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.
Fructans are storage carbohydrates found in many temperate grasses. The first enzyme in the biosynthetic pathway of most fructans is sucrose:sucrose fructosyl transferase (SST). In this report, we demonstrate that K+ and ionic strength noncompetitively inhibit the activity of SST from wheat (Triticum aestivum L.) stems. The Ki for this inhibition is high, 122 mM, but in the range of concentrations of K+ found in the tissue (205-314 mM). Addition of KCl to the assay system had no effect on the pH optimum (5.5) or the Km for sucrose (266 mM) but reduced the Vmax. At equivalent ionic strengths, inhibition by choline chloride was about half that of KCl, indicating that inhibition by ionic strength might be responsible for approximately 50% of the KCl inhibition. Inhibition by LiCl and (NH4)2SO4 was similar to that by choline chloride. Soluble invertase activity found in the SST preparations was less sensitive to KCl and more sensitive to choline chloride than was SST. SST from barley (Hordeum vulgare L.) stems and leaves, as well as SST from leaves of orchardgrass (Dactylis glomerata), was also inhibited by KCl. SST from onion (Allium cepa L.) bulbs and asparagus (Asparagus officinalis L.) stems was not inhibited by KCl; thus, inhibition of activity by KCl is not a universal characteristic of SST from all sources.  相似文献   

2.
1,4-benzoquinone (BQ) and 2,5-dimethyl-1,4-benzoquinone (DMBQ) were studied as inhibitors of jack bean urease in 50 mM phosphate buffer, pH 7.0. The mechanisms of inhibition were evaluated by progress curves studies and steady-state approach to data achieved by preincubation of the enzyme with the inhibitor. The obtained reaction progress curves were time-dependent and characteristic of slow-binding inhibition. The effects of different concentrations of BQ and DMBQ on the initial and steady-state velocities as well as the apparent first-order velocity constants obeyed the relationships of two-step enzyme-inhibitor interaction, qualified as mechanism B. The rapid formation of an initial BQ-urease complex with an inhibition constant of Ki = 0.031 mM was followed by a slow isomerization into the final BQ-urease complex with the overall inhibition constant of Ki* = 4.5 x 10(-5) mM. The respective inhibition constants for DMBQ were Ki = 0.42 mM, Ki* = 1.2 x 10(-3) mM. The rate constants of the inhibitor-urease isomerization indicated that forward processes were rapid in contrast to slow reverse reactions. The overall inhibition constants obtained by the steady-state analysis were found to be 5.1 x 10(-5) mM for BQ and 0.98 x 10(-3) mM for DMBQ. BQ was found to be a much stronger inhibitor of urease than DMBQ. A test, based on reaction with L-cysteine, confirmed the essential role of the sulfhydryl group in the inhibition of urease by BQ and DMBQ.  相似文献   

3.
3-Hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase activity in microsomes isolated from cultured lymphoid (IM-9) cells or freshly isolated human leukocytes was markedly decreased by either ascorbic acid or its oxidized derivative, dehydroascorbate. Inhibition of IM-9 leukocyte HMG-CoA reductase activity was log linear between 0.01 and 10 mM ascorbic acid (25 and 81% inhibition, respectively) and 0.1 and 10 mM dehydroascorbate (5 and 75% inhibition, respectively). Inhibition was noncompetitive with respect to HMG-CoA (Km = 10.2 microM (RS); ascorbic acid, Ki = 6.4 mM; dehydroascorbate, Ki = 15 mM) and competitive with respect to NADPH (Km = 16.3 microM; acetic acid, Ki = 6.3 mM; dehydroascorbate, Ki = 3.1 mM). Ascorbic acid and dehydroascorbate are interconverted through the free radical intermediate monodehydroascorbate. Reducing agents are required to convert dehydroascorbate to monodehydroascorbate, but prevent formation of the free radical from ascorbate. In microsomes from IM-9 cells, the reducing agent, dithiothreitol, abolished HMG-CoA reductase inhibition by ascorbate but enhanced inhibition by dehydroascorbate. In addition, the concentration of monodehydroascorbate present in ascorbate solutions was directly proportional to the degree of HMG-CoA reductase inhibition by 1.0 mM ascorbate. Fifty per cent inhibition of enzyme activity occurred at a monodehydroascorbate concentration of 14 microM. These data indicate that monodehydroascorbate mediates inhibition of HMG-CoA reductase by both ascorbate and dehydroascorbate. This effect does not appear to be due to free radical-induced membrane lipid modification, however, since both ascorbate and dehydroascorbate inhibited the protease-solubilized, partially purified human liver enzyme. Since inhibition of HMG-CoA reductase occurs at physiological concentrations of ascorbic acid in the human leukocyte (0.2-1.72 mM), this vitamin may be important in the regulation of endogenous cholesterol synthesis in man.  相似文献   

