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 31 毫秒
1.
Summary Citrate synthase (EC 4.1.3.7), aconitate hydratase (EC 4.2.1.3), NADP specific isocitrate dehydrogenase (EC 1.1.1.42), fumarate hydratase (EC 4.2.1.2) and malate dehydrogenase (EC 1.1.1.37) were detected in cell-free preparations of Sclerotinia sclerotiorum (Lib.) D By. grown on liquid glucose-salts medium in stationary culture. Isocitrate lyase (EC 4.1.3.1) was present when the fungus grew on a carbohydrate-free medium but was not detected when the cultures grew on the glucose-salts medium. The amount of oxalate in the culture filtrate declined as the specific activity of citrate synthase and malate dehydrogenase in the mycelium declined. Increasing the initial pH of the medium resulted in an increase of the dicarboxylic acids in the culture filtrate and the specific activity of malate dehydrogenase in the mycelium. The specific reaction(s) leading to oxalic acid formation were not identified.  相似文献   

2.
Citrate Metabolism by Pediococcus halophilus   总被引:3,自引:0,他引:3       下载免费PDF全文
Several strains of non-citrate-metabolizing Pediococcus halophilus have previously been isolated from soy sauce mash or moromi. The factors controlling the metabolism of citrate in soy pediococci were studied. All the soy pediococcal strains tested which failed to decompose citrate did not possess citrate lyase [citrate (pro-3S)-lyase; EC 4.1.3.6] activity. In P. halophilus, citrate lyase was an inducible enzyme, and the optimum pH for activity was 7.0. The metabolism of citrate in P. halophilus was different from that observed in lactic streptococci. The main products from citrate were acetate and formate, and this bacterium produced no acetoin or diacetyl. Formate production from citrate was greatly reduced in the presence of glucose. P. halophilus 7117 (Cit+) was proved to contain citrate lyase, pyruvate formate-lyase (EC 2.3.1.54) phosphotransacetylase (phosphate acetyltransferase; EC 2.3.1.8), and acetate kinase (EC 2.7.2.1), i.e., all the enzymes necessary to convert citrate to acetate and formate.  相似文献   

3.
The objectives of this study were to determine the cause of the crystallization in a large volume creatine supplement solution made from effervescent powders containing di-creatine citrate, and to characterize these crystals using thermal analyses and x-ray diffractometry. Creatine effervescent powders were dissolved in deionized water (pH 6.2) and stored both at room temperature (RT) (25°C) and refrigerated condition (4°C) over a period of 45 days. Creatine concentration was determined using high-performance liquid chromatography (HPLC). Intrinsic dissolution and saturated solubility of creatine, creatine monohydrate, and di-creatine citrate in water were determined and compared. Crystal growth was detected only in the refrigerated samples on the seventh day of storage. Differential Scanning Calorimetry (DSC) and x-ray diffraction (XRD) studies revealed that the crystals formed were of creatine monohydrate. Ninety percent creatine degradation was observed within 45 days for RT samples. However, at refrigerated condition this degradation was 80% within the same time period. The pH of the RT samples also increased from 3.6 to 4.5 during storage. No such increase was observed in the case of refrigerated samples. The intrinsic dissolution rate constants of the compounds decreased in the following order: dicreatine citrate>creatine>creatine monohydrate. In conclusion, di-creatine citrate used in effervescent formulation dissociates to creatine in aqueous solution and eventually crystallizes out as creatine monohydrate. Significant decrease in solubility and effect of pH contribute to this crystallization process.  相似文献   

