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.
Guanine deaminase in rat liver and mouse liver and brain   总被引:2,自引:2,他引:0       下载免费PDF全文
1. The guanine deaminase in rat liver supernatant preparations was resolved into two fractions, A and B, on DEAE-cellulose columns. The two differed in electrophoretic mobility and in various properties. The most noteworthy distinction between A and B components was that the enzyme A activity showed a sigmoid dependence on substrate concentration whereas the enzyme B showed classical Michaelis-Menten kinetics. The K(m) value of enzyme A for guanine was 5.3mum and that of enzyme B 20mum. 2. The entire guanine deaminase activity of mouse liver was contained in the 15000g supernatant of iso-osmotic homogenates. 3. A reinvestigation of the behaviour of rat brain 15000g supernatant guanine deaminase isoenzymes revealed that one enzyme had sigmoidal kinetics and the other enzyme showed a hyperbolic response. 4. Of the guanine deaminase in mouse brain iso-osmotic sucrose homogenate 80% was recovered in the 15000g supernatant and the rest from the particles. The supernatant guanine deaminase was resolvable into two fractions on DEAE-cellulose columns. One enzyme showed sigmoidal kinetics whereas the other showed a hyperbolic response to increasing substrate concentration; the K(m) values for the reaction with guanine were respectively 5 and 66mum. 5. The particulate fractions of mouse liver and brain were devoid of any overt inhibitory activity.  相似文献   

2.
Vick B  Beevers H 《Plant physiology》1977,59(3):459-463
Enzyme assays on organelles isolated from the endosperm of castor bean (Ricinus communis var. Hale) by sucrose density gradient centrifugation showed that palmitoyl-CoA:sn-glycerol 3-phosphate acyltransferase (EC 2.3.1.15) was localized in the membranes of the endoplasmic reticulum. Mn(2+) was required for activity, but Ca(2+) and Mg(2+) could substitute for Mn(2+) at higher concentrations. The apparent Km was 170 mum for sn-glycerol 3-phosphate and approximately 8 mum for palmitoyl-CoA. The optimum pH range was 7 to 7.5 and the principal reaction product was diacyl-sn-glycerol 3-phosphate (phosphatidic acid). Monoacyl-sn-glycerol 3-phosphate (lysophosphatidic acid) was not released as a free intermediate in the reaction. The maximum activity of the enzyme occurred immediately after imbibition, preceding the development of mitochondria and glyoxysomes.  相似文献   

3.
Palmitoyl-CoA hydrolase (EC 3.1.2.2) catalyses the irreversible hydrolysis of long-chain acyl-CoA thioesters. This enzyme is found primarily in the postmicrosomal supernatant fraction prepared from homogenates of rat brain. Either of two forms of the hydrolase, a lower-molecular-weight species of approx. 70000 or a higher-molecular-weight species of approx. 130000 can be isolated by gel filtration. The higher-molecular-weight form is obtained from columns of Sephadex G-200 eluted with buffer containing 10mum-palmitoyl-CoA or 20% (v/v) glycerol, whereas the lower-molecular-weight form is obtained when the eluting buffer does not contain palmitoyl-CoA or glycerol. The two forms of the hydrolase have the same pH optimum of 7.5, are equally sensitive to the thiol-blocking reagents p-hydroxymercuribenzoate, HgCl(2), and 5,5'-dithiobis-(2-nitrobenzoic acid), and exhibit the same K(m) (1.8mum) with palmitoyl-CoA as substrate. The two forms differ in the availability or reactivity of certain external thiol groups, as determined by covalent chromatography with activated thiol Sepharose. Dilute solutions of the lower-molecular-weight form of the hydrolase rapidly lose activity (50% in 60min at 0 degrees C), but there is no change in the K(m) with palmitoyl-CoA as substrate during this progressive inactivation. Dilutions of the hydrolase in buffer containing 10mum-palmitoyl-CoA retain full activity. However, addition of palmitoyl-CoA to solutions of the lower-molecular-weight form will not restore previously lost hydrolase activity. The evidence supports the conclusion that the substrate palmitoyl-CoA promotes the formation of a relatively stable dimer from two unstable subunits. This process may not be reversible, since the removal of palmitoyl-CoA or glycerol from solutions of the higher-molecular-weight form does not result in the appearance of the lower-molecular-weight form of the hydrolase.  相似文献   

