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.
The effect of gene knockout on metabolism in the pflA, pflB, pflC, and pflD mutants of Escherichia coli was investigated. Batch cultivations of the pfl mutants and their parent strain were conducted using glucose as a carbon source. It was found that pflA and pflB mutants, but not pflC and pflD mutants, produced large amounts of d-lactate from glucose under the microaerobic condition, and the maximum yield was 73%. In order to investigate the metabolic regulation mechanism, we measured enzyme activities for the following eight enzymes: glucose 6-phosphate dehydrogenase, 6-phosphogluconate dehydrogenase, glyceraldehyde 3-phosphate dehydrogenase (GAPDH), pyruvate kinase, lactate dehydrogenase (LDH), phosphoenolpyruvate carboxylase, acetate kinase, and alcohol dehydrogenase. Intracellular metabolite concentrations of glucose 6-phosphate, fructose 1,6-bisphosphate, phosphoenolpyruvate, pyruvate, acetyl coenzyme A as well as ATP, ADP, AMP, NADH, and NAD+ were also measured. It was shown that the GAPDH and LDH activities were considerably higher in pflA and pflB mutants, which implies coupling between NADH production and consumption between the two corresponding reactions. The urgent energy requirement was shown by the lower ATP/AMP level due to both oxygen limitation and pfl gene knockout, which promoted significant stepping-up of glycolysis when using glucose as a carbon source. It was shown that the demand for energy is more important than intracellular redox balance, thus excess NADH produced through GAPDH resulted in a significantly higher intracellular NADH/NAD+ ratio in pfl mutants. Consequently, the homolactate production was achieved to meet the requirements of the redox balance and the energy production through glycolysis. The effect of using different carbon sources such as gluconate, pyruvate, fructose, and glycerol was investigated.  相似文献   

2.
A selection system based on the phosphomannose-isomerase gene (pmi) as a selectable marker and mannose as the selective agent was evaluated for the transformation of apple (Malus domestica Borkh.). Mannose is an unusable carbon source for many plant species. After uptake, mannose is phosphorylated by endogenous hexokinases to mannose-6-phosphate. The accumulation of mannose-6-phosphate leads to a block in glycolysis by inhibition of phosphoglucose-isomerase, resulting in severe growth inhibition. The phosphomannose-isomerase is encoded by the manA gene from Escherichia coli and catalyzes the conversion of mannose-6-phosphate to fructose-6-phosphate, an intermediate of glycolysis. Transformed cells expressing the manA gene can therefore utilize mannose as a carbon and survive on media containing mannose. The manA gene along with a β-glucuronidase (GUS) gene was transferred into apple cv. ‘Holsteiner Cox’ via Agrobacterium tumefaciens-mediated transformation. Leaf explants were selected on medium supplemented with different concentrations and combinations of mannose and sorbitol to establish an optimized mannose selection protocol. Transgenic lines were regenerated after an initial selection pressure of 1–2 g l−1 mannose in combination with 30 g l−1 sorbitol followed by a stepwise increase in the mannose concentration up to 10 g l−1 and simultaneous decrease in the sorbitol concentration. Integration of transgenes in the apple genome of selected plants was confirmed by PCR and southern blot analysis. GUS histochemical and chlorophenol red (CPR) assays confirmed activity of both transgenes in regenerated plants. The pmi/mannose selection system is shown to be highly efficient for producing transgenic apple plants without using antibiotics or herbicides.  相似文献   