4.
Activities of five enzymes (pyruvate dehydrogenase complex; citrate synthase, EC 4.1.3.7; carnitine acetyltransferase, EC 2.3.1.7; acetyl-CoA synthetase, EC 6.2.1.1; and ATP citrate lyase, EC 4.1.3.8) were determined in cell bodies of anterior horn cells and dorsal root ganglion cells from the rabbit. For comparison, molecular layer, granular layer and white matter from rabbit and mouse cerebella and cerebral cortex and striatum from the mouse were analyzed. Samples (3–85 ng dry weight) were assayed in 180 to 370 ml of assay reagents containing CoASH and other substrates in excess. By using ‘CoA cycling’, the assay systems were devised to amplify and measure small amounts of acetyl-CoA formed during the enzyme reactions. Carnitine acetyltransferase was the most active enzyme in single nerve cell bodies and all layer samples, except for rabbit and mouse cerebellar white matter. Citrate synthetase was the lowest in single cell bodies. The activities of carnitine acetyltransferase and acetyl-CoA synthetase (656 and 89.8 mmoles of acetyl-CoA formed/kg of dry weight/h at 38°C) from dorsal root ganglion cells were about 2-fold higher than those from anterior horn cells. The activity of ATP citrate lyase (134mmol of acetyl-CoA formed/kg of dry weight/h at 38°C) from anterior horn cells was approximately twice that from dorsal root ganglion cells. The activity of this enzyme was distributed in a wider range in anterior horn cells than dorsal root ganglion cells. The second highest activity (80.0 mmol of acetyl-CoA formed/kg of dry weight/h at 38°C) of ATP citrate lyase was found in striatum where cholinergic interneurones are abundant. Relatively higher activities of this enzyme were found in cerebellar granular layer and white matter which are known to contain the cholinergic mossy fibers. These results suggested that cholinergic neurones contain higher activity of ATP citrate lyase which is thought to supply acetyl-CoA to choline acetyltransferase (EC 2.3.1.6) as a substrate to form acetylcholine.  相似文献   

5.
This paper describes the first measurement of enzyme activities in cluster roots under –Fe stress, at different stages of cluster root development and function. In Lupinus albus L., Cluster roots are produced both under iron- and phosphorus-deficient conditions. In both cases the structure is similar, but the level of exudation is much greater in iron-deficient plants. Much work has been done on the enzyme kinetics of P-deficient cluster roots, but none on enzyme activities of Fe-deficient cluster roots. The enzymes investigated were citrate synthase (EC 4.1.3.7), aconitase (EC 4.2.1.3), isocitrate dehydrogenase [IDH(NAD) (EC 1.1.1.41) and IDH (NADP) (EC 1.1.1.42)] and lactate dehydrogenase (LDH) (EC 1.1.1.27). In cluster roots, citrate synthase activity was initially lower than in lateral roots but, after 5 days, recovered to the lateral root level. Cluster root aconitase levels initially increased, but fell sharply on day 3, and no activity was detected after day 5. IDH (NAD) levels were much lower in cluster roots than in laterals, dropping to a low on day 3, and then rising throughout development. IDH (NADP) levels were always higher in cluster roots than in lateral roots, increasing throughout development. LDH levels in cluster roots fell throughout development. Internal tissue concentrations of citrate were markedly higher in –Fe laterals than in +Fe lateral roots and in cluster roots. Cluster root levels of citrate increased dramatically after day 3. Results are discussed within the context of previous work on enzyme kinetics under –P, and the importance of a block in aconitase activity is highlighted.  相似文献   

