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.
  • 1.1. Malate dehydrogenase has been purified from the foot muscle of Patella caerulea by ion-exchange chromatography on DEAE-cellulose, affinity chromatography on Blue Agarose and gel filtration on Sephadex G-150.
  • 2.2. The yield was 23.5% of the initial activity with a final specific activity of 257 U/mg of protein.
  • 3.3. The apparent mol. wt of the native enzyme is approx. 75,000 and it consists of two subunits of mol. wts in the range of 36,000–39,000.
  • 4.4. The enzyme exhibits hyperbolic kinetics with respect to oxaloacetate, NADH and l-malate. The Km values were determined to be 0.055 mM for oxaloacetate, 0.010 mM for NADH and 0.37 mM for l-malate. The pH optima are around 8.4 for the reduction of oxaloacetate and 9.2–9.6 for the reduction of oxaloacetate and 9.2–9.6 for the l-malate oxidation. Vmax and Km values for oxaloacetate change in an opposite manner with respect to pH values.
  • 5.5. Of the various compounds tested, only α-ketoglutarate, citrate and adenylate phosphates were found to inhibit the enzyme activity.
  • 6.6. From the above properties it appears that the reaction of cytoplasmic malate dehydrogenase of P. caerulea foot muscle is a key reaction in the anaerobic pathway and it occurs with the production of malate.
  相似文献   

2.
Lactate and malate dehydrogenases (EC 1.1.1.27 and EC 1.1.1.37, respectively) were precipitated with ammonium sulfate, redissolved in 100 mM phosphate buffer, and the kinetic parameters of each enzyme determined. Lactate dehydrogenase: The enzyme preparation had a specific activity of 0.35 μmole NADH oxidized/min/mg protein for pyruvate reduction, and 0.10 μmole NAD reduced/min/mg protein for lactate oxidation. Km values for the substrates and cofactors were as follows: pyruvate = 0.51, mM; lactate = 3.8 mM; NADH = 0.011 mM; and NAD = 0.17 mM. NADPH, NADP, or d(?)-lactate would not replace NADH, NAD, or l(+)-lactate, respectively. The enzyme was relatively stable at 50 C for 45 min, but much less stable at 60 C; repeated freezing and thawing of the enzyme preparation had little effect on LDH activity. Both p-chloromercuribenzoate (p-CMB) and N-ethylmaleimide (NEM) significantly inhibited LDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least two LDH isoenzymes in the unpurified enzyme preparation. The molecular weight was estimated at 160,000 by gel chromatography. Malate dehydrogenase: The enzyme preparation had a specific activity of 6.70 μmole NADH oxidized/min/mg protein for oxaloacetate reduction, and 0.52 μmole NAD reduced/ min/mg protein for malate oxidation. Km values for substrates and cofactors were as follows: l-malate = 1.09 mM; oxaloacetate = 0.0059 mM; NADH = 0.017 mM; and NAD = 0.180 mM. NADP and NADPH would not replace NAD and NADH, respectively, d-malate was oxidized slowly when present in high concentrations (>100 mM). Significant substrate inhibition occurred with concentrations of l-malate and oxaloacetate above 40 mM and 0.5 mM, respectively. The enzyme was unstable at temperatures above 40 C, but repeated freezing and thawing of the enzyme preparation had little effect on MDH activity. Only p-CMB inhibited MDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least three MDH isoenzymes in the unpurified enzyme preparation, and the molecular weight was estimated at 49,000 by gel chromatography.  相似文献   

3.
Aspartate β-decarboxylase (AspD), which catalyses the β-decarboxylation of aspartate (Asp) to alanine (Ala), was found in significant quantities only in the brain, kidney and liver. This enzyme has an optimum pH at 7.4. Addition of exogenous pyridoxal 5′-phosphate did not increase enzyme activity presumably because of firmly bound cofactor. However, aminooxyacetic acid is a potent inhibitor.There is an apparent 8-fold variation in AspD in the seven brain regions studied, with the highest activities in the cortex and the lowest in the striatum and hippocampus. In the presence of α-ketoglutarate, the production of 14CO2 from [14C]Asp may no longer represent AspD activity due to active transamination of Asp, presumably by aspartate aminotransferase, to oxaloacetate. Under such conditions, comparable AspD activities were observed in all seven brain regions.Kinetic analysis showed that the liver and kidney enzymes have identical affinity for Asp (Km = 3.5 mM) while the brain enzyme has a higher affinit (Km = 1.3 mM). The Vmax values obtained indicated that the enzyme populations in liver, kidney and brain are in the ratio 18:4:1. Various amino acids were found to inhibit both brain and liver AspD. Serine, however, activated the liver enzyme but inhibited competitively the kidney and brain enzymes. These results indicate that AspD may exist as two or more isozymes.  相似文献   