4.
K L Kim  D S Kang  L B Vitello  J E Erman 《Biochemistry》1990,29(39):9150-9159
The steady-state kinetics of the cytochrome c peroxidase catalyzed oxidation of horse heart ferrocytochrome c by hydrogen peroxide have been studied at both pH 7.0 and pH 7.5 as a function of ionic strength. Plots of the initial velocity versus hydrogen peroxide concentration at fixed cytochrome c are hyperbolic. The limiting slope at low hydrogen peroxide give apparent bimolecular rate constants for the cytochrome c peroxidase-hydrogen peroxide reaction identical with those determined directly by stopped-flow techniques. Plots of the initial velocity versus cytochrome c concentration at saturating hydrogen peroxide (200 microM) are nonhyperbolic. The rate expression requires squared terms in cytochrome c concentration. The maximum turnover rate of the enzyme is independent of ionic strength, with values of 470 +/- 50 s-1 and 290 +/- 30 s-1 at pH 7.0 and 7.5, respectively. The limiting slope of velocity versus cytochrome c concentration plots provides a lower limit for the association rate constant between cytochrome c and the oxidized intermediates of cytochrome c peroxidase. The limiting slope varies from 10(6) M-1 s-1 at 300 mM ionic strength to 10(8) M-1 s-1 at 20 mM ionic strength and extrapolates to 5 x 10(8) M-1 s-1 at zero ionic strength. The data are discussed in terms of both a two-binding-site mechanism and a single-binding-site, multiple-pathway mechanism.  相似文献   

5.
The force development by calcium-activated skinned frog skeletal muscle fibers and the motion on a slow time base after a quick decrease in load were studied at 0-1 degrees C as a function of the ionic strength and the degree of activation. The ionic strength was varied between 50 and 190 mM by adding appropriate concentrations of KCl to the bathing solution. Under these conditions, the fibers could be maximally activated for several cycles at low ionic strength without developing residual tension. We found that the steady isometric force in fully activated fibers linearly decreased when the KCl concentration was increased from 0 to 140 mM. The steady isotonic motion at a given relative load in fully activated fibers was almost the same at KCl concentration greater than or equal to 50 mM. In 0 and 20 mM KCl, the isotonic velocity decreased continuously for more than 300 ms. At a given relative load, the initial velocity of the motion in 0 and 20 mM KCl was about 0.6 and 0.9 times, respectively, that in 140 mM KCl. The initial velocity decreased further when residual tension developed; this observation provides additional evidence that residual tension may reflect the presence of an internal load. The effect of calcium on the motion was examined at 70 mM KCl. In this solution, the motion during the velocity transient at a given relative load appeared to be the same at different levels of activation. The speed of the subsequent motion was almost steady at high calcium levels but decreased continuously in low calcium levels. These results support the idea that at low ionic strength the response of the fiber to calcium is switch-like, but that other factors also affect the contraction mechanism under these conditions.  相似文献   