3.
Ehrlich ascites carcinoma cells depleted of K+ and provided with 5.5 mM K+ in isosmotic 50 mM tris(hydroxymethyl)methylglycine buffer at pH 7.4 and 38 °C take up K+ from the medium at a rate of 6 μmoles/ml intracellular fluid per min. Depleted cells exposed to K+ for 2 min prior to glucose addition exhibit a higher initial rate of glycolysis, a lower glycose-6-P accumulation, and a higher fructose-1,6-P2 accumulation than depleted cells incubated in a K+-free medium. Both the K+ transport and the effect of K+ on glycolysis are blocked by 2 mM oubain.Calculation of thein vitro velocities of glycolytic enzymes from the rates of accumulation of lactate and glycolytic intermediates shows that the presence of K+ accelerates the velocities of fructose-6-phosphate kinase and lactate dehydrogenase about 2-fold and the velocity of hexokinase about 1.5-fold during the first 15 s. In either the presence or absence of K+, the hexokinase velocity is highest immediately after glucose addition and declines sharply with time; this decline is greater than would be predicted by product inhibition by the accumulated glucose-6-P. The maximal stimulation of fructose-6-phosphate kinase attibutable to the increasing intarcellular K+ concentration is only 1.25-fold. These observations indicate that the initial acceleration in glycolysis is not simply mediated through a direct K+ activation of fructose-6-phosphate kinase.The calculated theoretical rate of ATP generation by glycolysis shows that glycolysis is an ATP-utilizing system for the first 5–10 s both in the presence and in the absence of K+. Hence, the initial stimulation of glycolysis by K+ is not a consequence of an increased rate of ATP hydrolysis associated with K+ transport, although this mechanism may be responsible for the stimulation of steady-state glycolysis.The initial rate of phosphate ester (hexose and triose phosphates) accumulation corresponds to be rate of ATP generation by the “tail-end” of glycolysis, or twice the rate of lactate accumulation, in either the absence or presence of K+, but both the rate and the maximal level of ester accumulated are higher in the presence of K+. This implies that the oxidatively generated pool of ATP which is diverted from endogenous reactions to hexokinase and fructose-6-phosphate kinase on the introduction of glucose is larger in the presence of K+.Valinomycin (0.27 μM) under certain conditions can produce effects on the glycolysis of non-depleted cells which superficially resemble the effects of K+ on depleted cells. However, unlike K+, valinomycin stimulates the initial rate of glycolytic ATP generation, and abolishes the initial correspondence between the ATP generation by the “tail-end” of glycolysis and phosphate ester accumulation. These observations are interpreted to mean that valinomycin introduces an ATPase activity effective on glycolytically generated ATP.Comparison of the theoretical ATP generation in the presence and absence of K+ indicates that approximately one ATP is hydrolyzed for each K+ transported.  相似文献   

4.
A new selection system for onion transformation that does not require the use of antibiotics or herbicides was developed. The selection system used the Escherichia coli gene that encodes phosphomannose isomerase (pmi). Transgenic plants carrying the manA gene that codes for pmi can detoxify mannose-6-phosphate by conversion to fructose-6-phosphate, an intermediate of glycolysis, via the pmi activity. Six-week-old embryogenic callus initiated from seedling radicle was used for transformation. Transgenic plants were produced efficiently with transformation rates of 27 and 23% using Agrobacterium and biolistic system, respectively. Untransformed shoots were eliminated by a stepwise increase from 10 g l−1 sucrose with 10 g l−1 mannose in the first selection to only10 g l−1 mannose in the second selection. Integrative transformation was confirmed by PCR, RT-PCR and Southern hybridization. An erratum to this article can be found at  相似文献   

5.
Clostridium thermoaceticum contains interesting enzymes suitable for redox reactions. Various AMAPOR (artificial-mediator-accepting pyridine-nucleotide oxidoreductase) activities were used for electromicrobial pyridine nucleotide regeneration. The combination of AMAPOR with commercially available pyridine-nucleotide-dependent oxidoreductases led to (S)-glutamate, (2R,3S)-isocitrate, (2S,3R)-isocitrate, 6-phosphogluconate and ribulose 5-phosphate. The redox equivalents were provided by electrochemically regenerated artificial mediators. Methylviologen or cobalt sepulchrate were used for NAD(P)H regeneration, whereas carboxamidomethylviologen (CAV) or anthraquinone sulphonates (AQ-S) were suitable for NAD(P)+ regeneration. With resting cells of C. thermoaceticum productivity numbers {mmol product/[biocatalyst (kg dry weight) × time (h)]} of about 30 000 for NADPH, 7000 for NADH and 14 000 for NADP+ regeneration could be reached. The cycle number for NADPH regeneration was up to 4300, that for NADP+ regeneration was at least 1600. An aldehyde and an alcohol oxidoreductase were used to reduce non-activated carboxylic acids to the alcohols and to dehydrogenate primary alcohols to the aldehydes or carboxylates. The electromicrobial reduction of 6-chloropyridine 3-carboxylate to the corresponding alcohol was compared with the reduction by CO as electron donor.The application of phenothiazine-dye-type mediators (thionine, methylene blue) converted primary alcohols to the aldehydes with productivity numbers up to 1400 in the presence of hydrazine as aldehyde scavenger. With CAV or AQ-S, alcohols were dehydrogenated to carboxylic acids with productivity numbers of almost 1700.  相似文献   

