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.
The acid end-products of glucose metabolism of oral and other haemophili   总被引:2,自引:0,他引:2  
The acids produced in broth culture by various species of oral haemophili and by stock strains of capsulated and other haemophili were identified and measured by gas-liquid chromatography. Succinic acid was the major acid end-product of all strains, with acetic acid also being regularly produced but in smaller amounts. A stock strain, Haemophilus parainfluenzae NCTC 4101, produced less succinic acid than other strains of haemophili. Strain NCTC 4101 possessed all the enzymes of the tricarboxylic acid cycle, as previously reported, but in the other haemophili examined only succinic dehydrogenase, fumarase and malate dehydrogenase could be detected. No other enzymes of the tricarboxylic acid cycle were detected and isocitrate lyase, malate synthase and pyruvate carboxylase were also absent. Phosphoenolpyruvate-carboxylase was present in all strains. A partial tricarboxylic acid cycle and marked malate dehydrogenase activity appear to be characteristic of haemophili. The pathway to succinate in haemophili appears to be via carboxylation of phosphoenolpyruvate to oxalacetate and thence via malate and fumarate. The results of tracer studies on a single oral strain of H. parainfluenzae using various labelled substrates were in keeping with this proposed metabolic pathway.  相似文献   

2.
The oxidation of exogenously added substrates has been studied in intact liver mitochondria isolated from the American eel, Anguilla rostrata. These data, coupled to determinations of the activity and localization of critical tricarboxylic acid (TCA) cycle enzymes, have been used to propose a pathway for the eel liver TCA cycle. (1) Isocitric, α-ketoglutaric, succinic, and malic acids are oxidized at essentially equivalent rates by eel mitochondria, with normal ADP:O and respiratory control ratios. No oxidation of citric, oxaloacetic, or pyruvic acids was detected when added alone or with malate, although oxaloacetic acid + pyruvic acid was oxidized but at a much reduced rate. (2) Radioactively labeled isocitrate was incorporated into at least α-ketoglutaric, succinic, and malic acids, indicating the eel liver TCA cycle is normal between isocitrate and malate. (3) No activity of the NAD-linked isocitrate dehydrogenase (IDH) could be detected, but NADP-IDH activities were higher in the mitochondria than cytosolic fractions. An active NADPH:NAD transhydrogenase was localized to the mitochondrial compartment. (4) These data suggest an important role for the NADP-IDH:transhydrogenase enzyme couple in eel liver TCA cycle function, and a pathway incorporating these ideas is proposed.  相似文献   

3.
The thermostability of four enzymes of the tricarboxylic acid cycle has been studied in the facultative thermophile,Bacillus coagulans. Although isocitrate dehydrogenase appeared to be more temperature-sensitive in whole-cell extracts of cultures grown at 30°C compared with that in cultures grown at 55°C, this difference could be largely eliminated by the removal of cell-wall material. The specific activity of each of the enzymes examined was approximately threefold higher in cultures grown at 55°C than in those grown at 30°C. The maximum temperature, Arrhenius plot and effect of stabilizing agents for each enzyme were examined and found to be independent of growth temperature. Sodium chloride (10% w/v) was an effective protective agent for fumarase, aconitase and malate dehydrogenase. Protection from thermal denaturation of isocitrate dehydrogenase, aconitase and fumarase but not malate dehydrogenase was also given when the enzymes were heated in the presence of their substrates. These results are discussed in light of the generalized theories of facultative thermophily which have been proposed.  相似文献   