6.
1. A method is described for the purification of a form of 3-deoxy-D-arabinoheptulosonate-7-phosphate synthase (tryptophan) that probably differs from that of the native enzyme. 2. The kinetics of the reaction catalysed by 3-deoxy-D-arabinoheptulosonate-7-phosphate synthase (tryptophan) shows that the reaction proceeds via a ping-pong bi-bi mechanism, with activation by phosphoenolpyruvate (P-Prv), the first substrate, and inhibition by erythrose 4-phosphate (Ery-P) the second substrate. At low substrate concentrations, KP-Prv is 0.1 mM and KEry-P is 0.13 mM. 3. The substrates phosphoenolpyruvate and erythrose 4-phosphate and the product inorganic phosphate can protect the purified enzyme against heat denaturation, whereas the inhibitor, tryptophan, has no effect, although it binds to the enzyme in the absence of other ligands. 4. Product inhibition by inorganic phosphate is linear non-competitive with respect to phosphoenolpyruvate (Ki, slope = 22 mM and Ki, intercept = 54 mM) and substrate-linear competitive with respect to erythrose 4-phosphate (Ki, slope = 25 mM). 5. The enzyme has an activity optimum at pH 7.3 and a tryptophan inhibition optimum at pH 6.4, Trp 0.5 is 4 microM. Inhibition by tryptophan is non-competitive with respect to phosphoenolpyrovate and substrate-parabolic competitive with respect to erythrose 4-phosphate. 6. The role of the enzyme in metabolic regulation is discussed.  相似文献   

7.
Glutathione reductase (E C: 1.8.1.7; GR) was purified from rainbow trout (Oncorhynchus mykiss) liver, and some characteristics of the enzyme were investigated. The purification procedure consisted of four steps: preparation of homogenate, ammonium sulfate fractionation, affinity chromatography on 2',5'-ADP Sepharose-4B and gel filtration chromatography on Sephadex G-200. The enzyme, with a specific activity of 27.45 U/mg protein, was purified 1,654-fold with a yield of 41%. Optimal pH, stable pH, optimal temperature, optimum ionic strength, molecular mass, KM and Vmax values for GSSG and NADPH were also determined for the enzyme. In addition, Ki values and inhibition types were determined for GSH and NADP+. Additionally, inhibitory effects of metal ions (Cd+2, Cu+2, Pb+2, Hg+2, Fe+3 and Al+3) on glutathione reductase were investigated. Ki constants and IC50 values for metal ions were determined by Lineweaver-Burk graphs and plotting activity % vs. [I], respectively. IC50 values of Cd+2,Cu+2, Pb+2, Hg+2, Fe+3 and Al+3 were 0.0655, 0.082, 0.122, 0.509, 0.797 and 0.804 mM, and the Ki constants for Cd+2 and Cu+2 were 0.104+/-0.001, 0.117+/-0.001, respectively.  相似文献   

8.
1. The effect of the interaction between the charged matrix and substrate on the kinetic behaviour of bound enzymes was investigated theoretically. 2. Simple expression is derived for the apparent Km. 3. The apparent Km can only be used for the characterization of the electrostatic effect of the ionic strength does not vary with the substrate concentration. 4. The deviations from Michaelis-Menton kinetics are graphically illustrated for cases when the ionic strength varies with the substrate concentration. 5. The inhibition of the bound enzyme by a charged inhibitor at constant ionic strength is characterized by an apparent Ki. 6. When both the inhibitor concentration and the ionic strength change there is no apparent Ki, and the inhibition profile is graphically illustrated for this case. 7. Under certain conditions the electrostatic effects manifest thenselves in a sigmoidal dependence of the enzyme activity on the concentration of the substrate or inhibitor.  相似文献   

