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.
Glutamate dehydrogenase (GDH, EC 1.4.1.2–4) and glutamine synthetase (GS, EC 6.3.1.2) activities as well as protein content and dry matter in developing kernels of winter Triticale were determined. The relatively low level of GS activity compared to high level of NAD(P)H-dependent GDH activity during intensive filling of grains with storage compounds may indicate the participation of GDH in reductive amination of 2-oxoglutarate. The amination activity of this enzyme in all grain development phases exceeded the deaminating activity several fold. Moreover, the dynamics in the change of NAD(P)H-GDH and NAD(P)+-GDH activities were analysed in various tissues of the developing grains. The high amination activity of the enzyme in the seed coat, where the intensive protein synthesis occurs would also be an indication of the anabolic function of this enzyme.  相似文献   

2.
The influence of selected factors on the activity of highly purified GDH in triticale roots was investigated in vitro. In the presence of 2-ME, NADH-GDH activity increased by 400 %, while NADPH-GDH activity rose by 500 %. No effect of reducing factors on NAD(P)+-GDH reaction was detected. The sulphydryl groups inhibitors, such as p-chloromercuribenzoate (p-CMB) and iodoacetamide, proved the strongest inhibitors of the aminative reaction. Metal-binding compounds: ethylenediaminetetraacetic acid disodium salt (EDTA) and Zinkov also considerably inhibited NAD(P)H-GDH activity. Diisopropylfluorophosphate (DFP) and pepstatin A, the inhibitors specific for -OH serine and COO aspartic acid groups respectively, caused a non-significant NAD(P)H-GDH activity decrease. Cd2+, Co2+, Hg2+, Mg2+, Pb2+ and Zn2+ ions strongly inhibited the amination reaction, whereas their inhibiting effect upon NAD+-GDH activity was negligible. Among the applied ions, only Ca2+ activated NADH-GDH.  相似文献   

3.
In the investigated 14 day old triticale seedlings a much higher GDH activity was observed in roots than in leaves. The enzyme from the roots was purified up to the state of homogeneity (about 400 fold). The purified enzyme showed a higher activity in the presence of reduced coenzyme forms (NAD(P)H) than their oxidated forms. In the presence of NAD(P)H the enzyme showed absolute specificity to 2-oxoglutarate and in cooperation with NAD(P)+ to L-glutamate. The Km values determined for particular substrates indicate a high affinity of NADPH-GDH to ammonium ions. Optimum pH, temperature and thermostability of GDH depended on the type and form of the coenzyme. Molecular mass of purified enzyme was 257 kDa. It seems that native GDH is composed of six identical subunits of the molecular mass 42.5 kDa.  相似文献   

4.
A hypothesis describing the mechanism of photoactive protochlorophyllide (P) photoreduction in vivo, relating mainly to the molecular nature of the intermediates, is proposed. The hypothesis is compatible with currently published experimental data. After illumination of etiolated barley leaves at 143 to 153 K, the absorption of P remains essentially unchanged, but a new absorption band at 690 nm is observed. Appearance of this new intermediate enables to distinguish between light and dark stages of the photoconversion reaction. When returned to the higher temperature in the dark, the treated leaves begin accumulating chlorophyllide (Chlide), concomitant with the disappearance of the 690-nm band. The decay time of the excited P (P*) is estimated at 300 ps, which approximates the time constant of photoinduced electron transfer (ET). It is suggested that the charge-transfer complex (CTC) in its ground state (GS) (ground state of CTC formed by the partial (δ) electron transfer), i.e. (Pδ−•••H–Dδ+), between P and NADPH – the electron and proton donor (H–D) – accumulates in the following sequence: P* + H–D → (P*•••H–D)→[(P*•••H–D)←(P•••H–D+)] → 1(P•••H–D+)] → 3(P•••H–D+) → (Pδ−•••H–D δ+), where an equilibrium state (ES) – [(P*•••H–D)←(P•••H–D+)] – with a lifetime of about 1 to 2 ns, exists between the local excited (LE) and ET states. The existence of a triplet ET state – 3(P•••H–D+) – is proposed because the time interval between recording of the ES and appearance of the CTC GS (35–250 ns) does not fit the lifetime of the singlet excited complex (exciplex). It is feasible that apart from NADPH, other intermediate proton carriers are contemporaneously involved in the dark reaction (Pδ−•••H–Dδ+) → Chlide, because proton binding to the C7–C8 bond in vivo takes place in the trans-configuration. The hydride ion may approach the C7–C8 bond from one side by heterolytic fission and an additional proton, donated by the protein group, may be simultaneously added to this bond from the opposite side of the porphyrin nucleus surface. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