4.
The particulate activities of Leishmania mexicana mexicana amastigote malate dehydrogenase (L-malate:NAD+ oxidoreductase, EC 1.1.1.37) and phosphoenolpyruvate carboxykinase (ATP:oxaloacetate carboxy-lyase (transphosphorylating) EC 4.1.1.49) have been purified to apparent electrophoretic homogeneity by hydrophobic interaction chromatography using Phenyl-Sepharose CL-4B, affinity chromatography using 5'AMP-Sepharose 4B, and gel filtration using Sephadex G-100. Malate dehydrogenase was purified 150-fold overall with a final specific activity of 1230 units/mg protein and a recovery of 63%. Phosphoenolpyruvate carboxykinase was purified 132-fold with a final specific activity of 30.3 units/mg protein and a recovery of 20%. Molecular weights determined by gel filtration and SDS-gel electrophoresis were 39 800 and 33 300 for malate dehydrogenase and 63 100 and 65 100 for phosphoenolpyruvate carboxykinase, respectively. Kinetic studies with malate dehydrogenase assayed in the direction of oxaloacetic acid reduction showed a Km(NADH) of 41 microM and a Km(oxaloacetic acid) of 39 microM. For malate oxidation there was a Km(malate) of 3.6 mM and a Km(NAD) of 0.79 mM. Oxaloacetic acid exhibited substrate inhibition at concentrations greater than 0.83 mM and malate was found to be a product inhibitor at high concentrations. However, there was no modification of enzyme activity by a number of glycolytic intermediates and cofactors, suggesting that malate dehydrogenase is not a major regulatory enzyme in L. m. mexicana. The results show that these L. m. mexicana amastigote enzymes are in several ways similar to their mammalian counterparts; nevertheless, their apparent importance and unique subcellular organization in the parasite make them potential targets for chemotherapeutic attack.  相似文献   

5.
Hanning I  Heldt HW 《Plant physiology》1993,103(4):1147-1154
The functioning of isolated spinach (Spinacia oleracea L.) leaf mitochondria has been studied in the presence of metabolite concentrations similar to those that occur in the cytosol in vivo. From measurements of the concentration dependence of the oxidation of the main substrates, glycine and malate, we have concluded that the state 3 oxidation rate of these substrates in vivo is less than half of the maximal rates due to substrate limitation. Analogously, we conclude that under steady-state conditions of photosynthesis, the oxidation of cytosolic NADH by the mitochondria does not contribute to mitochondrial respiration. Measurements of mitochondrial respiration with glycine and malate as substrates and in the presence of a defined malate:oxaloacetate ratio indicated that about 25% of the NADH formed in vivo during the oxidation of these metabolites inside the mitochondria is oxidized by a malate-oxaloacetate shuttle to serve extramitochondrial processes, e.g. reduction of nitrate in the cytosol or of hydroxypyruvate in the peroxisomes. The analysis of the products of the oxidation of malate indicates that in the steady state of photosynthesis the activity of the tricarboxylic acid cycle is very low. Therefore, we have concluded that the mitochondrial oxidation of malate in illuminated leaves produces mainly citrate, which is converted via cytosolic aconitase and NADP-isocitrate dehydrogenase to yield 2-oxoglutarate as the precursor for the formation of glutamate and glutamine, which are the main products of photosynthetic nitrate assimilation.  相似文献   

6.
Tricarboyxlic acid cycle activity was examined in Neisseria gonorrhoeae CS-7. The catabolism of glucose in N. gonorrheae by a combination of the Entner-Doudoroff and pentose phosphate pathways resulted in the accumulation of acetate, which was not further catabolized until the glucose was depleted or growth became limiting. Radiorespirometric studies revealed that the label in the 1 position of acetate was converted to CO2 at twice the rate of the label in the 2 position, indicating the presence of a tricarboxylic acid cycle. Growth on glucose markedly reduced the levels of all tricarboxylic acid cycle enzymes except citrate synthase (EC 4.1.3.7). Extracts of glucose-grown cells contained detectable levels of all tricarboxylic acid cycle enzymes except aconitase (EC 4.2.1.3), isocitrate dehydrogenase (EC 1.1.1.42), and a pyridine nucleotide-dependent malate dehydrogenase (EC 1.1.1.37). Extracts of cells capable of oxidizing acetate lacked only the pyridine nucleotide-dependent malate dehydrogenase. In lieu of this enzyem, a particulate pyridine nucleotide-independent malate oxidase (EC 1.1.3.3) was present. This enzyme required flavin adenine dinucleotide for activity and appeared to be associated with the electron transport chain. Radiorespirometric studies utilizing labeled glutamate demonstrated that a portion of the tricarboxylic acid cycle functioned during glucose catabolism. In spite of the presence of all tricarboxylic acid cycle enzymes, N. gonorrhoeae CS-7 was unable to grow in medium supplemented with cycle intermediates.  相似文献   