9.
The kinetic properties and inhibitor sensitivity of human sperm phospholipase A2 (PLA2; EC 3.1.1.4) were studied. Phospholipase activity was isolated from human spermatozoa by acid extraction. Hydrolysis of dipalmitoyl phosphatidylcholine was specific to the sn-2 position. Activity was sensitive to product inhibition (60% inhibition by 0.1 mM lysophosphatidylcholine). The effects of Ca2+ and sodium deoxycholate on enzyme activity were biphasic; maximal activities were observed at 0.5 mM concentration of each agent. PLA2 was stimulated (135%) by 3% dimethylsulfoxide and was inhibited by elevated ionic strength (approximately 70% inhibition with either 0.2 M NaCl or 0.2 M KCl). Two molecular forms of PLA2 were kinetically distinguishable, one with an apparent Michaelis constant and maximal reaction velocity of 3.0 microM and 0.64 mlU/mg protein and the other with respective constants of 630 microM and 32.0 mlU/mg protein. Both forms of the enzyme were Ca2+ dependent and heat stable; however, the low-Km activity was less resistant to 60 degrees C preincubation at pH 7.5 (28% inactivation of low-Km activity after 45 min, as compared to no effect on high-Km activity). Quinacrine was a noncompetitive PLA2 inhibitor with Kis for low- and high-Km activities of 0.42 mM and 0.49 mM, respectively. Trifluoperazine (calmodulin antagonist) inhibited the high-Km activity noncompetitively (Ki = 87 microM) and the low-Km activity by a mechanism consistent with the removal of a nonessential activator. Dissociation and rate constants for inactivation of low- and high-Km activities by p-bromophenacyl bromide were 0.28 mM and 0.032 min-1, and 0.73 mM and 0.066 min-1, respectively. PLA2 was inhibited by p-nitrophenyl-p'-guanidinobenzoate, at higher concentrations (10(-4)-10(-3) M) than required to inhibit trypsinlike proteinases; p-aminobenzamidine, another potent trypsin/acrosin inhibitor, stimulated (approximately 40%) PLA2 at concentrations from 2-5 mM but inhibited PLA2 (40-50%) at a concentration of 10 mM. MnCl2 (5mM) inhibited low- and high-Km PLA2 activities by 77% and 76%, respectively. Quinacrine (0.4 mM), trifluoperazine (20 microM), p-bromophenacyl bromide (20 microM), and MnCl2 (5 mM) were tested as inhibitors of the ionophore A23187-induced human acrosome reaction. Inhibition was noted only with quinacrine (32%) and MnCl2 (93%). The effect of MnCl2 was restricted to an interaction with A23187, rather than with PLA2; p-Bromophenacyl bromide inhibited (P less than 0.05) PLA2 (29%) when added to intact spermatozoa but had no effect on the acrosome reaction. PLA2 inhibition was poorly correlated with the acrosome reaction.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
The recombinant forms of the two human isozymes of glutamate decarboxylase, GAD65 and GAD67, are potently and reversibly inhibited by molecular oxygen (Ki = 0.46 and 0.29 mM, respectively). Inhibition of the vesicle-associated glutamate decarboxylase (GAD65) by molecular oxygen is likely to result in incomplete filling of synaptic vesicles with gamma-aminobutyric acid (GABA) and may be a contributing factor in the genesis of oxygen-induced seizures. Under anaerobic conditions, nitric oxide inhibits both GAD65 and GAD67 with comparable potency to molecular oxygen (Ki = 0.5 mM). Two forms of porcine cysteine sulfinic acid decarboxylase (CSADI and CSADII) are also sensitive to inhibition by molecular oxygen (Ki = 0.30 and 0.22 mM, respectively) and nitric oxide (Ki = 0.3 and 0.2 mM, respectively). Similar inhibition of glutamate decarboxylase and cysteine sulfinic acid decarboxylase by two different radical-containing compounds (O2 and NO) is consistent with the notion that these reactions proceed via radical mechanisms.  相似文献   