4.
Mitochondria isolated from pea (Pisum sativum L.) leaves are able to transport the keto acid, oxaloacetate, from the reaction medium into he mitochondrial matrix at high rates. The rate of uptake by the mitochondria was measured as the rate of disappearance of oxaloacetate from the reaction medium as it was reduced by matrix malate dehydrogenase using NADH provided by glycine oxidation. The oxaloacetate transporter was identifed as being distinct from the dicarboxylate and the α-ketoglutarate transporters because of its inhibitor sensitivities and its inability to interact with other potential substrates. Phthalonate and phthalate were competitive inhibitors of oxaloacetate transport with Ki values of 60 micromolar and 2 millimolar, respectively. Butylmalonate, an inhibitor of the dicarboxylate and α-ketoglutarate transporters, did not alter the rate of oxaloacetate transport. In addition, a 1000-fold excess of malate, malonate, succinate, α-ketoglutarate, or phosphate had little effect on the rate of oxaloacetate transport. The Km for the oxaloacetate transporter was about 15 micromolar with a maximum velocity of over 500 nanomoles per milligram mitochondrial protein/min at 25°C. No requirement for a counter ion to move against oxaloacetate was detected and the highest rates of uptake occurred at alkaline pH values. An equivalent transporter has not been reported in animal mitochondria.  相似文献   

5.
Properties of Ornithine Carbamoyltransferase from Pisum sativum L   总被引:4,自引:3,他引:1       下载免费PDF全文
Some properties of ornithine carbamoyltransferase from chloroplasts isolated from leaves of Pisum sativum L. (cv Marzia) were compared with those of the enzyme partially purified (316-fold) from shoots of seedlings after 3 weeks of cultivation.

Both preparations showed a pH optimum at pH 8.3 and had the same affinity to ornithine (Km = 1.2 millimolar) as well as to carbamoyl phosphate (Km = 0.2 millimolar). The approximate molecular weight determined by gel sieving was 77,600.

A desalted ammonium sulfate precipitate from 14-day seedlings (inclusive roots and senescing cotyledons) was applied on a column of anion exchanger. The elution pattern showed one peak of ornithine carbamoyl-transferase activity. This elution pattern was the same as observed for the enzyme from chloroplasts.

The results suggest the presence of one form of ornithine carbamoyl-transferase in pea seedlings.

  相似文献   

6.
Ornithine carbamoyltransferase (EC 2.1.3.3) activity was detected in apple (Pyrus malus L.) leaf tissue from early June to November. Total activity remained relatively constant at 4.1 μmoles citrulline produced per hour per 10 cm2 until mid-October when it sharply doubled. Following the first frost of the autumn, the enzyme lost about 80% of its former activity. The enzyme from apple leaf exhibited two pH optima, one at pH 8.6 and the other at pH 7.8, indicating the presence of isozymes or two forms of the enzyme. At pH 8.6, a partially-purified enzyme preparation had binding contrasts for its substrates of 6 mm for carbamyl-phosphate and 4.8 mm for ornithine. At pH 7.8, the Km for carbamyl-phosphate was 1.9 mm and the Km for ornithine was 1.22 mm.  相似文献   

7.
In a previous study we showed that ornithine aminotransferase (OAT) exhibits concentration-dependent self-association in two stages (45,000 Mr monomers aggregate to form 140,000–150,000 Mr trimers; the trimers then aggregate to form higher-molecular-weight complexes). In an attempt to characterize further the molecular mechanisms involved in OAT aggregation, the present study examined the effects of basic amino acids and keto acids on the aggregation process. These experiments showed that basic amino acids (ornithine and lysine) inhibit the association of monomers to form trimers, apparently by interaction with carboxyl groups on the surfaces of the monomers. The aggregation of trimers to form higher-molecular-weight assemblies is not affected by basic amino acids, and neither aggregation stage is affected by the keto acids, α-ketoglutarate, or oxaloacetate. We also found that two different OAT preparations (one fresh, the other 18 months old) differed in aggregation characteristics; the older preparation showed reduced self-affinity at both aggregation stages, but both preparations had similar catalytic efficiencies. Electrophoretic studies indicated that the older preparation contained variants of the enzyme monomer with greater electronegativity than did the fresh preparation. We conclude, therefore, that OAT purification exposes ionically labile but catalytically insignificant domains on the monomer surface, and the loss of positively charged groups from such regions diminishes the OAT aggregation potential.  相似文献   