5.
Treponema denticola convertedl-ornithine, a product ofl-arginine catabolism, to putrescine via a decarboxylation reaction and to proline via a deamination reaction. Ornithine decarboxylation byT. denticola extracts was stimulated by pyridoxal 5′-phosphate. In the absence of pyridoxal 5′-phosphate, (NH4)2SO4-fractionated extracts converted ornithine to proline and ammonia. This activity was not stimulated by α-keto acids, nicotinamide adenine dinucleotide, reduced nicotinamide adenine dinucleotide or ADP. Neither ornithine δ-transaminase (l-ornithine: 2-oxoacid aminotransferase, EC 2.6.1.13) nor Δ1 reductase [l-proline: NAD(P) 5-oxidoreductase, EC 1.5.1.2.] activity was detectable in cell extracts. These results indicate that formation of proline from ornithine inT. denticola is catalyzed by an enzyme system analogous to the ornithine cyclase (deaminating) ofClostridium sporogenes. Exogenous ornithine inhibited the growth ofT. denticola. Thus, in addition to generating putrescine and proline, the ornithine dissimilatory pathways may serve to prevent accumulation of inhibitory concentrations of ornithine in the spirochete's environment.  相似文献   

6.
7.
An N-acetyl-d-lactosamine (LacNAc) specific lectin from tubers of Alocasia cucullata was purified by affinity chromatography on asialofetuin-linked amino activated silica. The pure lectin showed a single band in SDS-PAGE at pH 8.8 and was a homotetramer with a subunit molecular mass of 13.5 kDa and native molecular mass of 53 kDa. It was heat stable up to 55 °C for 15 min and showed optimum hemagglutination activity from pH 2 to 11. The lectin was affected by denaturing agents such as urea (2 m), thiourea (2 m) and guanidine–HCl (0.5 m) and did not require Ca2+ and Mn2+ for its activity. It was a potent mitogen at 10 μg/ml towards human peripheral blood mononuclear cells with 50% growth inhibitory potential towards SiHa (human cervix ) cancer cell line at 100 μg/ml.  相似文献   

8.
 Single isolates of a mycobiont isolated from Pisonia grandis R. Br., Pisolithus tinctorius (Pers.) Coker & Couch and Tylospora fibrillosa (Burt.) Donk were compared with regard to their relative abilities to produce key enzymes of inorganic nitrogen assimilation. Nitrate reductase (NR) activities in the P. grandis mycobiont and T. fibrillosa were significantly lower than in P. tinctorius. While specific activities for glutamate dehydrogenase (GDH) were higher in P. tinctorius than the other two fungi following NH4 + pre-treatment, glutamine synthetase (GS) activity did not differ significantly between the three fungi. In all three fungi, specific activities for GS were significantly higher than for GDH. NR activity was expressed in all three fungi regardless of the nitrogen source in the medium, but in P. tinctorius diminished following continued exposure to either NO3 , NH4 +, glutamine or NO3 + glutamine. The data are discussed in relation to nitrogen utilisation by the P. grandis mycobiont. Accepted: 16 October 1997  相似文献   