11.
1,2-Diacyl-sn-glycerol : CDPcholine cholinephosphotransferase (EC 2.7.8.2) and acyl-CoA : 1-acyl-sn-glycero-3-phosphocholine acyltransferase (EC 2.3.1.23) activities of rat liver microsomes can be inhibited by centrophenoxine (N,N-dimethylaminoethyl p-chlorophenoxyacetate). This inhibition is brought about by the intact centrophenoxine molecule rather than by the products of hydrolysis. A nonhydrolyzable ether analog of centrophenoxine was synthesized (neophenoxine; N,N-dimethylaminoethyl p-chlorophenoxyethyl ether) and proved most effective in inhibiting the two routes of phosphatidylcholine biosynthesis. While 50% inhibition of the cholinephosphotransferase was attained at 5 mM neophenoxine, 50% inhibition of the acyltransferase required 0.6 mM neophenoxine levels only. Inhibition of the cholinephosphotransferase (Ki approximately 1.5 mM) and the acyltransferase (Ki approximately 1 mM) by neophenoxine was shown to be noncompetitive. Other membrane-bound enzymes, such as glucose-6-phosphatase, monoacylglycerol lipase, alkaline phosphatase or phospholipase A2 were not affected by the inhibitors. Because of this specificity, and because of the high affinity of the microsomal membrane for such agents, centrophenoxine and neophenoxine should prove useful for controlling phosphatidylcholine synthesis and for modulating the phosphatidylcholine deacylation-reacylation cycle.  相似文献   

12.
The kinetic mechanism of homoserine kinase, purified to homogeneity from Escherichia coli, was examined by initial velocity techniques at pH 7.6. Whereas ATP displayed normal Michaelis-Menten saturation kinetics (Km = 0.2 mM), L-homoserine showed hyperbolic saturation kinetics only up to a concentration of 0.75 mM (Km = 0.15 mM). Above this concentration, L-homoserine caused marked but partial inhibition (Ki approximately 2 mM). The kinetic data indicated that the addition of substrates to homoserine kinase occurs by a preferred order random mechanism, with ATP preferentially binding before L-homoserine. When the ATP concentration was varied at several fixed inhibitory concentrations of L-homoserine, the resulting inhibition pattern indicated hyperbolic mixed inhibition. This suggested a second binding site for L-homoserine. L-Aspartate semialdehyde, an amino acid analog of L-homoserine, proved to be an alternative substrate of homoserine kinase (Km = 0.68 mM), and was subsequently used as a probe of its kinetic mechanism. In aqueous solution, at pH 7.5, this analog was found to exist predominantly (ca 85%) as its hydrated species. When examined as an inhibitor of the physiological reaction, L-aspartate semialdehyde showed mixed inhibition versus both L-homoserine and ATP. Although the pH profiles for the binding of L-homoserine as a substrate (Km) and as an inhibitor (Ki) were identical, the kinetic data were best fit to a two-site model, with separate catalytic and inhibitory sites for L-homoserine.  相似文献   

13.
Highly purified enzymes from Alcaligenes eutrophus H 16 were used for kinetic studies. Chorismate mutase was feedback inhibited by phenylalanine. In the absence of the inhibitor, the double-reciprocal plot was linear, yielding a Km for chorismate of 0.2 mM. When phenylalanine was present, a pronounced deviation from the Michaelis-Menten hyperbola occurred. The Hill coefficient (n) was 1.7, and Hill plots of velocity versus inhibitor concentrations resulted in a value of n' = 2.3, indicating positive cooperativity. Chorismate mutase was also inhibited by prephenate, which caused downward double-reciprocal plots and a Hill coefficient of n = 0.7, evidence for negative cooperativity. The pH optimum of chorismate mutase ranged from 7.8 to 8.2; its temperature optimum was 47 C. Prephenate dehydratase was competitively inhibited by phenylalanine and activated by tyrosine. Tyrosine stimulated its activity up to 10-fold and decreased the Km for prephenate, which was 0.67 mM without effectors. Tryptophan inhibited the enzyme competitively. Its inhibition constant (Ki = 23 muM) was almost 10-fold higher than that determined for phenylalanine (Ki = 2.6 muM). The pH optimum of prephenate dehydratase was pH 5.7; the temperature optimum was 48 C. Prephenate dehydrogenase was feedback inhibited by tyrosine. Inhibition was competitive with prephenate (Ki = 0.06 mM) and noncompetitive with nicotinamide adenine dinucleotide. The enzyme was further subject to product inhibition by p-hydroxyphenylpyruvate (Ki = 0.13 mM). Its Km for prephenate was 0.045 mM, and that for nicotinamide adenine dinucleotide was 0.14 mM. The pH optimum ranged between 7.0 and 7.6; the temperature optimum was 38 C. It is shown how the sensitive regulation of the entire enzyme system leads to a well-balanced amino acid production.  相似文献   