6.
Despite highly conserved active-site structures, members of the plant peroxidase superfamily exhibit a wide range of pH optima. Horseradish peroxidase isozyme C (HRPC) is an ideal peroxidase to investigate the structural determinants of pH stability and activity in superfamily members. Conflicting reports exist on the low-pH stability of HRPC and consequently the pKa of the catalytic distal histidine, which is neutral in active peroxidases. Towards resolving such discrepancies, acid-induced changes in HRPC from two popular commercial suppliers were systematically analyzed. Specifically, FTIR v(CO) and Soret-CD spectra of HRPC-CO and Soret absorption of ferric HRPC were recorded to probe time-dependent heme-pocket changes at pH 3.0 in phosphate, citrate and formate buffers, while the FTIR amide I' and far-UV CD spectra were examined to probe changes in secondary structure. Both HRPC-CO samples exhibited identical pH 7.0 v(CO) bands at 1934 and 1905 cm-1. In the pH 3.0 spectrum of sample A, the 1934 cm-1 band was dominant while a broad 1969 cm-1 band appeared in sample B. The intensity of this band, which is assigned to solvent-exposed heme, was greater in citrate than phosphate buffer, but in formate the 1934 cm-1 band remained dominant. Other spectral changes mirrored the v(CO) trends. No time- or buffer-anion-dependent conformation changes were detected in 1 mM CaCl2, revealing that buffer-anion-dependent leaching of stabilizing Ca2+ from HRPC occurs at pH 3.0. Since the N-glycans present in HRPC are of the flexible protein-surface-shielding type, the variation in low-pH conformational stability of the HRPC samples could be attributed to heterogeneous glycosylation, which was detected by SDS-PAGE. It is further proposed that glycosylation patterns may affect the low-pH stability of class II and III plant peroxidases.  相似文献   

7.
Citrate synthase (EC 4.1.3.7), which is present in all living organisms as a key enzyme in aerobic energy metabolism, is one of the most highly phylogenetically conserved enzymes known in terms of its primary and active site structure. However, in terms of other parameters such as in vitro stability, tolerance to changes in pH, degree of self-polymerization, etc., citrate synthases from different sources are markedly different. These divergences can be observed even between isoforms of the enzyme within the same species. Data documenting these diversities suggest that a high degree of difference in tertiary structures may occur. Therefore, the surface profiles of citrate synthase enzymes from yeast, pig, rat, tomato and Escherichia coli were investigated with immunological methods using monoclonal antibody families generated against either pig citrate synthase (alpha-PCS) or yeast citrate synthase-2 (alpha-YCS-2). A high degree of homology of enzyme epitopes was detected on the mitochondrial citrate synthases originating from yeast, tomato, pig and rat cells. Major differences were found between the hexameric citrate synthase originating from E. coli compared with those dimeric forms prepared from eukaryotic cells. Only modest similarities were detected between the highly homologous peroxisomal and mitochondrial yeast citrate synthases. Furthermore, a point mutation of one of the catalytic residues (H274R on recombinant pig and H313R on yeast enzyme) of mitochondrial citrate synthase (CS-1) resulted in a significant increase in immunological similarity with the peroxisomal isoenzyme (CS-2). These findings are discussed in terms of the possible mechanism of evolution of CS-2 in yeast.  相似文献   

8.
The effect of pH on chorismate mutase/prephenate dehydratase (chorismate pyruvate mutase/prephenate hydro-lyase (decarboxylating) EC 5.4.99.5/EC 4.2.1.51) from Escherichia coli K12 has been studied. While the maximum velocity of both activities is independent of pH, Km for chorismate or prephenate shows a complex pH dependence. Differences in mutase activity in acetate/phosphate/borate and citrate/phosphate/borate buffers were traced to inhibition by citrate. When a variety of analogues of citrate were tested as possible inhibitors of the enzyme, several were found to inhibit mutase and dehydratase activities to different extents, and by different mechanisms. Thus citrate competitively inhibits mutase activity, but inhibits dehydratase activity by either a non-competitive or an uncompetitive mechanism. Conversely, cis- and trans-aconitate competitively inhibit dehydratase activity, but are partially competitive inhibitors of mutase activity. The differential effects of these inhibitors on the two activities are consistent with the existence of two distinct active sites, but additionally suggest some degree of interconnection between them. The implications of these results for possible mechanisms of catalysis by chorismate mutase/prephenate dehydratase are discussed.  相似文献   