9.
 The synthesis of L-isoleucine by Corynebacterium glutamicum involves 11 reaction steps, with at least 5 of them regulated in activity or expression. Using gene replacement we constructed a vector-free C. glutamicum strain having feedback-resistant aspartate kinase and feedback-resistant homoserine dehydrogenase activity. Isogenic strains carrying in addition one or several copies of feedback-resistant threonine dehydratase were made and their product accumulations compared. With strain SM1, with high threonine dehydratase activity, accumulation of 50 mM L-isoleucine was achieved, whereas with the parent strain only 4 mM L-isoleucine was obtained. Applying a closed-loop control fed-batch strategy to strain SM1 a final titre of 138 mM L-isoleucine was achieved with an integral molar yield of 0.11 mol/mol, and a maximal specific productivity of 0.28 mmol (g h)-1. This shows that high L-isoleucine yields can be obtained in the presence of one copy of feedback-resistant homoserine dehydrogenase by applying the appropriate fermentation strategy. In addition, the specific profiles of 2-oxoglutarate and pyruvate accumulation during fermentation revealed a major transition of the metabolism of C. glutamicum during the fermentation process. Received: 16 October 1995/Received revision: 21 December 1995/Accepted: 8 January 1996  相似文献   

10.
Incorporation of l- or d-Tic into position 7 of oxytocin (OT) and its deamino analogue ([Mpa1]OT) resulted in four analogues, [l-Tic7]OT (1), [d-Tic7]OT (2), [Mpa1,l-Tic7]OT (3) and [Mpa1,d-Tic7]OT (4). Their biological properties were described by Fragiadaki et al. (Eur J Med Chem 42:799–806, 2007). Their NMR study (NOESY, TOCSY, 1H–13C HSQC spectra) is presented here. Analogues 1, 3 and 4 showed partial agonistic activity, analogue 2 was pure antagonist, suggesting that a cis conformation between residues 6 and 7 of the molecule does not result in antagonistic activity. However, the reduction in agonistic activity of analogues 1, 3 and 4 in comparison to oxytocin is consistent with the reduction of the trans conformation form. Binding affinity for the human oxytocin receptor with IC50 value of 130, 730, 103, and 380 nM for peptides 1, 2, 3, and 4, respectively, showed lower affinity in the case of d analogues. Deamination slightly increased the affinity. The existence of both cis and trans configurations of the Cys6-d-Tic7 bond is supported by observation of two sets of cross-peaks for 1H and 13C nuclei for most of the residues of the peptide not only in NOESY and TOCSY but also in 1H–13C HSQC spectra. The MS and HPLC indicate the presence of a single molecule/peptide, and NMR data thus suggest that this second set of peaks is due to the cis conformation.  相似文献   

11.
Glutamate dehydrogenase (L-glutamate: NAD+ oxidoreductase, EC 1.4.1.2) was purified from Brassica napus leaves. Isoenzyme 1 (GDH1), with the lowest, and isoenzyme 7 (GDH7) with the highest electrophoretic mobility were characterized. The native GDH was estimated to have a molecular mass of about 239 kDa and consisted of six identical 41.4-kDa subunits for GDH1 and 42.4-kDa subunits for GDH7. The pH optima of both isoenzymes in amination and deamination reactions were 9.0 and 9.5, respectively. At optimum pH, the Km values for ammonium, 2-oxoglutarate, NADH, NAD and glutamate did not differ between the two isoenzymes. Addition of 10 mM EGTA inhibited the amination activity of GDH1, but that of GDH7 remained at about 30 %. Cellular fractionation experiments showed that both GDH1 and GDH7 localized in mitochondria with a loose association with the mitochondrial membrane.  相似文献   