8.
The incorporation of 15N into washed cells of Derxia gummosa from labelled-(NH4)2SO4 and -KNO3 respectively was inhibited by both L-methionine-DL-sulphoximine and azaserine. Glutamine synthetase purified to homogeneity from this bacterium had a molecular weight of 708 000 and was composed of 12 similar subunits each of 59 000. The enzyme assayed by γ-glutamyltransferase method had Km values for L-glutamine and hydroxylamine of 12.5 and 1.2 mM, respectively. Optimal pH values for adenylylated and deadenylylated forms were pH 7.0 and pH 8.0, respectively. The adenylylated enzyme was deadenylylated by treatment with snake venom phosphodiesterase. The inhibitions by both glutamate and ammonia were competitive. The activity was markedly inhibited by L-methionine-DL-sulphoximine, alanine, glycine and serine and to a lesser extent by aspartate, phenylalanine and lysine. Various tri-, di- and mono-phosphate nucleotides, organic acids (pyruvate, oxalate and oxaloacetate) were also inhibitory. Glutamate synthase purified 167-fold had specific requirements for NADH, L-glutamine and 2-ketoglutarate. The Km values for NADH, glutamine and 2-ketoglutarate were 9.6, 270 and 24 μM respectively. Optimal pH range was 7.2–8.2. The enzyme was inhibited by azaserine, methionine, aspartate, AMP, ADP and ATP.  相似文献   

9.
A study on the response of the stability and activity of crystalline ox liver nuclear and mitochondrial glutamate dehydrogenases to temperature variations has been carried out. The thermodynamic properties of the heat inactivation process and of the reaction with the substrates glutamate and α-ketoglutarate have been investigated. The heat inactivation of nuclear glutamate dehydrogenase proceeds at a faster rate than that of the mitochondrial enzyme in the temperature range 40–51 °C; the enthalpy of activation of the inactivation process is higher and the entropy is almost double, compared to the values of mitochondrial glutamate dehydrogenase. The effect of temperature on the maximal velocity shows that, with both glutamate and α-ketoglutarate, the enthalpy of activation with nuclear glutamate dehydrogenase is double and the decrease in entropy almost half of the values of the mitochondrial enzyme. The variation of the apparent Km with temperature shows a decrease of the affinity of both enzymes for glutamate, with no major difference in the thermodynamic properties of the reaction. With α-ketoglutarate, on the other hand, the affinity of nuclear glutamate dehydrogenase decreased, whereas that of the mitochondrial enzyme increased with temperature. The process is therefore exothermic with the former enzyme, endothermic with the latter; furthermore, it occurs with a decrease in enthropy with nuclear glutamate dehydrogenase, but with a large increase with the mitochondrial enzyme. The studies on the effect of temperature on the activity were carried out in the range 20–44 °C.  相似文献   

10.
《Experimental mycology》1989,13(3):294-298
Succinate dehydrogenase (EC 1.3.99.1) fromDictyostelium discoideum was purified 40-fold. The pH optimum for the reaction underin vitro conditions was 7.4. Divalent cations showed no effect on the enzyme activity. Lineweaver-Burk plots of initial velocity data were linear. The Km value for succinate was calculated to be 0.22 mM. Apparent Ki values for fumarate, malonate, and oxaloacetate were 0.4, 0.02, and 0.003 mM, respectively. All three showed a competitive inhibition pattern. A comparison of the reaction ratein vivo with the calculated enzyme activity requiredin vivo (Vv) suggests that succinate dehydrogenase may be rate controlling to flux through the citric acid cycle.  相似文献   