6.
Effects of glucose concentration and anoxia upon the metabolite concentrations and rates of glycolysis and respiration have been investigated in the perfused liver of the fetal guinea pig. In most cases the metabolite concentrations in the perfused liver were similar to those observed in vivo. Between 50 days and term there was a fall in the respiratory rate and in the concentration of ATP and fructose 1,6-diphosphate and an increase in the concentration of glutamate, glycogen and glucose. Reducing the medium glucose concentration from 10 mM to 1 mM or 0.1 mM depressed lactate production and the concentration of most of the phosphorylated intermediates (except 6-phosphogluconate) in the liver of the 50-day fetus. This indicates a fall in glycolytic rate which is not in accord with the known kinetic properties of hexokinase in the fetal liver. Anoxia increased lactate production by, and the concentrations of, the hexose phosphates ADP and AMP in the 50-day to term fetal liver, while the concentration of ribulose 5-phosphate, ATP and some triose phosphates fell. These results are consistent with an activation of glycolysis, particularly at phosphofructokinase and of a reduction in pentose phosphate pathway activity, particularly at 6-phosphogluconate dehydrogenase.The calculated cytosolic NAD+/NADH ratio for the perfused liver was similar to that measured in vivo and evidence is presented to suggest that the dihydroxyacetone phosphate/glycerol 3-phosphate ratio gives a better indication of cytosolic redox than the lactate/pyruvate ratio. The present observations indicate that phosphofructokinase and hexokinase and possibly pyruvate kinase control the glycolytic rate and that glyceraldehyde-3-phosphate dehydrogenase is at equilibrium in the perfused liver of the fetal guinea pig.  相似文献   

7.
The β-proteobacterium Ralstonia eutropha H16 utilizes fructose and gluconate as carbon sources for heterotrophic growth exclusively via the Entner–Doudoroff pathway with its key enzyme 2-keto-3-deoxy-6-phosphogluconate (KDPG) aldolase. By deletion of the responsible gene eda, we constructed a KDPG aldolase-negative strain, which is disabled to supply pyruvate for energy metabolism from fructose or gluconate as sole carbon sources. To restore growth on fructose, an alternative pathway, similar to the fructose-6-phosphate shunt of heterofermentative bifidobacteria, was established. For this, the xfp gene from Bifidobacterium animalis, coding for a bifunctional xylulose-5-phosphate/fructose-6-phosphate phosphoketolase (Xfp; Meile et al. in J Bacteriol 183:2929–2936, 2001), was expressed in R. eutropha H16 PHB4 Δeda. This Xfp catalyzes the phosphorolytic cleavage of fructose 6-phosphate to erythrose 4-phosphate and acetylphosphate as well as of xylulose 5-phosphate to glyceralaldehyde 3-phosphate and acetylphosphate. The recombinant strain showed phosphoketolase (PKT) activity on either substrate, and was able to use fructose as sole carbon source for growth, because PKT is the only enzyme that is missing in R. eutropha H16 to establish the artificial fructose-6-phosphate shunt. The Xfp-expressing strain R. eutropha H16 PHB4 Δeda (pBBR1MCS-3::xfp) should be applicable for a novel variant of a plasmid addiction system to stably maintain episomally encoded genetic information during fermentative production processes. Plasmid addiction systems are often used to ensure plasmid stability in many biotechnology relevant microorganisms and processes without the need to apply external selection pressure, like the addition of antibiotics. By episomal expression of xfp in a R. eutropha H16 mutant lacking KDPG aldolase activity and cultivation in mineral salt medium with fructose as sole carbon source, the growth of this bacterium was addicted to the constructed xfp harboring plasmid. This novel selection principle extends the applicability of R. eutropha H16 as production platform in biotechnological processes.  相似文献   