7.
Shortly after the injection of glucagon, epinephrine, norepinephrine, vasopressin, or angiotensin II into fasted rats, mitochondria isolated from their livers contained elevated concentrations of malate and oxidized citrate, alpha-ketoglutarate, and, in some cases, succinate more rapidly than mitochondria from fasted, control rats. The administration of tryptophan, lactate, or ethanol and refeeding of rats fasted 24 h result in similar elevations of mitochondrial malate concentration and oxidation of added substrates. Treatments that resulted in elevated mitochondrial malate resulted also in increased uptake of added citrate, alpha-ketoglutarate, pyruvate, and, in some cases, succinate. It is postulated that the well-documented effect of gluconeogenic hormones on mitochondrial oxidation of carboxylic substrates may be mediated by malate which not only yields oxalacetate to support the tricarboxylic acid cycle but also facilitates the transport of added substrates, and which is regenerated in the tricarboxylic acid cycle.  相似文献   

8.
Electrophoretically homogenous isoforms of malate dehydrogenase with different quaternary structure were prepared from Rhodopseudomonas palustris strain f8pt cultured photolithoheterotrophically on malate and acetate. By selective inhibition of the tricarboxylic acid cycle or glyoxylate cycle, it was shown that the dimeric isoform of the enzyme is responsible for Krebs cycle functioning and the tetrameric isoform is involved in functioning of the glyoxylate cycle.  相似文献   

9.
The role of the metabolite disposal mechanisms in the regulation of the tricarboxylic acid cycle pool size was studied in isolated perfused rat hearts oxidizing 2 mM propionate. Malate and succinate accumulated during the propionate metabolism. A further 118% increase in the malate concentration and 600% increase in the succinate concentration and a slight inhibition of the propionate uptake were observed during a subsequent KCl-induced arrest of the heart metabolizing propionate. When the mechanical activity of the heart was restored, the malate and succinate concentrations returned to the same levels as before the arrest of the heart, but the propionate uptake did not rise significantly. The mean disposal rates of the tricarboxylic acid cycle metabolites during the cardiac arrest and subsequent restoration of the activity were 1.4 and 2.4 μmol/min per g dry weight, respectively. During cardiac arrest the malate carbon disposed was almost totally recovered as C3 compounds, whereas after the increase in the ATP-consumption most of it was oxidized. The results show that propionate is oxidized by heart muscle at an appreciable rate but the disposal rate of the tricarboxylic acid cycle intermediates is not tightly regulated by the cellular energy state. Although the metabolite pool size of the tricarboxylic acid cycle responds to change in the ATP consumption, the energy state appears to have a greater effect on the fate of the C3 compounds formed than on the actual rate of C4 compound disposition.  相似文献   

10.
Summary Determinations of the momentary levels of various intermediates related to the activity of the tricarboxylic acid cycle have been made during citric acid production in high-accumulating (manganese deficient) and lowaccumulating (manganese supplemented) mycelia of Aspergillus niger. During the growth period the levels of almost all TCA cycle acids, with the exception of 2-oxo-acids, were unusually high; during the induction phase of citrate accumulation malate, fumarate, and isocitrate decreased, whereas pyruvate, oxalacetate, and citrate increased. The presence of succinate could not be demonstrated. The interrelations of the momentary concentrations of the intermediates mainly demonstrate a lack in activity of 2-oxoglutarate dehydrogenase, representing a block in the TCA cycle concomitant with a strongly operating glycolysis as a prerequisite for citrate accumulation. Inhibition studies with crude enzyme preparations suggest that an inhibition of malate dehydrogenase by citrate and also inhibition of isocitrate dehydrogenase by citrate and 2-oxoglutarate occur during the production phase as additional factors.  相似文献   