9.
The mitochondrial matrix subfractions from rat liver, kidney cortex, brain, heart, and skeletal muscle were isolated and their protein components were resolved by two-dimensional polyacrylamide gel electrophoresis, revealing between 120 and 150 components for each matrix subfraction. Excellent resolution was obtained utilizing a pH 5 to 8 gradient in the first dimension and in 8 to 13% exponential acrylamide gradient in the second dimension, increasing the number of mitochondrial matrix proteins observed 3-fold over one-dimensional systems. Protein components tentatively identified by co-migration with pure enzymes and by known tissue distributions are carbamoyl-phosphate synthetase (EC 2.7.2.5), ornithine transcarbamylase (EC 2.1.3.3), glutamate dehydrogenase (EC 1.4.1.3), pyruvate carboxylase (EC 6.4.1.1), citrate synthase (EC 4.1.3.7), fumarase (EC 4.2.1.2), aconitase (EC 4.2.1.3), alpha-ketoglutarate dehydrogenase (EC 1.2.4.2), dihydrolipoyl transsuccinylase (EC 2.3.1.12), lipoamide dehydrogenase (EC 1.6.4.3), glutamate-aspartate aminotransferase (EC 2.6.1.1), and the two subunits of pyruvate dehydrogenase (EC 1.2.4.1). Protein components unambiguously identified by peptide mapping are citrate synthase, aconitase, and pyruvate carboxylase. The inner membrane subfraction from rat liver mitochondria was also resolved two dimensionally; the alpha and beta subunits of ATPase (F1) (EC 3.6.1.3) were identified by peptide mapping.  相似文献   

10.
The mitochondrial and glyoxysomal citrate synthase (EC 4.1.3.7) from the endosperm of germinating castor beans ( Ricinus communis L., type Sanzibaricnsis) were purified to a final specific activity of 76 and 78 U (mg protein)−1, respectively. Both citrate synthases could be bound to ATP-Sepharose. However, only the mitochondrial enzyme could be eluted by either 100 μ M oxaloacetate or 100 μ M coenzyme A (indicative of affinity chromatography), while the glyoxysomal enzyme was only eluted by 0.5 M KCI (indicative of ion-exchange chromatography). Many properties of the two isoenzymes were similar including the pH dependence and temperature dependence of activity, the pH stability, and the inactivation of the enzyme at elevated temperatures. The most pronounced differences between the two citrate synthases were the isolelectric points of pH 5.9 for the mitochondrial and of pH 9.1 for the glyoxysomal enzyme. Both citrate synthases are dimers in the native form with a molecular weight of 95000 each, as determined by gel filtration on Sepharose CL-6B and by polyacrylamide gel electrophoresis in the presence of 0.1% sodium dodecyl sulfate. However, the glyoxysomal citrate synthase existed also as a tetramer with a molecular weight of 200000 in the presence of 10 m M MgCl2.  相似文献   

11.
Aconitase (citrate(isocitrate)hydro-lyase, EC 4.2.1.3) prior to activation demonstrates a single binding site for substrates and inhibitors. On the basis of kinetic experiments, at pH 8.5 and 37 degrees C, with monomeric butanedione in borate, this binding site was found to contain a single arginine residue. Dissociation constants at pH 8.5 and 37 degrees C, determined from inhibitory effects on butanedione inactivation rates are: citrate, 0.74 mM; D-isocitrate, 0.33 mM: cis-aconitate, 0.52 mM; tricarballytate, 0.42 mM; trans-aconitate, 0.025 mM. Corresponding dissociation constants for the active enzyme are: tricarballylate, 0.39 mM; trans-aconitate, 0.14 mM. Active site Fe2+ added to the enzyme on activation is therefore not required for binding. Km values are: citrate, 0.23 mM and cis-aconitate 0.012 mM. Binding to active enzyme is considered to be transition state binding.  相似文献   