8.
It has been suggested by some authors that during amphibian development, due to the higher glucose-6-phosphate dehydrogenase (EC 1.1.1.49) activity compared to that of 6-phosphogluconate dehydrogenase (EC 1.1.1.43), 6-phosphogluconate could accumulate in the embryo tissues and regulate the channelling of glucose-6-phosphate into glycolysis. Here, on the base of the specific activities of glucose-6-phosphate dehydrogenase, 6-phosphogluconate dehydrogenase and glucose-6-phosphate isomerase (EC 5.3.1.9) found in the embryos of Bufo bufo during development, it is discussed whether 6-phosphogluconate can accumulate and play a regulative role on glucose-6-phosphate metabolism in the anuran embryo.  相似文献   

9.
1. Substrate cycling of fructose 6-phosphate through reactions catalysed by phosphofructokinase and fructose diphosphatase was estimated in bumble-bee (Bombus affinis) flight muscle in vivo. 2. Estimations of substrate cycling of fructose 6-phosphate and of glycolysis were made from the equilibrium value of the 3H/14C ratio in glucose 6-phosphate as well as the rate of 3H release to water after the metabolism of [5-3H,U-14C]glucose. 3. In flight, the metabolism of glucose proceeded exclusively through glycolysis (20.4μmol/min per g fresh wt.) and there was no evidence for substrate cycling. 4. In the resting bumble-bee exposed to low temperatures (5°C), the pattern of glucose metabolism in the flight muscle was altered so that substrate cycling was high (10.4μmol/min per g fresh wt.) and glycolysis was decreased (5.8μmol/min per g fresh wt.). 5. The rate of substrate cycling in the resting bumble-bee flight muscle was inversely related to the ambient temperature, since at 27°, 21° and 5°C the rates of substrate cycling were 0, 0.48 and 10.4μmol/min per g fresh wt. respectively. 6. Calcium ions inhibited fructose diphosphatase of the bumble-bee flight muscle at concentrations that were without effect on phosphofructokinase. The inhibition was reversed by the presence of a Ca2+-chelating compound. It is proposed that the rate of fructose 6-phosphate substrate cycling could be regulated by changes in the sarcoplasmic Ca2+ concentration associated with the contractile process.  相似文献   

10.
Yeom SJ  Kim YS  Lim YR  Jeong KW  Lee JY  Kim Y  Oh DK 《Biochimie》2011,93(10):1659-1667
Mannose-6-phosphate isomerase catalyzes the interconversion of mannose-6-phosphate and fructose-6-phosphate. The gene encoding a putative mannose-6-phosphate isomerase from Thermus thermophilus was cloned and expressed in Escherichia coli. The native enzyme was a 29 kDa monomer with activity maxima for mannose 6-phosphate at pH 7.0 and 80 °C in the presence of 0.5 mM Zn2+ that was present at one molecule per monomer. The half-lives of the enzyme at 65, 70, 75, 80, and 85 °C were 13, 6.5, 3.7, 1.8, and 0.2 h, respectively. The 15 putative active-site residues within 4.5 Å of the substrate mannose 6-phosphate in the homology model were individually replaced with other amino acids. The sequence alignments, activities, and kinetic analyses of the wild-type and mutant enzymes with amino acid changes at His50, Glu67, His122, and Glu132 as well as homology modeling suggested that these four residues are metal-binding residues and may be indirectly involved in catalysis. In the model, Arg11, Lys37, Gln48, Lys65 and Arg142 were located within 3 Å of the bound mannose 6-phosphate. Alanine substitutions of Gln48 as well as Arg142 resulted in increase of Km and dramatic decrease of kcat, and alanine substitutions of Arg11, Lys37, and Lys65 affected enzyme activity. These results suggest that these 5 residues are substrate-binding residues. Although Trp13 was located more than 3 Å from the substrate and may not interact directly with substrate or metal, the ring of Trp13 was essential for enzyme activity.  相似文献   