12.
The relationship between nutrient composition, crop biomass, and glutamate dehydrogenase (GDH) isoenzyme pattern was investigated in soybean (Glycine max) and maize (Zea mays) by monitoring the nutrient induced isomerization of the enzyme from the seedling stage to the mature crop. GDH was extracted from the leaves of the plants, and the isoenzymes were fractionated by isoelectric focusing followed by native polyacrylamide gel electrophoresis. The isomerization Vmax values for soybean GDH, similar to maize GDH increased curvilinearly from 200 – 400 μmol mg−1 min−1 as the inorganic phosphate nutrient applied to the soil decreased from 50 − 0 mM. In soybean, combinations of N and K, P, or S nutrients induced the acidic and neutral isoenzymes, and gave biomass increases 25 – 50 % higher than the control plant. GDH isoenzymes were suppressed in soybean that received nutrients without N, K, or P and accordingly the biomass was about 30 % lower than the control. Treatment of maize with NPK nutrients increased the GDH Vmax values from 138.9 at the vegetative to 256.4 μmol mg−1 min−1 at the reproductive phase, and suppressed the basic isoenzymes, but induced both the acidic and neutral isoenzymes thereby inducing seed production (27.0 ± 1.4 g per plant); whereas both the acidic and basic isoenzymes were suppressed in the control maize, and seeds did not develop. Simultaneous induction of the acidic, neutral, and basic isoenzymes of GDH indicated the occurrence of senescence. Therefore in maize and soybean, the induction of the acidic and basic isoenzymes of GDH led to the enhancement of biomass. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

13.

Background  

There are three isoforms of glutamate dehydrogenase. The isoform EC 1.4.1.4 (GDH4) catalyses glutamate synthesis from 2-oxoglutarate and ammonium, using NAD(P)H. Ammonium assimilation is critical for plant growth. Although GDH4 from animals and prokaryotes are well characterized, there are few data concerning plant GDH4, even from those whose genomes are well annotated.  相似文献   

14.
The effects of l-arginine, and its analogues N ω-nitro-l-arginine methyl ester and N ω-nitro-l-arginine on vascular resistance were investigated in the intact coronary system of an isolated non-working trout heart preparation. l-Arginine, at 10–8 mol · l–1induced a slight vasodilatory effect (max 10%). N ω-nitro-l-arginine methyl ester and N ω-Nitro-l-arginine in the range 10–8–10–4 mol · l–1 caused dose-dependent increases in coronary resistance. The vasodilatory action of l-arginine was abolished when the preparation was pretreated with 10–4 mol · l–1 N ω-nitro-l-arginine or N ω-nitro-l-arginine methyl ester. Nitroprusside alone at 1 mmol · l–1 induced a maximum vasodilation (30%) of the coronary system. Methylene blue a known inhibitor of guanylate cyclase, induced a strong vasoconstriction (already significant at 10–5 mol · l–1) and was able to overcome the vasodilative effect of nitroprusside. The endothelial nitric oxide agonists acetylcholine and serotonin, established in mammalian vessels, also mediate vasodilation in trout coronary system. In 50% of preparations, acetylcholine induced a biphasic response with vasodilation at low concentration (max 15% at 10–8 mol · l–1). Serotonin displayed a dose-response vasodilation in the range 10–8–10–4 mol · l–1 (max 20%). These vasodilative effects were reduced or abolished by 10–4 mol · l–1 l-NA. These data support the existence of NO-mediated vasodilation mechanisms in the trout coronary system. Accepted: 1 July 1996  相似文献   