4.
This report includes results demonstrating the existence of orotidine-5'-phosphate decarboxylase and orotidine-5'-phosphate pyrophosphorylase in plant leaves. The decarboxylase enzyme, purified 8 fold from leaves of etiolated pinto beans (Phaseolus vulgaris L.), had a pH optimum of 6.3. It was strongly inhibited by 6-azauridine-5'-phosphate; a concentration of 12 mum decreased the reaction rate 60%. The enzyme was not dependent upon magnesium ions or inhibited by p-chloromercuribenzoate. It was present in other parts of the bean plant and was found in young leaves of tomato (Lycopersicon esculentum Mill.) and Canada thistle (Cirsium arvense L.)The enzyme orotidine-5'-phosphate pyrophosphorylase, which catalyzes the formation of orotidine-5'-phosphate from orotic acid and 5-phosphoribosyl-1-pyrophosphate, was found in the etiolated bean leaves, and was also present in the leaves of tomato and Canada thistle. It was stimulated by manganous or magnesium ions and had a pH optimum of 7.2. The K(m) value obtained by varying the concentrations of 5-phosphoribosyl-1-pyrophosphate was 75 mum, and when orotic acid was varied the resulting K(m) was 3.5 mum.The presence of these 2 enzymes in higher plants, combined with previous results with inhibitors and labeled metabolites, indicates that the normal pathway of pyrimidine nucleotide synthesis in higher plants proceeds through orotic acid and OMP.  相似文献   

5.
The syntheses of 6,7-dihydrogeraniol and of its pyrophosphate are described. It is shown that this analogue of geranyl pyrophosphate is a substrate for liver prenyltransferase and that the product synthesized by this enzyme from it and isopentenyl pyrophosphate is 10,11-dihydrofarnesyl pyrophosphate. The K(m) value for 6,7-dihydrogeranyl pyrophosphate was determined to be 1.11+/-0.19mum as compared with 4.34+/-1.71mum for geranyl pyrophosphate. The maximum reaction velocity with the artifical substrate was, however, only about one-fourth of that observed with geranyl pyrophosphate. The binding of isopentenyl pyrophosphate to the enzyme was not affected by the artificial substrate.  相似文献   

6.
Lineweaver-Burk plots of reduced nicotinamide adenine dinucleotide (NADH) oxidation by membrane preparations from Bacillus subtilis are biphasic, with two K(m) values for NADH. The higher K(m) corresponds to the only K(m) observed for NADH oxidation by whole cells, whereas the lower K(m) corresponds to that observed with open cell envelopes. Membrane preparations apparently contain a small fraction of open or inverted vesicles which is responsible for the low K(m) reaction, whereas entry of NADH into the larger portion of closed, normally oriented vesicles is rate limiting and responsible for the high K(m) reaction. In contrast, the oxidation of l-alpha-glycerol-phosphate (glycerol-P) by membrane preparations shows only one K(m) that corresponds to that of glycerol-P oxidation by whole cells or lysates. Since glycerol-P dehydrogenase (NAD independent) has the same K(m), this enzyme reaction rather than entry of glycerol-P into vesicles represents the rate-limiting step for glycerol-phosphate oxidation. The K(m) for amino acid uptake by vesicles in the presence of NADH corresponds to the high K(m) for NADH oxidation, indicating that NADH energizes transport only if it enters closed, normally oriented vesicles. Studies with rotenone and proteolytic enzymes support this interpretation. The apparent efficiency of NADH in energizing uptake seems to be lower than that of glycerol-P because, under the experimental conditions usually employed, open or inverted vesicles that do not participate in amino acid uptake are responsible for the major portion of NADH oxidation. When the results are corrected for this effect, the efficiency of NADH is essentially the same as that of l-alpha-glycerol-P.  相似文献   