11.
Cell-free preparations of Chlorella pyrenoidosa Chick, van Niel's strain, were assayed for oxidative enzymes, utilizing isotopic and spectrophotometric techniques. The enzyme activity of heterotrophic and autotrophic cells was compared. The study was divided into categories, one concerned with the spectrophotometric detection of enzymes involved in the initial reactions of glycolysis and the hexose monophosphate shunt, and the other with the direct oxidation of glucose as compared with that oxidized via glycolysis. The reduction of pyridine nucleotides in crude extracts was studied with glucose, glucose-6-phosphate, 6-phosphogluconate, and fructose-1-6-diphosphate as substrates. Enzymes detected in both heterotrophic and autotrophic cells were hexokinase, fructose-diphosphate-aldolase, NAD-linked 3-phosphoglyceraldchyde dehydrogenase, glucose-6-phosphate dehydrogenase, 6-phosphogluconate dehydrogenase, and a NADP-linked 3-phosphoglyceraldchyde dehydrogenase. In addition to isotopic studies designed to make an appraisal of the hexose monophosphate shunt, a comparison of the rate of reduction of NADP by glucose-6-phosphate and 6-phosphogluconate in relation to the reduction of NAD by 3-phosphoglyceraldehyde was made in light- and dark-grown cells. The rate of reduction of NADP appeared to be lowered in the light-grown cells, suggesting, as did also the isotopic studies, that the hexose monophosphate shunt is less active in autotrophic metabolism than in heterotrophic metabolism.  相似文献   

12.
The consecutive genes BF0771–BF0774 in the genome of Bacteroidesfragilis NCTC 9343 were found to constitute an operon. The functional analysis of BF0772 showed that the gene encoded a novel enzyme, mannosylglucose phosphorylase that catalyzes the reaction, 4-O-β-d-mannopyranosyl-d-glucose + Pi → mannose-1-phosphate + glucose. Here we propose a new mannan catabolic pathway in the anaerobe, which involves 1,4-β-mannanase (BF0771), a mannobiose and/or sugar transporter (BF0773), mannobiose 2-epimerase (BF0774), and mannosylglucose phosphorylase (BF0772), finally progressing to glycolysis. This pathway is distributed in microbes such as Bacteroides, Parabacteroides, Flavobacterium, and Cellvibrio.  相似文献   

13.
A new selection system using mannose has been evaluated for germ-line transformation ofArabidopsis thaliana. Although mannose itself has no adverse effects on plant cells, it leads to an accumulation of mannose-6-phosphate, which depletes intracellular stores of inorganic phosphate. This results in an inhibition of plant cell growth. The selection system uses theEscherichia coli pmi gene that encodes phosphomannose isomerase (PMI). Transgenic plants carrying thepmi gene can detoxify mannose-6-phosphate by conversion to fructose-6-phosphate, an intermediate of glycolysis, via the PMI activity. Germ-line transformation ofA. thaliana followed by sterile selection on 2–5 mM of mannose resulted in the isolation of mannose-6-phosphate-resistant progeny in about 2.5% of the treated seed, consistent with transformation rates using other selection schemes. Integrative transformation was confirmed by Southern hybridization. Analysis of PMI enzyme activity demonstrated a 5-fold range of activity levels, although these differences had little effect on the ability to select transformed plants or on the growth of transformed plants on mannose. Finally, mannose selection using thepmi gene could be accomplished in sterile plates and in soil, making this an extremely versatile tool forA. thaliana transformation.  相似文献   

14.
The genes xpk1 and xpk2(Δ1–21) encoding phosphoketolase-1 and (Δ1–7)-truncated phosphoketolase-2 have been cloned from Lactobacillus plantarum and expressed in Escherichia coli. Both gene-products display phosphoketolase activity on fructose-6-phosphate in extracts. A N-terminal His-tag construct of xpk2(Δ1–21) was also expressed in E. coli and produced active His-tagged (Δ1–7)-truncated phosphoketolase-2 (hereafter phosphoketolase-2). Phosphoketolase-2 is activated by thiamine pyrophosphate (TPP) and the divalent metal ions Mg2+, Mn2+, or Ca2+. Kinetic analysis and data from the literature indicate the activators are MgTPP, MnTPP, or CaTPP, and these species activate by an ordered equilibrium binding pathway, with Me2+TPP binding first and then fructose-6-phosphate. Phosphoketolase-2 accepts either fructose-6-phosphate or xylulose-5-phosphate as substrates, together with inorganic phosphate, to produce acetyl phosphate and either erythrose-4-phosphate or glyceraldehyde-3-phosphate, respectively. Steady state kinetic analysis of acetyl phosphate formation with either substrate indicates a ping pong kinetic mechanism. Product inhibition patterns with erythrose-4-phosphate indicate that an intermediate in the ping pong mechanism is formed irreversibly. Background mechanistic information indicates that this intermediate is 2-acetyl-TPP. The irreversibility of 2-acetyl-TPP formation might explain the overall irreversibility of the reaction of phosphoketolase-2.  相似文献   

