH). From these results, we propose that L-glutamate transport by C. burnetii is energized via a proton motive force.  相似文献   

14.
Bacillus subtilis natto is widely used in industry to produce natto, a traditional and popular Japanese soybean food. However, during its secondary fermentation, high amounts of ammonia are released to give a negative influence on the flavor of natto. Glutamate dehydrogenase (GDH) is a key enzyme for the ammonia produced and released, because it catalyzes the oxidative deamination of glutamate to alpha-ketoglutarate using NAD+ or NADP+ as co-factor during carbon and nitrogen metabolism processes. To solve this problem, we employed multiple computational methods model and re-design GDH from Bacillus subtilis natto. Firstly, a structure model of GDH with cofactor NADP+ was constructed by threading and ab initio modeling. Then the substrate glutamate were flexibly docked into the structure model to form the substrate-binding mode. According to the structural analysis of the substrate-binding mode, Lys80, Lys116, Arg196, Thr200, and Ser351 in the active site were found could form a significant hydrogen bonding network with the substrate, which was thought to play a crucial role in the substrate recognition and position. Thus, these residues were then mutated into other amino acids, and the substrate binding affinities for each mutant were calculated. Finally, three single mutants (K80A, K116Q, and S351A) were found to have significant decrease in the substrate binding affinities, which was further supported by our biochemical experiments.  相似文献   

15.
This study reports the results of experiments on continuous adsorption and desorption of Cr(VI) ions by a chemically modified and polysulfone-immobilized biomass of the fungus Rhizopus nigricans. A fixed quantity of polymer-entrapped biomass beads corresponding to 2 g of dry biomass powder was employed in packed bed, fluidized bed, and stirred tank reactor for monitoring the continuous removal and recovery of Cr(VI) ions from aqueous solution and synthetic chrome plating effluent. Parameters such as flow rate (5, 10 and 15 mL/min), inlet concentration of Cr(VI) ions (50, 100, 150 and 250 mg/L) and the depth of biosorbent packing (22.8, 11.2 and 4.9 cm) were evaluated for the packed bed reactor. The breakthrough time and the adsorption rates in the packed bed column were found to decrease with increasing flow rate and higher Cr inlet concentrations and to increase with higher depths of sorbent packing. To have a comparative analysis of Cr adsorption efficiency in different types of reactors, the fluidized bed reactor and stirred tank reactor were operated using the same quantities of biosorbent material. For the fluidized bed reactor, Cr(VI) solution of 100 mg/L was pumped at 5 mL/min and fluidized by compressed air at a flow rate of 0.5 kg/cm.(2) The stirred tank reactor had a working volume of 200 mL capacity and the inlet/outlet flow rate was 5 mL/min. The maximum removal efficiency (mg Cr/g biomass) was obtained for the stirred tank reactor (159.26), followed by the fluidized reactor (153.04) and packed bed reactor (123.33). In comparison to the adsorption rate from pure chromate solution, approximately 16% reduction was monitored for synthetic chrome plating effluent in the packed bed. Continuous desorption of bound Cr ions from the reactors was effective with 0.01 N Na(2)CO(3) and nearly 80-94% recoveries have been obtained for all the reactors.  相似文献   

16.
Human glutamate dehydrogenase (GDH), an enzyme central to the metabolism of glutamate, is known to exist in housekeeping and nerve tissue-specific isoforms encoded by the GLUD1 and GLUD2 genes, respectively. As there is evidence that GDH function in vivo is regulated, and that regulatory mutations of human GDH are associated with metabolic abnormalities, we sought here to characterize further the functional properties of the two human isoenzymes. Each was obtained in recombinant form by expressing the corresponding cDNAs in Sf9 cells and studied with respect to its regulation by endogenous allosteric effectors, such as purine nucleotides and branched chain amino acids. Results showed that L-leucine, at 1.0 mM:, enhanced the activity of the nerve tissue-specific (GLUD2-derived) enzyme by approximately 1,600% and that of the GLUD1-derived GDH by approximately 75%. Concentrations of L-leucine similar to those present in human tissues ( approximately 0.1 mM:) had little effect on either isoenzyme. However, the presence of ADP (10-50 microM:) sensitized the two isoenzymes to L-leucine, permitting substantial enzyme activation at physiologically relevant concentrations of this amino acid. Nonactivated GLUD1 GDH was markedly inhibited by GTP (IC(50) = 0.20 microM:), whereas nonactivated GLUD2 GDH was totally insensitive to this compound (IC(50) > 5,000 microM:). In contrast, GLUD2 GDH activated by ADP and/or L-leucine was amenable to this inhibition, although at substantially higher GTP concentrations than the GLUD1 enzyme. ADP and L-leucine, acting synergistically, modified the cooperativity curves of the two isoenzymes. Kinetic studies revealed significant differences in the K:(m) values obtained for alpha-ketoglutarate and glutamate for the GLUD1- and the GLUD2-derived GDH, with the allosteric activators differentially altering these values. Hence, the activity of the two human GDH is regulated by distinct allosteric mechanisms, and these findings may have implications for the biologic functions of these isoenzymes.  相似文献   