7.
1. The effects of ATP, inorganic phosphate and citrate on the relationship between fructose 6-phosphate concentration and initial velocity of reaction has been investigated with a partially purified preparation of rat-heart phosphofructokinase. 2. At low concentrations of ATP (<80mum) rate curves for fructose 6-phosphate approximated to Michaelis-Menten kinetics. At higher ATP concentrations rate curves were sigmoid, the K(m) for fructose 6-phosphate increased and the reaction appeared to be first-order with respect to fructose 6-phosphate at concentrations above its K(m) and of a higher order at concentrations below its K(m). Inorganic phosphate lowered the K(m) for fructose 6-phosphate and the concentration at which the apparent kinetic order decreased. 3. At 40mum-ATP, citrate was an activator at low concentration (<100mum) and an inhibitor at higher concentrations. At 0.5mm-ATP, citrate was inhibitory at all concentrations tested. 4. A new method for phosphofructokinase assay using [U-(14)C]fructose 6-phosphate is described which allows measurements to be made of the velocity of the forward reaction at known concentrations of the products of the reaction. With this method confirmatory evidence has been obtained that concentrations of ATP, AMP, phosphate and citrate may regulate phosphofructokinase in the perfused rat heart.  相似文献   

8.
With enzyme preparations from Phaseolus aureus seedlings, the initial rate of (14)C-labelled polysaccharide formation from GDP-alpha-d-[(14)C]glucose is not increased by additions of GDP-alpha-d-mannose. However, final incorporation is increased by addition of GDP-alpha-d-mannose, since the total reaction-time is extended. In contrast, the initial rate of (14)C-labelled polysaccharide formation from GDP-alpha-d-[(14)C]mannose is increased by all concentrations of GDP-alpha-d-glucose that are less than that of the GDP-alpha-d-[(14)C]mannose. Maximum stimulation of the initial rate occurs at a GDP-alpha-d-[(14)C]mannose/GDP-alpha-d-glucose concentration ratio of about 4:1. However, eventual incorporation from GDP-alpha-d-[(14)C]mannose is decreased by the addition of GDP-alpha-d-glucose, since the reaction rate falls off sharply after about 2min. Reciprocal plots of (14)C-labelled polysaccharide formation from GDP-alpha-d-[(14)C]mannose result in biphasic graphs. The two straight-line portions of the plot are joined by a curved line in the concentration range between 2-3 and 50mum. Extrapolated K(m) values for the two linear components are 0.4-1.0 and 700-1500mum. The effect of GDP-alpha-d-glucose on the kinetics of (14)C-labelled polysaccharide formation from GDP-alpha-d-[(14)C]mannose is complex, and depends on relative concentrations of the two sugar nucleotides. (14)C-labelled polysaccharide formation from GDP-alpha-d-[(14)C]glucose also results in biphasic reciprocal plots. One component appears to have K(m) about 2-3mum, the other about 200-400mum. In this reaction, GDP-alpha-d-mannose appears to be a competitive inhibitor with K(i) 20-30mum. With particulate preparations of P. aureus, GDP-alpha-d-[(14)C]glucose appears to be a precursor for the synthesis of one polysaccharide, a glucomannan, the mannose moieties of which are derived from an intermediate existing in the particulate preparation. From the rate results, GDP-alpha-d-[(14)C]mannose appears to be a precursor for at least two polysaccharides, one of which is a glucomannan.  相似文献   