15.
 Rat liver arginase contains a dinuclear Mn2(II,II) center in each subunit having EPR properties similar to those observed in Mn-catalases. The principal physiologic role of arginase is catalyzing the hydrolytic cleavage of l-arginine to produce l-ornithine and urea. Here we demonstrate that arginase catalyzes the disproportionation of hydrogen peroxide by a redox mechanism analogous to Mn-catalases, but at rates that are 10–5 to 10–6 of k cat for the Mn-catalases, and also exhibits peroxidase activity. The dinuclear Mn2(II,II) center is essential for maximal catalase activity, since both the H101N and H126N mutant arginases containing only one Mn(II)/subunit have catalase activities that are <3% of that for the wild-type enzyme. Like the Mn-catalases, the catalase activity of arginase is not inhibited by millimolar concentrations of CN, the most potent inhibitor of heme catalases, or by EDTA, a chelator of free metal ions. The catalase activity of arginase is not significantly inhibited by Cl or F, in contrast to Mn-catalases, while potent inhibitors of the hydrolytic activity are also effective inhibitors of the catalase activity. These results suggest that lower affinity of hydrogen peroxide to the active site of arginase contributes to the lower catalase activity. EPR spectroscopy reveals that potent inhibitors of the hydrolytic reaction, including N ω-hydroxy-l-arginine, l-lysine, and l-valine, decouple the electronic interaction between the Mn2+ ions, most probably by removing a μ-bridging ligand or by increasing the intermanganese separation. The capacity for arginase to deliver a hydroxide ion to hydrolyze the l-arginine substrate is suggested to arise from a "dinuclear effect", wherein the two metal ions contribute more or less equivalently in deprotonation of metal-bound water molecule. Structure-reactivity analyses of these reactions will provide insights into the factors that control redox versus hydrolytic function in dimanganese clusters. Received: 18 November 1996 / Accepted: 7 April 1997  相似文献   

16.
The gene encoding an α-l-arabinofuranosidase that could biotransform ginsenoside Rc {3-O-[β-d-glucopyranosyl-(1–2)-β-d-glucopyranosyl]-20-O-[α-l-arabinofuranosyl-(1–6)-β-d-glucopyranosyl]-20(S)-protopanaxadiol} to ginsenoside Rd {3-O-[β-d-glucopyranosyl-(1–2)-β-d-glucopyranosyl]-20-O-β-d-glucopyranosyl-20(S)-protopanaxadiol} was cloned from a soil bacterium, Rhodanobacter ginsenosidimutans strain Gsoil 3054T, and the recombinant enzyme was characterized. The enzyme (AbfA) hydrolyzed the arabinofuranosyl moiety from ginsenoside Rc and was classified as a family 51 glycoside hydrolase based on amino acid sequence analysis. Recombinant AbfA expressed in Escherichia coli hydrolyzed non-reducing arabinofuranoside moieties with apparent K m values of 0.53 ± 0.07 and 0.30 ± 0.07 mM and V max values of 27.1 ± 1.7 and 49.6 ± 4.1 μmol min−1 mg−1 of protein for p-nitrophenyl-α-l-arabinofuranoside and ginsenoside Rc, respectively. The enzyme exhibited preferential substrate specificity of the exo-type mode of action towards polyarabinosides or oligoarabinosides. AbfA demonstrated substrate-specific activity for the bioconversion of ginsenosides, as it hydrolyzed only arabinofuranoside moieties from ginsenoside Rc and its derivatives, and not other sugar groups. These results are the first report of a glycoside hydrolase family 51 α-l-arabinofuranosidase that can transform ginsenoside Rc to Rd.  相似文献   