14.
GK (glucokinase) is an enzyme central to glucose metabolism that displays positive co-operativity to substrate glucose. Small-molecule GKAs (GK activators) modulate GK catalytic activity and glucose affinity and are currently being pursued as a treatment for Type 2 diabetes. GK progress curves monitoring product formation are linear up to 1 mM glucose, but biphasic at 5 mM, with the transition from the lower initial velocity to the higher steady-state velocity being described by the rate constant kact. In the presence of a liver-specific GKA (compound A), progress curves at 1 mM glucose are similar to those at 5 mM, reflecting activation of GK by compound A. We show that GKRP (GK regulatory protein) is a slow tight-binding inhibitor of GK. Analysis of progress curves indicate that this inhibition is time dependent, with apparent initial and final Ki values being 113 and 12.8 nM respectively. When GK is pre-incubated with glucose and compound A, the inhibition observed by GKRP is time dependent, but independent of GKRP concentration, reflecting the GKA-controlled transition between closed and open GK conformations. These data are supported by cell-based imaging data from primary rat hepatocytes. This work characterizes the modulation of GK by a novel GKA that may enable the design of new and improved GKAs.  相似文献   

15.
Effects of the non-hydrolyzable nucleotide analogue magnesium pyrophosphate (MgPPi) on cross-bridge properties were investigated in skinned smooth muscle of the guinea pig Taenia coli. A "high" rigor state was obtained by removing MgATP at the plateau of an active contraction. Rigor force decayed slowly towards an apparent plateau of approximately 25-35% of maximal active force. MgPPi markedly increased the rate of force decay. The initial rate of the force decay depended on [MgPPi] and could be described by the Michaelis-Menten equation with a dissociation constant of 1.6 mM. The decay was irreversible amounting to approximately 50% of the rigor force. Stiffness decreased by 20%, suggesting that the major part of the cross-bridges were still attached. The results can be interpreted as "slippage" of PPi-cross-bridges to positions of lower strain. The initial rate of MgPPi-induced force decay decreased with decreasing ionic strength in the range 45-150 mM and was approximately 25% lower in thiophosphorylated fibers. MgADP inhibited the MgPPi-induced force decay with an apparent Ki of 2 microM. The apparent Km of MgATP for the maximal shortening velocity in thiophosphorylated fibers was 32 microM. This low Km of MgATP suggests that steps other than MgATP-induced detachment are responsible for the low shortening velocity in smooth muscle. No effects were observed of 4 mM MgPPi on the force-velocity relation, suggesting that cross-bridges with bound MgPPi do not constitute an internal load or that binding of MgPPi is weaker in negatively strained cross-bridges during shortening.  相似文献   