9.
1. Kinetic experiments suggested the possible existence of at least two different NAD(+)-dependent aldehyde dehydrogenases in rat liver. Distribution studies showed that one enzyme, designated enzyme I, was exclusively localized in the mitochondria and that another enzyme, designated enzyme II, was localized in both the mitochondria and the microsomal fraction. 2. A NADP(+)-dependent enzyme was also found in the mitochondria and the microsomal fraction and it is suggested that this enzyme is identical with enzyme II. 3. The K(m) for acetaldehyde was apparently less than 10mum for enzyme I and 0.9-1.7mm for enzyme II. The K(m) for NAD(+) was similar for both enzymes (20-30mum). The K(m) for NADP(+) was 2-3mm and for acetaldehyde 0.5-0.7mm for the NADP(+)-dependent activity. 4. The NAD(+)-dependent enzymes show pH optima between 9 and 10. The highest activity was found in pyrophosphate buffer for both enzymes. In phosphate buffer there was a striking difference in activity between the two enzymes. Compared with the activity in pyrophosphate buffer, the activity of enzyme II was uninfluenced, whereas the activity of enzyme I was very low. 5. The results are compared with those of earlier investigations on the distribution of aldehyde dehydrogenase and with the results from purified enzymes from different sources.  相似文献   

10.
Methods are described for the assay and partial purification of mevalonate kinase from superovulated rat ovary. The total activity of mevalonate kinase in superovulated rat ovary was 1.6+/-0.14units/g wet wt.; it was unchanged by the administration of luteinizing hormone in vivo. The K(m) of a partially purified preparation of mevalonate kinase for dl-Mevalonate was 3.6+/-0.5mum; its K(m) for MgATP(2-) was 120+/-7.7mum. The enzyme was inhibited by geranyl pyrophosphate and farnesyl pyrophosphate, but not by isopentenyl pyrophosphate or 3,3'-dimethylallyl pyrophosphate. dl-mevalonate 5-phosphate inhibited at high concentrations. With both geranyl pyrophosphate and farnesyl pyrophosphate the inhibition was competitive with respect to MgATP(2-). The K(i) for inhibition by geranyl pyrophosphate was 1.3+/-0.2mum; the K(i) for inhibition by farnesyl pyrophosphate was 1.0+/-0.3mum. These findings are discussed with reference to the control by luteinizing hormone of steroidogenesis from acetate.  相似文献   

11.
The interaction of palmitoyl-CoA with porcine glucokinase was studied by the gel permeation technique. The finding that glucokinase "bound" up to 60 molecules was unexpected from the specific inhibition of rat glucokinase by long chain acyl-CoA (Tippett & Neet, J. Biol. Chem. (1982) 287, 12839-12845). Sephacryl S-200 gel filtration in the presence of palmitoyl-CoA demonstrated a protein peak without enzyme activity that was eluted earlier than the active enzyme peak, indicating a large molecular weight shift for the inactivated enzyme form and confirming a large number (greater than or equal to 30) of associated palmitoyl-CoA molecules. The binding was also verified by analyzing the absorption characteristics of the inactivated enzyme peak. In the presence of glycerol, the size of the inactivated peak greatly decreased, but the separation between the two peaks remained unchanged. Therefore, the amphiphile bound predominantly to the inactive enzyme and not to the active form, suggesting that the rapid inhibitory interactions between palmitoyl-CoA and glucokinase previously observed are specific. Parallel enzyme activity studies showed that in the time range of the column experiments (4-20 h), both the rat and pig enzyme were greatly inactivated (greater than 90%) in the presence of palmitoyl-CoA (15 microM) in the absence of glycerol. This slow inactivation is different from the immediate specific inhibition previously reported and depends on both enzyme and palmitoyl-CoA concentrations. The presence of up to 20% glycerol slowed this inactivation process. These results demonstrated that even below the critical micelle concentration, partial inactivation of glucokinase occurs in the presence of palmitoyl-CoA over a long period of time.  相似文献   