12.
The cepA putative gene encoding a cellobiose phosphorylase of Thermotoga maritima MSB8 was cloned, expressed in Escherichia coli BL21-codonplus-RIL and characterized in detail. The maximal enzyme activity was observed at pH 6.2 and 80 degrees C. The energy of activation was 74 kJ/mol. The enzyme was stable for 30 min at 70 degrees C in the pH range of 6-8. The enzyme phosphorolyzed cellobiose in an random-ordered bi bi mechanism with the random binding of cellobiose and phosphate followed by the ordered release of D-glucose and alpha-D-glucose-1-phosphate. The Km for cellobiose and phosphate were 0.29 and 0.15 mM respectively, and the kcat was 5.4 s(-1). In the synthetic reaction, D-glucose, D-mannose, 2-deoxy-D-glucose, D-glucosamine, D-xylose, and 6-deoxy-D-glucose were found to act as glucosyl acceptors. Methyl-beta-D-glucoside also acted as a substrate for the enzyme and is reported here for the first time as a substrate for cellobiose phosphorylases. D-Xylose had the highest (40 s(-1)) kcat followed by 6-deoxy-D-glucose (17 s(-1)) and 2-deoxy-D-glucose (16 s(-1)). The natural substrate, D-glucose with the kcat of 8.0 s(-1) had the highest (1.1 x 10(4) M(-1) s(-1)) kcat/Km compared with other glucosyl acceptors. D-Glucose, a substrate of cellobiose phosphorylase, acted as a competitive inhibitor of the other substrate, alpha-D-glucose-1-phosphate, at higher concentrations.  相似文献   

13.
1. In order to assess whether the potential ability of heart ventricular muscle and liver to metabolise substrates such as alanine, aspartate and lactate varies as the sheep matures and its nutrition changes, the activities of the following enzymes were determined in tissues of lambs obtained at varying intervals between 50 days after conception to 16 weeks after birth and in livers from adult pregnant ewes: lactate dehydrogenase (EC 1.1.1.27), alanine aminotransferase (EC 2.6.1.2), pyruvate kinase (EC 2.7.1.40), pyruvate carboxylase (EC 6.4.1.1), phosphoenolpyruvate carboxykinase (GTP)(EC 4.1.1.32), malate dehydrogenase (EC 1.1.1.37), aspartate aminotransferase (EC 2.6.1.1) and citrate (si)-synthase (EC 4.1.3.7). 2. In the heart a most marked increase in alanine aminotransferase activity was found throughout development. During this period the activities of citrate (si)-synthase, lactate dehydrogenase and pyruvate carboxylase also increased. There were no substantial changes in the activities of aspartate aminotransferase, malate dehydrogenase or pyruvate kinase. Pyruvate kinase activities were five times greater in the heart compared with those found in the liver. No significant activity of phosphoenolpyruvate carboxykinase (GTP) was detected in heart muscle. 3. In the liver the activities of both alanine aminotransferase and aspartate aminotransferase increased immediately following birth although the activity of alanine aminotransferase was lower in livers of pregnant ewes than in any of the lambs. As with alanine aminotransferase the highest activities of lactate dehydrogenase were found during the period of postnatal growth. No marked changes were observed in malate dehydrogenase or citrate (si)-synthase activities during development. A small decline in pyruvate kinase activity occurred whilst the activities of pyruvate carboxylase and phosphoenolpyruvate carboxykinase (GTP) tended to rise during development.  相似文献   

14.
A procedure is described for the purification of phosphoenolpyruvatecarboxylase (EC 4.1.1.31 [EC] ) and NADP-dependent malic enzyme (EC1.1.1.40 [EC] ) from sugar cane leaves. Each enzyme was purified tohomogeneity as judged by sodium dodecyl sulfate-polyacrylamidegel electro-phoresis, with about 30% yield. Phosphoenolpyruvatecarboxylase was purified 54-fold. A molecular weight of 400,000and a homotetrameric structure were determined for the nativeenzyme. The purified carboxylase had a specific activity of20.0 {diaeresis}mol (mg protein)–1 min–1, and wasactivated by glucose-6-phosphate and inhibited by L-malate.Km values at pH 8.0 for phosphoenolpyruvate and bicarbonatewere 0.25 and O.l0 mM, respectively. NADP-malic enzyme, 356-foldpurified, exhibited a specific activity of 71.2 {diaeresis}mol(mg protein)–1 min–1 and was characterized as ahomotetramer with native molecular weight of 250,000. Purifiedmalic enzyme showed an absolute specificity for NADP+ and requireda divalent metal ion for activity. Km values of 0.33 and 0.008mM for L-malate and NADP+, respectively, were determined. Thisenzyme was inhibited by several organic acids, including ketoand amino acids; while succinate and citrate increased the enzymeactivity when assayed with 10{diaeresis}M L-malate. The effectsshown by amino acids and by citrate were dependent on pH, beinghigher at pH 8.0 than at pH 7.0. (Received October 26, 1988; Accepted February 3, 1989)  相似文献   