16.
The present work demonstrates that the high-activity zinc metalloenzyme, carbonic anhydrase (CA II) from bovine erythrocytes is inhibited by the cyclic sulfimide, saccharin, and 2- and 4-carbobenzoxybenzene sulfonamide. A spectrophotometric method was employed to monitor the enzymatically catalyzed hydrolysis of p-nitrophenyl acetate by following the increase in absorbance at 410 nm which accompanies p-nitrophenoxide/p-nitrophenol formation. The more rapid enzymatic hydration of CO2 was monitored by using a stopped-flow spectrophotometer as well as by a modified colorimetric method of Wilbur and Anderson. The studies show that, at a given molar ratio of inhibitor to enzyme, the degree of inhibition of the enzymaic hydration of CO2 and hydrolysis of p-nitrophenyl acetate by the inhibitory compounds is essentially the same. Kinetic analyses were made at 25.0 degrees at pH 6.5 (MES buffers), pH 6.9 (HEPES buffers) and pH 7.9 (HEPES buffers) with ionic strength regulated by the addition of appropriate quantities of sodium sulfate. Lineweaver-Burk plots were used to evaluate apparent inhibition constants for each of the three inhibitors. For all the inhibitors studied, inhibition appears to be mixed (competitive/noncompetitive). For saccharin in the presence of sodium sulfate, the extent of inhibition is considerably decreased. It was found for the three inhibitors that the inhibitory potency decreases with increasing pH, and that the inhibitory potency is extremely sensitive to the shape of these rather closely related molecules. For example, apparent inhibition constants for the enzymatic hydrolysis of p-nitrophenyl acetate at pH 6.9 were Ki (saccharin) = 0.20 mM, Ki (2-carbobenzoxybenzene sulfonamide) = 0.54 mM and Ki (4-carbobenzoxybenzene sulfonamide) = 1.6 microM. For the enzymatic hydration of CO2 at pH 6.9, 0.10 mM saccharin caused 50% inhibition while 7.0 nM 4-carbobenzoxybenzene sulfonamide resulted in 50% inhibition. The results suggest that sulfonamide inhibition is caused by formation of a monodentate ligand at the zinc ion of the enzyme active site and that the more linear 4-carbobenzoxybenzene sulfonamide is better able to enter a conical enzyme active site than is 2-carbobenzoxybenzene sulfonamide or saccharin.  相似文献   

17.
The haemolymph pNP-alpha-D-glucosidase activity of emerging worker bees shows a tendency towards negative cooperativity (n = 0.92). The relation between initial velocity and enzyme concentration is not exactly linear (bilogarithmic exponent = 0.91). Between 25 degrees and 31 degrees C, the activation energy, EA = 38.2 kJ/mol. Chloramphenicol administered in vitro decreases the maximum velocity and re-establishes pure Michaelian kinetics (n congruent 1.0). The Hill coefficient for the binding of chloramphenicol to the enzyme, ni = 1.13; the values of Ki = 20.7 mM, K'i = 17.3 mM, and I50 = 17.6 mM are not significantly different from one another. These data indicate that inhibition by chloramphenicol is of the pure, non-competitive type.  相似文献   

18.
Effects of monovalent cations on the neutral rabbit liver fructose-1,6-bisphosphatase are multifunctional and dependent on their nonhydrated ionic size. (a) The maximal velocity is increased by addition of monovalent cations with the optimum stimulation occurring with a nonhydrated ionic radius of 1.2 A in the presence of a chelating agent such as EDTA. (B) Activation curves are sigmoidal with n values varying from 1.5 to 2.3 as ionic radius of monovalent cation increases. The apparent Ka values from 16.0 to 180 mM, obtained for various monovalent cations, have a linear relationship to ionic radii of cations. (c) At lower concentrations of fructose 1,6-bisphosphate monovalent cations show the inhibitory effect and the apparent Km for fructose 1,6-bisphosphate is increased as the concentration of monovalent cation is increased. A linear relationship is obtained between the slopes of increase in the Km and the reciprocals of ionic volume of monovalent cations. (d) The apparent Ka for Mg2+ is also increased as the concentration of monovalent cation is increased, and a linear relationship is obtained again between the increases in Ka and the reciprocals of ionic volume of monovalent cations. The cooperative nature for Mg2+ saturation is decreased as the Ka increases. (e) The apparent Ki for AMP is also linearly altered as the concentration of monovalent cation is varied. However, the alteration of the Ki is unusual, that is, the smaller cations than K+ increase the Ki (Li+ greater than Na+ greater than NH4+), whereas the larger cations decrease the value ((CH2CH2OH)3N+ greater than Cs+ greater than Rb+). The effect of K+ is insignificant. Alterations in the Ki are also linearly related to the reciprocals of ionic volume of monovalent cations. The cooperative nature for AMP inhibition is decreased or increased as the Ki increased or decreased. (f) In the absence of the chelating agent, the curves for Mg2+ saturation and AMP inhibition were hyperbolic without monovalent cations. By addition of monovalent cation the Ka for Mg+2+ or Ki for AMP is increased and cooperative natures for binding of both ligands are induced. For nonspherical monovalent cations, the application of "functional ionic radius" is proposed. Functional ionic radii of NH4+, (CH2OH)3CNH3+, and (CH2CH2OH)3N+ are estimated to be 1.17, 2.55, and 2.87 A, respectively. The presence of two distinct sites for the actions of monovalent cations is suggested.  相似文献   