15.
Summary We used intracellular microelectrode techniques to study the mechanisms responsible for Cl secretion by canine tracheal epithelium. Tissues were treated with indomethacin (10–6 m, added to the mucosal solution) to reduce the baseline rate of Cl secretion and then stimulated by addition of epinephrine (10–6 m) or prostaglandin E1 (10–6 m) to the submucosal solution.Three conclusions emerged from our findings: First, secretagogues enhance the rate of transepithelial Cl transport primarily by increasing apical membrane Cl permeability, since: (i) stimulation of secretion produced parallel decreases in transepithelial resistance (R t) and the membrane resistance ratioR a/Rb, whereR a andR b refer to the resistances of the apical and basolateral membranes; (ii) there was an inverse relation between the short-circuit current andR a/Rb; (iii) secretagogues depolarized the electrical potential difference across the apical membrane (a) and produced an equivalent hyperpolarization of the transepithelial electrical potential difference (1) so that, in the steady-state, the basolateral membrane potential (b) was unchanged; and (iv) substitution of sulfate or gluconate for Cl in the bathing solutions prevented secretagogue-induced changes inR t, Ra/Rb, (a) and (1).Second, Cl entry into the cell across the basolateral membrane appears to be electrically-neutral since omission of Cl from the submucosal solution had no effect on (b) and did not decreaseR a/Rb as would be expected if Cl entered the cell by a conductive process.Third, secretagogues decreaseR b. Approximately 20 sec after the onset of the secretory responseR a/Rb underwent a secondary increase whileR t continued to fall. The decrease inR b may reflect an increase in basolateral membrane K permeability.  相似文献   

16.
Summary The course of glycerol biosynthesis, initiated by exposure to –4°C, was monitored in larvae of the goldenrod gall moth,Epiblema scudderiana, and accompanying changes in the levels of intermediates of glycolysis, adenylates, glycogen, glucose, fructose-2,6-bisphosphate, and fermentative end products were characterized. Production of cryoprotectant was initiated within 6 h after a switch from +16° to –4°C, with halfmaximal levels reached in 30 h and maximal content, 450–500 mol/g wet weight, achieved after 4 days. Changes in the levels of intermediates of the synthetic pathway within 2 h at –4°C indicated that the regulatory sites involved glycogen phosphorylase, phosphofructokinase, and glycerol-3-phosphatase. A rapid increase in fructose-2,6-bisphosphate, an activator of phosphofructokinase and inhibitor of fructose-1,6-bisphosphatase, appeared to have a role in maintaining flux in the direction of glycerol biosynthesis. Analysis of metabolite changes as glycerol production slowed suggested that the inhibitory restriction of the regulatory enzymes was slightly out of phase. Inhibition at the glycerol-3-phosphatase locus apparently occurred first and resulted in a build-up of glycolytic intermediates and an overflow accumulation of glucose. Glucose inhibition of phosphorylase, stimulating the conversion of the activea to the inactiveb forms, appears to be the mechanism that shuts off phosphorylase function, counteracting the effects of low temperature that are the basis of the initial enzyme activation. Equivalent experiments carried out under a nitrogen gas atmosphere suggested that the metabolic make-up of the larvae in autumn is one that obligately routes carbohydrate flux through the hexose monophosphate shunt. The consequence of this is that fermentative ATP production during anoxia is linked to the accumulation of large amounts of glycerol as the only means of maintaining redox balance.Abbreviations G6P glucose-6-phosphate - F6P fructose-6-phosphate - F1, 6P fructose-1,6-bisphosphate - F2,6P 2 fructose-2,6-bisphosphate - G3P grycerol-3-phosphate - DHAP dinydroxyacetonephosphate - GAP glyceraldehyde-3-phosphate - PEP phosphoenolpyruvate - PFK phosphofructokinase - FBPase fructose-1,6-bisphosphatase - PK pyruvate kinase  相似文献   