17.
Glutamate dehydrogenase (GDH) activity was determined in high-speed fractions (100,000 g for 60 min) obtained from whole rat brain homogenates after removal of a low-speed pellet (480 g for 10 min). Approximately 60% of the high-speed GDH activity was particulate (associated with membrane) and the remaining was soluble (probably of mitochondrial matrix origin). Most of the particulate GDH activity resisted extraction by several commonly used detergents, high concentration of salt, and sonication; however, it was largely extractable with the cationic detergent cetyltrimethylammonium bromide (CTAB) in hypotonic buffer solution. The two GDH activities were purified using a combination of hydrophobic interaction, ion exchange, and hydroxyapatite chromatography. Throughout these purification steps the two activities showed similar behavior. Kinetic studies indicated similar Km values for the two GDH fractions for the substrates alpha-ketoglutarate, ammonia, and glutamate; however, there were small but significant differences in Km values for NADH and NADPH. Although the allosteric stimulation by ADP and L-leucine and inhibition by diethylstilbestrol was comparable, the two GDH components differed significantly in their susceptibility to GTP inhibition in the presence of 1 mM ADP, with apparent Ki values of 18.5 and 9.0 microM GTP for the soluble and particulate fractions, respectively. The Hill plot coefficient, binding constant, and cooperativity index for the GTP inhibition were also significantly different, indicating that the two GDH activities differ in their allosteric sites. In addition, enzyme activities of the two purified proteins exhibited a significant difference in thermal stability when inactivated at 45 degrees C and pH 7.4 in 50 mM phosphate buffer.  相似文献   

18.
In C. pagurus exposed to air for 18 h, blood ammonia content decreased within the 2 first hours, then increased at a relatively constant rate (25 microM/h); blood urate content increased at a lower rate (10 microM/h) and a classical blood acidosis was observed. In the cheliped muscle, a transient 22% decrease in GDH activity for ammonia formation and a 48% increase in GDH activity in the reverse reaction (glutamate synthesis) occurred following 6 and 12 h of emersion, respectively. Changes in LDH activity, used as an indicator of anaerobic potential of muscle, were not observed, except for an 18% increase in crabs exposed to air for 24 h. The increase in blood urate content, not known as a response to emersion in decapods, appeared to be different from that observed in response to hypoxia. The relatively low blood ammonia overload and the GDH increased activity for glutamate synthesis suggested that part of the produced ammonia was stored under a bound form in some tissues. The response of C. pagurus to air exposure is discussed on account of the Storey and Storey ('90) theory.  相似文献   

19.
A new class of glutamate dehydrogenase (GDH) is reported. The GDH of Streptomyces clavuligerus was purified to homogeneity and characterized. It has a native molecular mass of 1,100 kDa and exists as an alpha(6) oligomeric structure composed of 183-kDa subunits. GDH, which requires AMP as an essential activator, shows a maximal rate of catalysis in 100 mm phosphate buffer, pH 7.0, at 30 degrees C. Under these conditions, GDH displayed hyperbolic behavior toward ammonia (K(m), 33 mm) and sigmoidal responses to changes in alpha-ketoglutarate (S(0.5) 1.3 mm; n(H) 1.50) and NADH (S(0.5) 20 microm; n(H) 1.52) concentrations. Aspartate and asparagine were found to be allosteric activators. This enzyme is inhibited by an excess of NADH or NH(4)(+), by some tricarboxylic acid cycle intermediates and by ATP. This GDH seems to be a catabolic enzyme as indicated by the following: (i) it is NAD-specific; (ii) it shows a high value of K(m) for ammonia; and (iii) when S. clavuligerus was cultured in minimal medium containing glutamate as the sole source of carbon and nitrogen, a 5-fold increase in specific activity of GDH was detected compared with cultures provided with glycerol and ammonia. GDH has 1,651 amino acids, and it is encoded by a DNA fragment of 4,953 base pairs (gdh gene). It shows strong sequence similarity to proteins encoded by unidentified open reading frames present in the genomes of species belonging to the genera Mycobacterium, Rickettsia, Pseudomonas, Vibrio, Shewanella, and Caulobacter, suggesting that it has a broad distribution. The GDH of S. clavuligerus is the first member of a class of GDHs included in a subfamily of GDHs (large GDHs) whose catalytic requirements and evolutionary implications are described and discussed.  相似文献   

20.
Glutamate dehydrogenase (GDH) catalyzes reversible oxidative deamination of l-glutamate to alpha-ketoglutarate. Enzyme activity is regulated by several allosteric effectors. Recognition of a new form of hyperinsulinemic hypoglycemia, hyperinsulinism/hyperammonemia (HI/HA) syndrome, which is caused by gain-of-function mutations in GDH, highlighted the importance of GDH in glucose homeostasis. GDH266C is a constitutively activated mutant enzyme we identified in a patient with HI/HA syndrome. By overexpressing GDH266C in MIN6 mouse insulinoma cells, we previously demonstrated unregulated elevation of GDH activity to render the cells responsive to glutamine in insulin secretion. Interestingly, at low glucose concentrations, basal insulin secretion was exaggerated in such cells. Herein, to clarify the role of GDH in the regulation of insulin secretion, we studied cellular glutamate metabolism using MIN6 cells overexpressing GDH266C (MIN6-GDH266C). Glutamine-stimulated insulin secretion was associated with increased glutamine oxidation and decreased intracellular glutamate content. Similarly, at 5 mmol/l glucose without glutamine, glutamine oxidation also increased, and glutamate content decreased with exaggerated insulin secretion. Glucose oxidation was not altered. Insulin secretion profiles from GDH266C-overexpressing isolated rat pancreatic islets were similar to those from MIN6-GDH266C, suggesting observation in MIN6 cells to be relevant in native beta-cells. These results demonstrate that, upon activation, GDH oxidizes glutamate to alpha-ketoglutarate, thereby stimulating insulin secretion by providing the TCA cycle with a substrate. No evidence was obtained supporting the hypothesis that activated GDH produced glutamate, a recently proposed second messenger of insulin secretion, by the reverse reaction, to stimulate insulin secretion.  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