11.
Gluconeogenesis in the kidney cortex. Effects of d-malate and amino-oxyacetate   总被引:15,自引:13,他引:2  
1. Rat kidney-cortex slices incubated with d-malate alone formed very little glucose. d-Malate, however, augmented gluconeogenesis from l-lactate and inhibited gluconeogenesis from pyruvate and l-malate. 2. d-Malate had little effect on the rate of the tricarboxylic acid cycle with or without other substrates added. 3. d-Malate inhibited the activity of the l-malate dehydrogenase in a high-speed-supernatant fraction from kidney cortex. 4. It was concluded that d-malate inhibited either the operation of the cytoplasmic l-malate dehydrogenase or malate outflow from the mitochondria in the intact kidney-cortex cell. This supports the hypothesis of Lardy, Paetkau & Walter (1965) and Krebs, Gascoyne & Notton (1967) on the role of malate as carrier for carbon and reducing equivalents in gluconeogenesis. 5. Gluconeogenesis from l-lactate in kidney-cortex slices was strongly inhibited by a low concentration (0.1mm) of amino-oxyacetate, whereas glucose formation from pyruvate, malate, aspartate and several other compounds was only slightly affected. 6. High concentrations of l-aspartate largely reversed the inhibition of gluconeogenesis from l-lactate caused by amino-oxyacetate. 7. Amino-oxyacetate inhibited strongly the glutamate-oxaloacetate transaminase in the 30000g supernatant fraction of a kidney-cortex homogenate. The presence of l-aspartate decreased the inhibition of the transaminase by amino-oxyacetate. 8. Detritiation of l-[2-(3)H]aspartate was inhibited by 90% during an incubation of kidney-cortex slices with l-lactate and amino-oxyacetate. 9. Low concentrations (10mum) of artificial electron acceptors such as Methylene Blue and phenazine methosulphate abolished most of the inhibition of gluconeogenesis from l-lactate by amino-oxyacetate. This is interpreted as an activation of net malate outflow from the mitochondria by-passing the inhibited transfer of oxaloacetate. 10. These findings support the concept that transamination to aspartate is involved in the transfer of oxaloacetate from mitochondria to cytosol required in gluconeogenesis from l-lactate.  相似文献   

12.
meso-Tartrate inhibited the growth of non-meso-tartrate-utilizing strains of Salmonella typhimurium in peptone water media and mineral salts media with some, but not all, carbon sources. C-R intermediates of the tricarboxylic acid cycle or compounds readily converted to them and substrates metabolized independently of the C-6 part of the cycle spared bacteria from the inhibitory effects of meso-tartrate when added to cultures along with meso-tartrate. Experiments with cell-free extracts of non-meso-tartrate-utilizing strains from batch and continuous cultures showed that meso-tartrate was a competitive inhibitor of isocitrate dehydrogenase and isocitrate lyase activities and also inhibited citrate synthase and malate synthase activities. The synthesis of these enzymes was not inhibited by meso-tartrate. The isocitrate enzymes of meso-tartrate-utilizing strains of S. typhimurium were similarly inhibited by meso-tartrate, but inhibition of the growth of meso-tartrate-utilizing strains was demonstrable only in uninduced cultures in which the intracellular concentrations of meso-tartrate were high.  相似文献   

13.
Quinolinic acid (Q.A.) which inhibits gluconeogenesis at the site of phosphoenolpyruvate (PEP) synthesis, reduced the content of PEP while elevating that of aspartate and malate in rat livers perfused with a medium containing 10 mM L-lactate. Glucagon at 10(-9) M did not affect Q.A. inhibition of lactate gluconeogenesis nor the depression of PEP level, but further elevated malate and aspartate accumulation. Exogenous butyrate had the same effect as glucagon on these parameters. Butylmalonate (BM), an inhibitor of mitochondrial malate transport, inhibited lactate and propionate gluconeogenesis to similar extents. The addition of 10(-9) M glucagon had no effect on BM inhibition of lactate gluconeogenesis, but almost completely reversed BM inhibition of propionate gluconeogenesis. These results suggest that glucagon may act on at least two sites, resulting in elevated hepatic gluconeogenesis. First, it may stimulate dicarboxylic acid synthesis (malate and oxaloacetate, specifically) through activation of pyruvate carboxylation. Secondly, it may stimulate synthesis of other dicarboxylic acids (fumarate, for example) by activating certain steps of the tricarboxylic acid cycle. The stimulatory effect of glucagon on gluconeogenesis in the perfused rat liver is well documented (1, 2). Exton et al., who earlier located the site of stimulation between pyruvate and PEP synthesis (3), proposed that glucagon stimulated PEP synthesis in the perfused rat liver (4), while reports from Williamson et al. (5) suggested the pyruvate-carboxylase reaction as the site of glucagon action. Stimulation at sites above PEP formation and of portions of the tricarboxylic acid cycle (4) by glucagon have also been suggested (6). In the present experiments, we have used substrates entering at different parts of the gluconeogenic pathway, and specific inhibitors to further resolve the action of glucagon.  相似文献   