12.
Willis RM  Wahlen BD  Seefeldt LC  Barney BM 《Biochemistry》2011,50(48):10550-10558
Fatty alcohols are of interest as a renewable feedstock to replace petroleum compounds used as fuels, in cosmetics, and in pharmaceuticals. One biological approach to the production of fatty alcohols involves the sequential action of two bacterial enzymes: (i) reduction of a fatty acyl-CoA to the corresponding fatty aldehyde catalyzed by a fatty acyl-CoA reductase, followed by (ii) reduction of the fatty aldehyde to the corresponding fatty alcohol catalyzed by a fatty aldehyde reductase. Here, we identify, purify, and characterize a novel bacterial enzyme from Marinobacter aquaeolei VT8 that catalyzes the reduction of fatty acyl-CoA by four electrons to the corresponding fatty alcohol, eliminating the need for a separate fatty aldehyde reductase. The enzyme is shown to reduce fatty acyl-CoAs ranging from C8:0 to C20:4 to the corresponding fatty alcohols, with the highest rate found for palmitoyl-CoA (C16:0). The dependence of the rate of reduction of palmitoyl-CoA on substrate concentration was cooperative, with an apparent K(m) ~ 4 μM, V(max) ~ 200 nmol NADP(+) min(-1) (mg protein)(-1), and n ~ 3. The enzyme also reduced a range of fatty aldehydes with decanal having the highest activity. The substrate cis-11-hexadecenal was reduced in a cooperative manner with an apparent K(m) of ~50 μM, V(max) of ~8 μmol NADP(+) min(-1) (mg protein)(-1), and n ~ 2.  相似文献   

13.
alpha-Isopropylmalate synthase, the first specific enzyme in leucine biosynthesis, was purified approximately 100-fold from extracts of Saccharomyces sp. (strain 60615), the most effective step being specific elution with the feedback inhibitor leucine from a hydroxyapatite column. In the early steps of purification, special care was taken to protect the synthase against proteolytic activities. The apparent molecular weight of the enzyme as determined from gel filtration on a calibrated column was 137,000 in the absence and 121,000 in the presence of leucine. Inhibition by leucine was specific and strongly pH-dependent, with the leucine concentration necessary for half-maximal inhibition increasing about 10-fold as the pH increased from 7.5 to 8.5. Within this pH range, catalytic activity remained almost unchanged. The apparent K(m) values for the two substrates were found to be 16 mum for alpha-ketoisovalerate and 9 mum for acetyl-coenzyme A. K(+) was required for full activity, the apparent K(a) value being 2 mm. Leucine inhibition was of the mixed type, resulting in decreased V(max) and increased apparent K(m) values forboth substrates. Whereas no cooperative effects were observed with either substrate, positive cooperativity was seen with leucine in the presence of saturating substrate concentrations. Leucine and, to a lesser extent, alpha-ketoisovalerate stabilized the purified enzyme against heat-inactivation. The presence of acetyl-coenzyme A, on the other hand, accelerated the inactivation. In subsequent experiments, coenzyme A was recognized as the actual inactivating ligand, being effective even at lower temperatures and in concentrations which were estimated to be in the range of the enzyme concentration.  相似文献   