17.
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   

18.
Pyranose 2-oxidase (P2O) was purified 43-fold to apparent homogeneity from the basidiomycete Phanerochaete chrysosporium using liquid chromatography on phenyl Sepharose, Mono Q (twice) and phenyl Superose. The native enzyme has a molecular mass of about 250 kDa (based on native PAGE) and is composed of four identical subunits of 65 kDa. It contains three isoforms of isoelectric point (pI) 5.0, 5.05 and 5.15 and does not appear to be a glycoprotein. P2O is optimally stable at pH 8.0 and up to 60 °C. It is active over a broad pH range (5.0–9.0) with maximum activity at pH 8.0–8.5 and at 55 °C, and a broad substrate specificity. d-Glucose is the preferred substrate, but 1-β-aurothioglucose, 6-deoxy-d-glucose, l-sorbose, d-xylose, 5-thioglucose, d-glucono-1,5-lactone, maltose and 2-deoxy-d-glucose are also oxidised at relatively high rates. A Ping Pong Bi Bi mechanism was demonstrated for the P2O reaction at pH 8.0, with a catalytic constant (k cat) of 111.0 s−1 and an affinity constant (K m) of 1.43 mM for d-glucose and 83.2 μM for oxygen. Whereas the steady-state kinetics for glucose oxidation were unaffected by the medium at pH ≥ 7.0, at low pH both pH and buffer composition affected the P2O kinetics with the k cat/K m value decreasing with decreasing pH. The greatest effect was observed in acetate buffer (0.1 M, pH 4.5), where the k cat decreased to 60.9 s−1 and the K m increased to 240 mM. The activity of P2O was completely inhibited by 10 mM HgCl2, AgNO3 and ZnCl2, and 50% by lead acetate, CuCl2 and MnCl2. Received: 28 August 1996 / Received revision: 25 November 1996 / Accepted: 29 November 1996  相似文献   

19.
Bradyrhizobium japonicum, the nitrogen-fixing symbiotic partner of soybean, was grown on various carbon substrates and assayed for the presence of the glyoxylate cycle enzymes, isocitrate lyase and malate synthase. The highest levels of isocitrate lyase [165–170 nmol min–1 (mg protein)–1] were found in cells grown on acetate or β-hydroxybutyrate, intermediate activity was found after growth on pyruvate or galactose, and very little activity was found in cells grown on arabinose, malate, or glycerol. Malate synthase activity was present in arabinose- and malate-grown cultures and increased by only 50–80% when cells were grown on acetate. B. japonicum bacteroids, harvested at four different nodule ages, showed very little isocitrate lyase activity, implying that a complete glyoxylate cycle is not functional during symbiosis. The apparent K m of isocitrate lyase for d,l-isocitrate was fourfold higher than that of isocitrate dehydrogenase (61.5 and 15.5 μM, respectively) in desalted crude extracts from acetate-grown B. japonicum. When isocitrate lyase was induced, neither the V max nor the d,l-isocitrate K m of isocitrate dehydrogenase changed, implying that isocitrate dehydrogenase is not inhibited by covalent modification to facilitate operation of the glyoxylate cycle in B. japonicum. Received: 10 October 1997 / Accepted: 16 January 1998  相似文献   

20.
A novel phosphorylase from Clostridium phytofermentans belonging to the glycoside hydrolase family (GH) 65 (Cphy1874) was characterized. The recombinant Cphy1874 protein produced in Escherichia coli showed phosphorolytic activity on nigerose in the presence of inorganic phosphate, resulting in the release of d-glucose and β-d-glucose 1-phosphate (β-G1P) with the inversion of the anomeric configuration. Kinetic parameters of the phosphorolytic activity on nigerose were k cat = 67 s−1 and K m = 1.7 mM. This enzyme did not phosphorolyze substrates for the typical GH65 enzymes such as trehalose, maltose, and trehalose 6-phosphate except for a weak phosphorolytic activity on kojibiose. It showed the highest reverse phosphorolytic activity in the reverse reaction using d-glucose as the acceptor and β-G1P as the donor, and the product was mostly nigerose at the early stage of the reaction. The enzyme also showed reverse phosphorolytic activity, in a decreasing order, on d-xylose, 1,5-anhydro-d-glucitol, d-galactose, and methyl-α-d-glucoside. All major products were α-1,3-glucosyl disaccharides, although the reaction with d-xylose and methyl-α-d-glucoside produced significant amounts of α-1,2-glucosides as by-products. We propose 3-α-d-glucosyl-d-glucose:phosphate β-d-glucosyltransferase as the systematic name and nigerose phosphorylase as the short name for this Cphy1874 protein.  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