14.
The transport of the tricarboxylic acid cycle C(4)-dicarboxylic acids was studied in both the wild-type strain and tricarboxylic acid cycle mutants of Bacillus subtilis. Active transport of malate, fumarate, and succinate was found to be inducible by these dicarboxylic acids or by precursors to them, whereas glucose or closely related metabolites catabolite-repressed their uptake. l-Malate was found to be the best dicarboxylic acid transport inducer in succinic dehydrogenase, fumarase, and malic dehydrogenase mutants. Succinate and fumarate are accumulated over 100-fold in succinic dehydrogenase and fumarase mutants, respectively, whereas mutants lacking malate dehydrogenase were unable to accumulate significant quantities of the C(4)-dicarboxylic acids. The stereospecificity of this transport system was studied from a comparison of the rates of competitive inhibition of both succinate uptake and efflux in a succinate dehydrogenase mutant by utilizing thirty dicarboxylic acid analogues. The system was specific for the C(4)-dicarboxylic acids of the tricarboxylic acid cycle, neither citrate nor alpha-ketoglutarate were effective competitive inhibitors. Of a wide variety of metabolic inhibitors tested, inhibiors of oxidative phosphorylation and of the formation of proton gradients were the most potent inhibitors of transport. From the kinetics of dicarboxylic acid transport (K(m) approximately 10(-4) M for succinate or fumarate in succinic acid dehydrogenase and fumarase mutants) and from the competitive inhibition studies, it was concluded that an inducible dicarboxylic acid transport system mediates the entry of malate, fumarate, or succinate into B. subtilis. Mutants devoid of alpha-ketoglutarate dehydrogenase were shown to accumulate both alpha-ketoglutarate and glutamate, and these metabolites subsequently inhibited the transport of all the C(4)-dicarboxylic acids, suggesting a regulatory role.  相似文献   

15.
High levels of aconitase and fumarase activities were found in Rhodobacter capsulatus E1F1 cells cultured with nitrate as the sole nitrogen source either under light-anaerobic or dark-aerobic conditions. Both activities were strongly and reversibly inhibited in vitro by nitrite or nitric oxide, whereas nitrate or hydroxylamine showed a lower effect. Other enzymes of the tricarboxylic acids cycle such as malate dehydrogenase or isocitrate dehydrogenase were not affected by these nitrogen compounds. When growing on nitrate in the dark R. capsulatus E1F1 cells accumulated nitrite intracellularly, so that an in vivo inhibition of aconitase and fumarase could account for the strong inhibition of growth observed in the presence of nitrite under dark-aerobic conditions.Abbreviations ACO aconitase - FUM fumarase - MDH malate dehydrogenase - ICDH isocitrate dehydrogenase - TCA tricarboxylic acid  相似文献   

16.
The maximum respiration rate of tomato fruit during the climacteric period was markedly increased when the plants were grown under potassium-deficient conditions. Whereas potassium deficiency had no effect on cytoplasmic glutamate-oxoloacetate transaminase, there was a significant increase in the activity of this enzyme from mitochondria once the fruit began to change colour. Malate dehydrogenase was reduced in activity by potassium deficiency. It is suggested that the augmented mitochondrial transiminase levels, coupled with reduced malate dehydrogenase activity in low potassium fruit, result in reduced levels of oxaloacetic acid which is a potent inhibitor of Krebs cycle oxidations, thus leading to higher respiration rates for the intact fruit.  相似文献   

17.
The possible existence of a malonate-sensitive dicarboxylate-mediated electron shuttle between microsomal NAD-linked fatty acid α-oxidation and the mitochondrial electron transport chain in uncoupled fresh potato slices was investigated. Uncoupled slice respiration is inhibited by benzylmalonate and butylmalonate, inhibitors of dicarboxylate transport into mitochondria. Uncoupled slice respiration is also inhibited by rotenone, an indication of intramitochondrial NADH oxidation. Since fatty acid α-oxidation per se is rotenone insensitive, rotenone and benzylmalonate inhibition of the oxidation of carboxyl-labeled myristate in slices points to a dicarboxylic acid shuttle linking microsomal fatty acid a-oxidation with intramitochondrial NADH dehydrogenase.
Malonute inhibits both respiration and 14CO2, release from carboxyl-labeled myristate in fresh uncoupled slices, as do inhibitors of dicarboxylate transport. Mitochondrial studies show that malonate inhibits malate oxidation but not malate dehydrogenase per se. Furthermore, malonate inhibits malate transport more severely than malate oxidation. Accordingly, mulonate inhibition of uncoupled slice respiration in the absence of tricarboxylic acid cycle activity is attributed to its interference with mitochondrial malate transport, and its consequent curtailment of a putative malate-OAA shuttle linked to cytosolic NAD-mediated fatty acid α-oxidation.  相似文献   