14.
Previous work with the bovine phenol sulfotransferase (bSULT1A1, EC ) demonstrated inhibition by CoA that was competitive with respect to the sulfuryl donor substrate, 3'-phosphoadenosine-5'-phosphosulfate (PAPS) (Leach, M., Cameron, E., Fite, N., Stassinopoulos, J., Palmreuter, N., and Beckmann, J. D. (1999) Biochem. Biophys. Res. Commun. 261, 815-819). Here we report that long chain acyl-CoAs are more potent inhibitors of bSULT1A1 and also of human dopamine sulfotransferase (SULT1A3) when compared with unesterified CoA and short chain-length acyl-CoAs. A complex pattern of inhibition was revealed by systematic variation of palmitoyl-CoA, PAPS, and 7-hydroxycoumarin, the acceptor substrate. Convex plots of apparent K(m)/V(max) versus [palmitoyl-CoA] were adequately modeled using an ordered rapid equilibrium scheme with PAPS as the leading substrate and by accounting for the possible binding of two equivalents of inhibitor to the dimeric enzyme. Interestingly, the first K(i) of 2-3 microm was followed by a second K(i) of only 0.01-0.05 microm, suggesting that positive subunit cooperativity enhances binding of long chain acyl-CoAs to this sulfotransferase. Simultaneous interaction of palmitoyl-CoA with both the nucleotide and phenol binding sites is suggested by two experiments. First, the acyl-CoA displaced 7-hydroxycoumarin from the highly fluorescent bSULT1A1.PAP.7-HC complex in a cooperative manner. Second, palmitoyl-CoA prevented the quenching of bSULT1A1 fluorescence observed with pentachlorophenol. Finally, titrations of bSULT1A1-pentachlorophenol complex with palmitoyl-CoA caused the return of protein fluorescence, and the binding of palmitoyl-CoA was highly cooperative (Hill constant of 1.9). Overall, these results suggest a model of sulfotransferase inhibition in which the 3'-phosphoadenosine-5'-diphosphate moiety of CoA docks to the PAPS domain, and the acyl-pantetheine group docks to the hydrophobic phenol binding domain.  相似文献   

15.
A soluble enzyme preparation, which catalyses the polymerization of mannose, was obtained by Triton X-100 extraction of a particulate fraction derived from Phaseolus aureus hypocotyls. The product that resulted when GDP-alpha-d-mannose was used as a substrate was a beta-(1-->4)-linked mannan, about three-quarters of which was alkali-insoluble. The mannose-polymerizing enzyme activity was at least as great in the soluble preparation as in the particulate preparation, and the specific activity of the solubilized enzyme was greater by a factor of at least 3.5. Kinetic studies of the soluble enzyme indicate that the apparent K(m) is 55-62mum, and a disproportionate increase in rate is observed at high concentrations. GDP-alpha-d-glucose is a strong competitive inhibitor of the mannose-polymerizing reaction, with an apparent K(i) of 6.2mum. The soluble enzyme is relatively unstable, losing about two-thirds of its original activity in 5h at 0 degrees C or in 24h at -20 degrees C. A solvent (acetone, butanol, diethyl ether)-extracted particulate preparation, which also exhibits the same enzyme activity, is more stable, retaining full activity for at least 5 days at -20 degrees C. There was no polymerizing-enzyme activity in the soluble enzyme preparation when UDP-d-glucose, UDP-d-galactose, UDP-d-xylose, UDP-l-arabinose or UDP-d-glucuronic acid were used as substrates. However, the soluble enzyme preparation would catalyse the polymerization of glucose, with GDP-d-glucose as substrate.  相似文献   