11.
Flavanone 3β-hydroxylase plays very important role in the biosynthesis of flavonoids. A putative flavanone 3β-hydroxylase gene (Pef3h) from Populus euphratica was cloned and over-expressed in Escherichia coli. Induction performed with 0.1 mM IPTG at 20°C led to localization of PeF3H in the soluble fraction. Recombinant enzyme was purified by Ni-NTA affinity. The optimal activity of PeF3H was revealed at pH 7.6 and 35°C. The purified enzyme was stable over pH range of 7.6–8.8 and had a half-life of 1 h at 50°C. The activity of PeF3H was significantly enhanced in the presence of Fe2+ and Fe3+. The K M and V max for the enzyme using naringenin as substrate were 0.23 mM and 0.069 μmoles mg–1min-1, respectively. The K m and V max for eriodictyol were 0.18 mM and 0.013 μmoles mg–1min–1, respectively. The optimal conditions for naringenin bioconversion in dihydrokaempferol were obtained: OD600 of 3.5 for cell concentration, 0.1 mM IPTG, 5 mM α-ketoglutaric acid and 20°C. Under the optimal conditions, naringenin (0.2 g/L) was transformed into 0.18 g/L dihydrokaempferol within 24 h by the recombinant E. coli with a corresponding molar conversion of 88%. Thus, this study provides a promising flavanone 3β-hydroxylase that may be used in biosynthetic applications.  相似文献   

12.
Cysteine sulfinate transaminase (E.C. 2.6.1,l-cysteine sulfinate:2 oxoglutarate aminotransferase) catalyzes the conversion of cysteine sulfinate and α-ketoglutarate to 3-sulfonyl pyruvate and glutamate. A simple two-step assay has been developed to measure the enzyme activity in the high speed supernatant of whole brain homogenate. In the first step, the supernatant is incubated in the presence of exogenous substrate, then glutamate dehydrogenase is added to catalyze the conversion of glutamate to α-ketoglutarate, and the concomitant production of NADH is fluorimetrically monitored. The apparent Km values of cysteine sulfinate transaminase for cysteine sulfinate and α-ketoglutarate are 1.24 and 0.22 mm, respectively. This assay is extremely rapid and has a high sensitivity, samples containing as low as 30 ng of protein may be accurately assayed.  相似文献   

13.
The specific activity of dihydroorotate dehydrogenase, catalysing the conversion of l-5,6-dihydroorotate (l-DHO) to orotate, in Leishmania mexicana mexicana was found to be 22.1 ± 3.5 nmole/hr/mg protein in the amastigote, and 28.7 ± 4.6 nmole/hr/mg protein in the promastigote. The enzyme was found to be soluble and to require molecular O2 for activity in both forms of the parasite. Oxygen utilisation was not mediated through the mitochondrial cytochrome-containing respiratory chain, and phenazine methosulphate and ferricyanide could be used as electron acceptors by the enzyme in both aerobic and anaerobic conditions. The enzyme from both amastigote and promastigote had a pH optimum of 7.0, and the Km values for l-DHO were 11.8 ± 4.9 and 2.3 ± 0.4 μM, respectively. The pyrimidine analogs 5-methylorotate (Ki = 8.8 μM) and 5-aminoorotate (Ki = 57 μM) were shown to be competitive inhibitors of the promastigote enzyme, as was the reaction product orotate (Ki = 10 μM).  相似文献   

14.
Soluble and mitochondrial malic dehydrogenases (MDH) were isolated from root tips of the halophyte Tamarix tetragyna L. grown in the presence and absence of NaCl. The activity of the enzymes isolated from root tips grown in the presence of NaCl was lower than that of the enzymes isolated from roots grown in absence of NaCl. The mitochondrial MDH was much more sensitive to salinity than the soluble MDH. The soluble enzyme from roots grown in NaCl had a higher Km for malate and lower Km for NAD than enzyme from the control roots. Addition of NaCl in vitro at 72 mM significantly stimulated the reductive activity of soluble MDH, while higher NaCl concentrations (240 mM and above) depressed enzyme activity. The inhibition of enzyme activity by various salts was found to be in the order MgCl2 > NaCl = KCl > Na2SO4. Mannitol at equiosmotic concentrations had no effect. Substrate inhibition, typical for oxaloacetate oxidation, was not observed at high NaCl concentrations in vitro and high substrate concentrations neutralized the inhibitory effect of NaCl. Increased coenzyme concentrations had no effect. In vitro NaCl increased the Km for malate and oxaloacetate already at relatively low concentrations. At the same time NaCl decreased the Km for NAD and NADH. The inhibitory effect of NaCl on enzyme activity seems not to be due to the effect on the Km alone. Soluble and mitochondrial MDH had different responses to pH changes, mitochondrial MDH being more sensitive. Mitochondrial MDH released from the particles had a similar response to that of the entire particles. Changes of pH modified the effect of NaCl on enzyme activity. It was postulated that NaCl apparently induces conformational changes in the enzyme.  相似文献   