15.
Intercept inhibition of rabbit-muscle phosphoglucomutase (alpha-D-glucose-1,6-bisphosphate: alpha-D-glucose-1-phosphate phosphotransferase, EC 2.7.5.1) produced by several nucleotide diphosphates and compounds related to coenzyme A was re-examined in order to re-evaluate an earlier suggestion that this enzyme has an allosteric regulatory site. However, in all cases intercept inhibition constants were much larger than those previously reported, and in all but two cases were too large to assess in the assay system, i.e., were greater than 10 mM. Most of the intercept inhibition previously observed apparently was caused by the use of the Li+ salts of inhibitors. Thus, Li+ binds competitively with the natural activator, Mg2+, and in the presence of glucose phosphates binds almost as well as Mg2+: Kd approximately 10 micrometer. The observation that glucose phosphates bind to the Li+ complex of phosphoglucomutase some 900 times more tenaciously than to the corresponding Mg2+ complex could provide a partial rationale for the lack of reactivity of the Le+ form of the enzyme. Attempts to verify the dimeric structure of phosphoglucomutase that was previously reported also produced negative results.  相似文献   

16.
广州市流溪河降水离子浓度与电导率关系   总被引:2,自引:0,他引:2  
通过对2006年7月~2008年5月在流溪河地区采集的106次降雨水样(雨量范围0.7~268.6 mm)进行化学测试,研究了降水中几类主要离子含量与电导率的关系。结果显示:(1)通过各化学离子含量计算的电导率与对应实测的电导率之间存在-2.94%的平均相对误差。(2)H+离子浓度与电导率之间的相关关系,在水样中pH<5.6时是正相关,在水样中pH≥5.6时呈现负相关规律。(3)电导率随离子总浓度的增加而增加,但当水样的总离子浓度相同或者H+和总离子浓度相同时,由于样品之间存在显著的其他化学组成差异,致使电导率的不一样。(4)将水样分成pH≥5.6和pH<5.6的两组后,每组中各离子浓度与电导率之间存在更好的回归关系,因此可以通过pH值、电导率的测定并利用回归方程来预测水中的其他化学成分含量。  相似文献   

17.
Carbon-14 was incorporated from citrate-1,5-14C, glyoxylate-14C(U), or glyoxylate-1-14C into oxalate by cultures of Aspergillus niger pregrown on a medium with glucose as the sole source of carbon. Glyoxylate-14C(U) was superior to glyoxylate-1-14C and citrate-1,5-14C as a source of incorporation. By addition of a great amount of citrate the accumulation of oxalate was accelerated and its maximum yield increased. In a cell-free extract from mycelium forming oxalate from citrate the enzyme oxaloacetate hydrolase (EC3.7.1.1) was identified. Its in vitro activity per flask exceeded the rate of in vivo accumulation of oxalate. Glyoxylate oxidizing enzymes (glycolate oxidase, EC1.1.3.1; glyoxylate oxidase, EC1.2.3.5;NAD(P)-dependent glyoxylate dehydrogenase; glyoxylate dehydrogenase, CoA-oxalylating, EC1.2.1.7) could not be detected in cell-free extracts. It is concluded that in cultures accumulating oxalate from citrate after pregrowth on glucose, oxalate arises by hydrolytic cleavage of oxaloacetate but not by oxidation of glyoxylate.  相似文献   