16.
Palmitic acid solubilized with Triton WR-1339 was converted to palmitoyl-CoA by microsomal membranes but lignoceric acid solubilized with Triton WR-1339 was not an effective substrate even though the detergent dispersed the same amount of these fatty acids and was also not inhibitory to the enzyme [I. Singh, R. P. Singh, A. Bhushan, and A. K. Singh (1985) Arch. Biochem. Biophys. 236, 418-426]. This observation suggested that palmitoyl-CoA and lignoceroyl-CoA may be synthesized by two different enzymes. We have solubilized the acyl-CoA ligase activities for palmitic and lignoceric acid of rat brain microsomal membranes with Triton X-100 and resolved them into three separate peaks (fractions) by hydroxylapatite chromatography. Fraction A (palmitoyl-CoA ligase) had high specific activity for palmitic acid and Fraction C (lignoceroyl-CoA ligase) for lignoceric acid. Specific activity of palmitoyl-CoA ligase for palmitic acid was six times higher than in Fraction C and specific activity of lignoceroyl-CoA ligase for lignoceric acid was four times higher than in Fraction A. At higher concentrations of Triton X-100 (0.5%), lignoceroyl-CoA ligase loses activity whereas palmitoyl-CoA ligase does not. Lignoceroyl-CoA ligase lost 60% of activity at 0.6% Triton X-100. Palmitoyl-CoA ligase (T1/2 of 4.5 min) is more stable at 40 degrees C than lignoceroyl-CoA ligase (T1/2 of 1.5 min). The pH optimum of palmitoyl-CoA ligase was 7.7 and that of lignoceroyl-CoA ligase was 8.4. Similar to our results with intact membranes, palmitic acid solubilized with Triton WR-1339 was converted to palmitoyl-CoA by palmitoyl-CoA ligase whereas lignoceric acid when solubilized with Triton WR-1339 was not able to act as substrate for lignoceroyl-CoA ligase. Since solubilized enzyme activities for synthesis of palmitoyl-CoA and lignoceroyl-CoA from microsomal membranes can be resolved into different fractions by column chromatography and demonstrate different properties, we suggest that in microsomal membranes palmitoyl-CoA and lignoceroyl-CoA are synthesized by two different enzymes.  相似文献   

17.
1. An enzyme (EC 2.8.2.1) that catalyses the transfer of sulphate from adenosine 3'-phosphate 5'-sulphatophosphate to phenols was purified approx. 2000-fold from male rat livers. 2. The purified preparation did not catalyse the sulphurylation of dehydroepiandrosterone, butan-1-ol, l-tyrosine methyl ester, 1-naphthylamine or serotonin. 3. At pH8.0 and 37 degrees C the K(m) values of the enzyme for p-nitrophenol and adenosine 3'-phosphate 5'-sulphatophosphate are 51 and 14mum respectively. The K(m) value for either substrate is independent of the concentration of the other. 4. The sulphurylation of phenol is inhibited by thiol compounds and glutathione at a concentration of 3mm caused an approx. 50% decrease in enzyme activity. 5. The K(m) of the enzyme for adenosine 3'-phosphate 5'-sulphatophosphate is unaffected by the presence of added glutathione but at a concentration of 5mm-glutathione the K(m) of the enzyme for its phenolic substrate is decreased.  相似文献   

18.
The mechanism of inhibition of yeast mitochondrial F(1)-ATPase by its natural regulatory peptide, IF1, was investigated by correlating the rate of inhibition by IF1 with the nucleotide occupancy of the catalytic sites. Nucleotide occupancy of the catalytic sites was probed by fluorescence quenching of a tryptophan, which was engineered in the catalytic site (beta-Y345W). Fluorescence quenching of a beta-Trp(345) indicates that the binding of MgADP to F(1) can be described as 3 binding sites with dissociation constants of K(d)(1) = 10 +/- 2 nm, K(d2) = 0.22 +/- 0.03 microm, and K(d3) = 16.3 +/- 0.2 microm. In addition, the ATPase activity of the beta-Trp(345) enzyme followed simple Michaelis-Menten kinetics with a corresponding K(m) of 55 microm. Values for the K(d) for MgATP were estimated and indicate that the K(m) (55 microm) for ATP hydrolysis corresponds to filling the third catalytic site on F(1). IF1 binds very slowly to F(1)-ATPase depleted of nucleotides and under unisite conditions. The rate of inhibition by IF1 increased with increasing concentration of MgATP to about 50 mum, but decreased thereafter. The rate of inhibition was half-maximal at 5 microm MgATP, which is 10-fold lower than the K(m) for ATPase. The variations of the rate of IF1 binding are related to changes in the conformation of the IF1 binding site during the catalytic reaction cycle of ATP hydrolysis. A model is proposed that suggests that IF1 binds rapidly, but loosely to F(1) with two or three catalytic sites filled, and is then locked in the enzyme during catalytic hydrolysis of ATP.  相似文献   