18.
The effects of concentrations of 2-phenoxyethanol of negligible bactericidal activity upon the rates of biosynthetic assimilation by Escherichia coli, of 14C-thymidine, 14C-uracil, 14C-phenylalanine and 14C-glucose, were assessed. Increasing the drug concentration from 0.05-0.4% w/v progressively increased inhibition of growth rate, measured as changes in optical density. Thymidine, uracil and glucose assimilation were inhibited to an extent similar to growth rate, whilst phenylalanine incorporation was markedly less sensitive at the lower concentrations (leads to 0.2% w/v). In addition to its previously observed roles in inhibiting oxidative phosphorylation and TCA cycle enzymes, it is suggested that 2-phenoxyethanol can exert a more direct inhibitory action upon DNA and RNA biosynthesis and possibly on protein biosynthesis.  相似文献   

19.
Less than 50% of theoretical oxygen uptake was observed when glucose was dissimilated by resting cells of Pseudomonas natriegens. Low oxygen uptakes were also observed when a variety of other substrates were dissimilated. When uniformly labeled glucose-(14)C was used as substrate, 56% of the label was shown to accumulate in these resting cells. This material consisted, in part, of a polysaccharide which, although it did not give typical glycogen reactions, yielded glucose after its hydrolysis. Resting cells previously cultivated on media containing glucose completely catabolized glucose and formed a large amount of pyruvate within 30 min. Resting cells cultivated in the absence of glucose catabolized glucose more slowly and produced little pyruvate. Pyruvate disappeared after further incubation. In this latter case, experimental results suggested (i) that pyruvate was converted to other acidic products (e.g., acetate and lactate) and (ii) that pyruvate was further catabolized via the tricarboxylic acid cycle. Growth on glucose repressed the level of key enzymes of the tricarboxylic acid cycle and of lactic dehydrogenase. Growth on glycerol stimulated the level of these enzymes. A low level of isocitratase, but not malate synthetase, was noted in extracts of glucose-grown cells. Isocitric dehydrogenase was shown to require nicotinamide adenine dinucleotide phosphate (NADP) as cofactor. Previous experiments have shown that reduced NADP (NADPH(2)) cannot be readily oxidized and that pyridine nucleotide transhydrogenase could not be detected in extracts. It was concluded that acetate, lactate, and pyruvate accumulate under growing conditions when P. natriegens is cultivated on glucose (i) because of a rapid initial catabolism of glucose via an aerobic glycolytic pathway and (ii) because of a sluggishly functioning tricarboxylic acid cycle due to the accumulation of NADPH(2) and to repressed levels of key enzymes.  相似文献   

20.
The composition and properties of the tricarboxylic acid cycle of the microaerophilic human pathogen Helicobacter pylori were investigated in situ and in cell extracts using [1H]- and [13C]-NMR spectroscopy and spectrophotometry. NMR spectroscopy assays enabled highly specific measurements of some enzyme activities, previously not possible using spectrophotometry, in in situ studies with H. pylori, thus providing the first accurate picture of the complete tricarboxylic acid cycle of the bacterium. The presence, cellular location and kinetic parameters of citrate synthase, aconitase, isocitrate dehydrogenase, alpha-ketoglutarate oxidase, fumarate reductase, fumarase, malate dehydrogenase, and malate synthase activities in H. pylori are described. The absence of other enzyme activities of the cycle, including alpha-ketoglutarate dehydrogenase, succinyl-CoA synthetase, and succinate dehydrogenase also are shown. The H. pylori tricarboxylic acid cycle appears to be a noncyclic, branched pathway, characteristic of anaerobic metabolism, directed towards the production of succinate in the reductive dicarboxylic acid branch and alpha-ketoglutarate in the oxidative tricarboxylic acid branch. Both branches were metabolically linked by the presence of alpha-ketoglutarate oxidase activity. Under the growth conditions employed, H. pylori did not possess an operational glyoxylate bypass, owing to the absence of isocitrate lyase activity; nor a gamma-aminobutyrate shunt, owing to the absence of both gamma-aminobutyrate transaminase and succinic semialdehyde dehydrogenase activities. The catalytic and regulatory properties of the H. pylori tricarboxylic acid cycle enzymes are discussed by comparing their amino acid sequences with those of other, more extensively studied enzymes.  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