15.
NADPH-cytochrome c reductase of vitamin D3-deficient chick kidney mitochondria has been purified approximately 1100-fold to a specific activity of 788 nmol cytochrome c reduced/min/mg protein. Analytical gel electrophoresis of the purified enzyme revealed two bands when stained for protein, but only the more anodic band demonstrated NADPH-tetrazolium reductase activity. The relative ease of solubilization of the reductase without the use of lipases, proteases, or detergents was the first line of evidence that suggested a novel mitochondrial localization for this previously unreported NADPH-linked cytochrome c reductase. The apparent properties of the reductase suggest that the enzyme is a component of kidney mitochondrial outer membrane. The kinetic determination of Michaelis constants with respect to NADPH, cytochrome c, and NADH gave the following values: KmNADPH = 1.7 μM, Kmcytc = 3.4 μM, and KmNADH = 20 mM. These constants were different from those of the intact kidney microsomal reductase: KmNADPH = 5.5 μM, Kmcytc = 10.5 μM, and KmNADH = 13.3 μM. The mitochondrial as well as the intact microsomal reductases exhibited a ping-pong kinetic mechanism for NADPH-mediated cytochrome c reduction. Spectrofluorometric measurements demonstrated the presence of equimolar amounts of FAD and FMN. The oxidized enzyme has absorption maxima at 280 and 450 nm with a shoulder at 415 nm. Upon reduction with NADPH a distinct loss in the 450-nm absorption was observed. Ouchterlony immunodiffusion studies with rabbit antiserum to chick renal mitochondrial ferredoxin did not reveal cross-reactivity when the purified reductase was the antigen. This excludes the involvement of a ferredoxin-type iron-sulfur protein in the NADPH-mediated reduction of cytochrome c by the purified reductase.  相似文献   

16.
Glutamate metabolism triggered by oxaloacetate in intact plant mitochondria   总被引:6,自引:0,他引:6  
In Percoll-purified potato tuber mitochondria, glutamate metabolism can be triggered by oxaloacetate, in the presence of ADP and thiamine pyrophosphate. There is a lag phase before O2 uptake is initiated. During this lag period, oxaloacetate is rapidly converted into α-ketoglutarate and succinate, or into malate at the expense of the NADH generated by α-ketoglutarate dehydrogenase. The ratio of the flux rates of both pathways is strongly dependent on the glutamate concentration in the medium. When all the oxaloacetate is consumed, a rapid O2 uptake is initiated. The effects of malonate on glutamate metabolism triggered by oxaloacetate and on α-ketoglutarate oxidation are reported. It is concluded that the inhibition of the succinate dehydrogenase by either malonate or oxaloacetate does not affect the rate of α-ketoglutarate dehydrogenase functioning. All the metabolites accumulated are excreted by the mitochondria in the supernatant. Some of them are then reabsorbed. These results emphasize the importance of the anion carriers in the overall process.  相似文献   

17.
K m for L-phenylalanine, L-glutamic acid, L-aspartic acid, and the corresponding keto acids were calculated, as well as V max was measured for the following pairs of substrates: L-phenylalanine-2-ketoglutarate, L-phenylalanine-oxaloacetate, L-glutamic acid-phenylpyruvate, and L-aspartic acid-phenylpyruvate for aminotransferases PAT1, PAT2, and PAT3 from Erwinia carotovora catalyzing transamination of phenylpyruvate. The ping-pong bi-bi mechanism was shown for the studied aminotransferases. The substrate inhibition (K s) of PAT3 with 2-ketoglutarate and oxaloacetate was 10.23 ± 3.20 and 3.73 ± 1.99 mM, respectively. It was shown that L-β-(N-benzylamino)alanine was a competitive inhibitor with respect to L-phenylalanine for PAT1 (K i = 0.32 ± 0.07 mM, K m = 0.45 ± 0.1 mM, V max = 11. 6 ± 0.4 U/mg) at 25 mM concentration of 2-ketoglutarate in the reaction medium. L-β-(N-methylamino)alanine is a noncompetitive inhibitor with respect to L-phenylalanine for PAT3 (K I = 138.4 ± 95.4 mM, K m = 13.7 ±3.9 mM, V max = 18.6 ± 4.1 U/mg) at 2 mM concentration of 2-ketoglutarate in the reaction medium. L-stereo isomers of nonprotein analogues of aromatic amino acids were studied as substrates for PAT1, PAT2, and PAT3. L-β-(2-Br-phenyl)alanine, L-β-(4-Br-phenyl)alanine, L-β-(2-F-phenyl)alanine, and L-(2-F)tryptophan were good substrates for all three aminotransferases; L-α-methyl-β-(2-Br-phenyl)alanine and L-O-benzyltyrosine were substrates only for PAT3; L-β-(4-F-phenyl)alanine was a substrate for PAT1 and PAT3. Thus, these analogues of aromatic amino acids can be stereoselectively synthesized using the studied aminotransferases in the presence of the corresponding keto acids.  相似文献   