19.
Two modes of inhibition of the Ca2+ pump in red cells by Ca2+   总被引:2,自引:0,他引:2  
Two different and independent modes of inhibition of the Ca2+ pump by Ca2+ can be detected measuring active Ca2+ extrusion from resealed ghosts of human red cells: one requires extracellular and the other requires intracellular Ca2+. Ki for inhibition by extracellular Ca2+ is about 10 mM. Extracellular Mg2+ replaces Ca2+ in inhibiting Ca2+ transport but with an apparent affinity for inhibition about 3-times less than that for Ca2+. Inhibition by external Ca2+ is not affected by Na+ or K+ at both surfaces of the cell membrane, external EGTA, internal Ca2+ or ATP. The apparent affinity for external Ca2+ progressively raises as pH increases. The effects of extracellular Ca2+ and Mg2+ are consistent with the idea that for Ca2+ pumping to proceed, external sites in the pump must be protonated and not occupied by extracellular Ca2+ or Mg2+. Inhibition by intracellular Ca2+ takes place with a Ki of about 1 mM and is independent of external Ca2+. The inhibitory effects of intracellular Ca2+ can be accounted for if Ca2+ and CaATP were competitive inhibitors of the activation of the pump by Mg2+ and MgATP, respectively.  相似文献   

20.
21S Dynein ATPase [EC 3.6.1.3] from axonemes of a Japanese sea urchin, Pseudocentrotus depressus, and its subunit fractions were studied to determine their kinetic properties in the steady state, using [gamma-32P]ATP at various concentrations, 5 mM divalent cations, and 20 mM imidazole at pH 7.0 and 0 degrees C. The following results were obtained. 1. 21S Dynein had a latent ATPase activity of about 0.63 mumol Pi/(mg . min) in 1 mM ATP, 100 mM KCl, 4 mM MgSO4, 0.5 mM EDTA, and 30 mM Tris-HCl at pH 8.0 and 25 degrees C. Its exposure to 0.1% Triton X-100 for 5 min at 25 degrees C induced an increase in the ATPase activity to about 3.75 mumol Pi/(mg . min) and treatment at 40 degrees C for 5 min also induced a similar activation. 2. The double-reciprocal plot for the ATPase activity of dynein activated by the treatment at 40 degrees C consisted of two straight lines, while that of nonactivated 21S dynein fitted a single straight line. 3. In low ionic strength solution, the Mg- and Mn-ATPase of 21S dynein showed substrate inhibition at ATP concentrations above 0.1 mM; the inhibition decreased with increasing ionic strength. Ca- and Sr-ATPase showed no substrate inhibition. 4. Both the Vmax and Km values of dynein ATPase decreased reversibly upon addition of about 40% (v/v) glycerol. In the presence of glycerol, the dynein ATPase showed an initial burst of Pi liberation. The apparent Pi-burst size was 1.0 mol/(10(6) g protein) and the true size was calculated to be 1.6 mol/1,250 K after correcting for the effect of Pi liberation in the steady state and the purity of our preparation. 5. One of the subunit fractions of 21S dynein which was obtained by the method of Tang et al. showed substrate inhibition and an initial burst of Pi liberation of 1.4 mol/(10(6) g protein) in the presence of 54% (v/v) glycerol.  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