19.
Purification of the alliin lyase of garlic, Allium sativum L   总被引:3,自引:0,他引:3       下载免费PDF全文
1. Alliin lyase (EC 4.4.1.4) was purified up to sevenfold from garlic-bulb homogenates. The enzyme was unstable to storage at -10 degrees , particularly in dilute concentrations, but the addition of glycerol (final concentration 10%, v/v) stabilized the activity completely for at least 30 days. 2. The purified enzyme had an optimum pH for activity at 6.5. The addition of pyridoxal phosphate stimulated the reaction rate and the stimulation became more marked as the purification proceeded. 3. Hydroxylamine (10mum) and cysteine (0.5mm) inhibited the enzyme activity by more than 80%. Spectral studies indicated that cysteine reacted with pyridoxal phosphate bound to the protein. 4. The K(m) values for S-methyl-, S-ethyl-, S-propyl-, S-butyl- and S-allyl-l-cysteine sulphoxides were determined. With S-allyl-l-cysteine sulphoxide the K(m) was 6mm and the V(max.) was greater than those with the other substrates tested. 5. The thioether analogues of the substrates were competitive inhibitors for the lyase reaction. The K(i) decreased with increasing chain length of the alkyl substituent. With S-ethyl-l-cysteine sulphoxide as substrate the K(i) was 33, 8 and 5mm respectively for S-methyl-, S-ethyl- and S-propyl-l-cysteine. 6. The addition of EDTA or Mg(2+), Mn(2+), Co(2+) or Fe(2+) stimulated the reaction rate. Other bivalent cations either had no effect or gave a strong inhibition. In the presence of EDTA no further increase of activity was observed with added Mg(2+).  相似文献   

20.
1. The 120-fold purification of ethanolamine ammonia-lyase from Escherichia coli extracts, to apparent homogeneity, is described. Ethanolamine, dithiothreitol, glycerol and KCl protected the apoenzyme from inactivation. 2. At the optimum pH7.5, K(m) values for ethanolamine and coenzyme B(12) were 44mum and 0.42mum respectively. The K(m) for ethanolamine was markedly affected by pH, transitions occurring at pH7.0 and 8.35. 3. The enzyme was specific for ethanolamine as substrate, none of the 18 analogues tested being active. l-2-Aminopropan-l-ol (K(i) 0.86mum), dl-1-aminopropan-2-ol (K(i) 2.2mum) and dl-1,3-diaminopropan-2-ol (K(i) 88.0mum) inhibited competitively. 4. Enzyme activity was inhibited, irreversibly and non-competitively, by the coenzyme analogues methylcobalamin (K(i) 1.4nm), hydroxocobalamin (K(i) 2.1nm) and cyanocobalamin (K(i) 4.8nm). 5. Iodoacetamide inhibited in the absence of ethanolamine, but only slightly in its presence. p-Hydroxymercuribenzoate inhibited markedly even in the presence of ethanolamine. Dithiothreitol and 2-mercaptoethanol (less effectively) restored activity to the enzyme dialysed against buffer containing ethanolamine. 6. Although K(+) ions stabilized the enzyme during dialysis or storage, they were not necessary for activity. 7. Gel filtration showed the enzyme to be of high molecular weight, ultracentrifugal studies giving s(20,w) of 16.4 and an estimated mol.wt. 560400. The isoelectric point for the apoenzyme was approx. pH5.0. inhibited enzyme activity at concentrations above 1m (95% inhibition at 3m) and sodium dodecyl sulphate/polyacrylamide-gel electrophoresis indicated protein subunits of mol.wt. 61400. 8. Immunological studies showed that the E.coli enzyme was closely related to those of other enterobacteria, but only distantly to that of Clostridium sp. A double precipitin band suggested that the apoenzyme may be made up of two protein components.  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