18.
Optimal activity was recorded at pH 4.5–5 and pH 9.0–9.5 and specific activity was seen to be 0.013 μmoles of p-nitrophenyl phosphate/min/mg protein at 37 C at pH 4.5 and 0.00169 μmoles at pH 9.0. The ratio of acid to alkaline phosphatase was 7.7:1.0. The Km for acid phosphatase (EC 3.1.3.2) was 0.5 mM with a Vmax of 0.0128 units/mg protein and 0.2mM for alkaline phosphatase (EC 3.1.3.1) with a Vmax of 0.00175 units/mg protein. Acid phosphatase activity was optimal at 60 C and alkaline at 37 C. Linearity of enzyme activity was observed with time after the first 15 min of incubation and with homogenate concentration. KCN at 20 mM inhibited 82% of activity at pH 9.0 but also 91.5% activity at pH 4.5. NaF at 10?2M inhibited 92% of activity at pH 4.5 but had no effect at pH 9.0. The two flukicides rafoxanide and nitroxynil at 20mM had little effect on activity at pH 9.0 and pH 4.5. Enzyme activity at pH 4.5 was found to be greatest in the microsomal fraction with high activity in the lysosomal and soluble fractions. Histochemically, alkaline phosphatase was restricted to the excretory system, vitellaria, and uterus while acid phosphatase was found in the integument and gastrodermis.  相似文献   

19.
In fertilized sea urchin eggs which are kept in sea water containing α-hydrazinoornithine (αHO) at a concentration above 1 mM from the time of fertilization, cleavage is delayed markedly. The third cleavage is almost completely blocked by 3 mM αHO. Hydrazine, as well as ornithine, exerts no harmful effect on egg cleavage. αHO causes competitive inhibition of ornithine decarboxylase (ODC) in the egg homogenate. Polyamine levels decrease in fertilized eggs treated with αHO. The addition of ornithine (above 3 mM) to an egg culture containing αHO prevents the αHO-induced delay of cleavage. Putrescine (0.2–0.5 mM), which is the product of the reaction catalyzed by ODC, also relieves egg cleavage from the inhibited state. The same effect occurs in the presence of spermidine (0.2–0.5 mM) or spermine (0.1–0.8 mM). Especially, spermine (0.5 mM) completely cancels the inhibitory effect of αHO on egg cleavage. Egg cleavage is delayed only slightly in the presence of each polyamine (above 2 mM).  相似文献   

20.
γ-Aminobutyric acid transaminase (GABA-T, EC 2.6.1.19) is a pyridoxal phosphate (PLP) dependent enzyme that catalyzes the degradation of γ-aminobutyric acid. The kinetics of this reaction are studied in vitro, both in the absence, and in the presence of two inhibitors: γ-vinyl GABA (4-aminohex-5-enoic acid), and a natural product, taurine (ethylamine-2-sulfonic acid). A kinetic model that describes the transamination process is proposed. GABA-T from Pseudomonas fluorescens is inhibited by γ-vinyl GABA and taurine at concentrations of 51.0 and 78.5?mM. Both inhibitors show competitive inhibition behavior when GABA is the substrate and the inhibition constant (Ki) values for γ-vinyl GABA and taurine were found to be 26±3?mM and 68±7?mM respectively. The transamination process of α-ketoglutarate was not affected by the presence of γ-vinyl GABA, whereas, taurine was a noncompetitive inhibitor of GABA-T when α-ketoglutarate was the substrate. The inhibition dissociation constant (Kii) for this system was found to be 96±10?mM. The Michaelis-Menten constant (Km) in the absence of inhibition, was found to be 0.79±0.11?mM, and 0.47±0.10?mM for GABA and α-ketoglutarate respectively.  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