17.
SYNOPSIS. Tetrahymena pyriformis, strain HSM, do not have glucose-6-phosphate dehydrogenase or 6-phosphogluconate dehydrogenase, but contain transaldolase, transketolase, ribose 5-phosphate isomerase, ribulose-5-phosphate 3-epimerase, and ribokinase. The nonoxidative enzymes of the pentose phosphate shunt function in metabolism as indicated by the incorporation of label from [1-14C]ribose into CO2 and glycogen and by the increase in total glycogen content of cultures supplemented with ribose.  相似文献   

18.
Synopsis Concentrations of glycolysis and Krebs cycle metabolites in the tail tissues of mosquitofish,Gambusia holbrooki, were measured in response to starvation and to exposure to 0.62 mg Hg l–1 (as HgCl2) for 48 h. In control fish, starvation caused decreased concentrations of glucose-6-phosphate (GO, –58%) and fructose-6-phosphate (F6P, –79%). Mercury exposure resulted in decreased concentrations of G6P (– 56%) and F6P (– 79%), and to increased concentrations of pyruvate (+ 75%), -ketoglutarate (+ 41%), succinate (+ 39%), and malate (+ 47%). Krebs cycle activity increased in response to mercury exposure, perhaps in response to greater energy needs associated with maintaining homeostasis under stressful conditions. We conclude that glycolytic activity is reduced in fish exposed to mercury and that this response is similar to that caused by a cessation of feeding.  相似文献   

19.
Addition of NO3 to N-limited Selenastrum minutum during photosynthesis resulted in an immediate drop in the NADPH/NADP ratio and a slower increase of the NADH/NAD ratio. These changes were accompanied by a rapid decrease in glucose-6-phosphate and increase in 6-phosphogluconate, indicating activation of glucose-6-phosphate dehydrogenase and a role for the oxidation pentose phosphate pathway during photosynthetic NO3 assimilation. In contrast, the short-term changes in pyridine nucleotides and metabolites during photosynthetic assimilation of NH4+ were not consistent with a stimulation of the oxidative pentose phosphate pathway.  相似文献   

20.
Di-nor-benzofuran neolignan aldehydes, Δ7-3,4-methylenedioxy-3′-methoxy-8′,9′-dinor-4′,7-epoxy-8,3′-neolignan-7′-aldehyde (ocophyllal A) 1, Δ7-3,4,5,3′-tetramethoxy-8′,9′-dinor-4′,7-epoxy-8,3′-neolignan-7′-aldehyde (ocophyllal B) 2, and macrophyllin-type bicyclo[3.2.1]octanoid neolignans (7R, 8R, 3′S, 4′S, 5′R)-Δ8′-4′-hydroxy-5′-methoxy-3,4-methylenedioxy-2′,3′,4′,5′-tetrahydro-2′-oxo-7.3′,8.5′-neolignan (ocophyllol A) 3, (7R, 8R, 3′S, 4′S, 5′R)-Δ8′-4′-hydroxy-3,4,5′-trimethoxy-2′,3′,4′,5′-tetrahydro-2′-oxo-7.3′,8.5′-neolignan (ocophyllol B) 4, (7R, 8R, 3′S, 4′S, 5′R)-Δ8′-4′-hydroxy-3,4,5,5′-tetramethoxy-2′,3′,4′,5′-tetrahydro-2′-oxo-7.3′,8.5′-neolignan (ocophyllol C) 5, as well as 2′-epi-guianin 6 and (+)-licarin B 7, were isolated and characterized from leaves of Ocotea macrophylla (Lauraceae). The structures and configuration of these compounds were determined by extensive spectroscopic analyses. Inhibition of platelet activating factor (PAF)-induced aggregation of rabbit platelets were tested with neolignans 1–7. Although compound 6 was the most potent PAF-antagonist, compounds 3–5 showed some activity.  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