18.
Heat-stable fructose 1,6-bisphosphate-dependent L-lactate dehydrogenase [EC 1.1.1.27] was purified from an extremely thermophilic bacterium, Thermus aquaticus YT-1. The amino acid composition and NH2-terminal 34 amino acid sequence of the enzyme were determined. Its NH2-terminal sequence shows high homology with those of Thermus caldophilus GK24 (82% identity) and some other bacterial L-lactate dehydrogenases (44-53% identity), indicating the close phylogenic relationship of the two Thermus species. At the same time, the two Thermus L-lactate dehydrogenases were found not to be identical not only chemically but also kinetically and immunologically. Citrate activated the T. aquaticus enzyme in the weak acidic pH region, while fructose 1,6-bisphosphate did in both acidic and neutral pH regions. The maximum activity obtained with citrate at pH 5.0 was about 2.5 times higher than that in the presence of fructose 1,6-bisphosphate at pH 6.7. The enzymes modified with 2,3-butanedione, acetic anhydride and diethyl pyrocarbonate in the presence of both NADH and oxamate were desensitized to fructose 1,6-bisphosphate, and the modified enzymes were active even in the absence of fructose 1,6-bisphosphate. All of the modified enzymes examined were still activated by citrate similarly to the native enzyme. These results suggest that the mechanism of activation by citrate is different from that by fructose 1,6-bisphosphate, and that the citrate-binding site is different from the fructose 1,6-bisphosphate-binding site.  相似文献   

19.
The cytosolic fumarase [EC 4.2.1.2[ of rat liver was bound, after dialysis, to the microsomal membrane in vitro. Binding of the enzyme was dependent on pH, and was facilitated in the pH range below 7.5. The binding reaction was completely inhibited by 0.5 mM fumarate, aurintricarboxylate or colchicine. The bound fumarase was released from the membrane by the substrates, isocitrate, citrate or 2,3-diphosphoglycerate at low concentrations. Desorption of the enzyme by metabolites was also dependent on pH, and was more rapid in the alkaline pH range. The enzyme desorption curves were sigmoidal, and kinetic studies suggested a biphasic cooperative mechanism for the action of the metabolites. The apparent desorption constants (concentrations necessary for 50% desorption of the enzyme) estimated at pH 7.3 for isocitrate, 2,3-diphosphoglycerate, L-malate, oxalacetate, fumarate, citrate, succinate, and KCl were 0.073, 0.074, 0.22, 0.39, 0.56, 2.9, and 19 mM, respectively. The bound fumarase showed little enzymatic activity, and its Km and Vmax values were fivefold and 31%, respectively, of those of the free enzyme.  相似文献   

20.
Carbon-14 was incorporated into oxalate and CO2 from either citrate-1,5-14C, succinate-1,4-14C, or fumarate-1,4-14C by cultures of Aspergillus niger pregrown on a medium which contained glucose as the sole carbon source and which did not allow citrate accumulation. In cell-free extracts of mycelium forming oxalate and CO2 from added citrate the following enzymes of the tricarboxylic acid (TCA) cycle were identified: citrate synthase CE 4.1.3.7), aconitate hydratase (EC4.2.1.3), NAD and NADP-dependent isocitrate dehydrogenase (EC 1.1.1.41, 1.1.1.42), (alpha-oxoglutarate dehydrogenase (EC 1.2.4.2), succinate dehydrogenase (EC 1.3.99.1), fumarate hydratase (EC 4.2.1.2), and malate dehydrogenase (EC 1.1.1.37). The in vitro activity of aconitate hydratase and of NADP-dependent isocitrate dehydrogenase was shown to be almost identical to the rate of in vivo degradation of citrate or to exceed this rate. The degradation of citrate to oxalate was inhibited completely by 9 mM fluoroacetate. It is concluded that the TCA cycle is involved in the formation of oxalate from citrate.  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