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.
Thermal stability of glucoamylase [α-1,4: 1,6-glucan-4: 6-glucohydrolase, EC 3.2.1.3] from Rhizopus niveus was investigated in detail mainly at 60.0±0.1°C (pH 4.5) in relation to kinetics both in the presence and in the absence of various ligands. With substrate analogues, such as glucose, lactose and gluconolactone, and with glycerol, thermostability of the enzyme greatly increased, whereas decreased with dioxane, ethanol and glycol. To elucidate such phenomena clearly, several model equations were propounded on the assumption that there were two main factors which would play an essential role in heat stability of the enzyme: binding of ligands and the dielectric constants of solvents. From the model equations, binding constants of the ligands were estimated in order to confirm the validity of the assumption; e.g., the inhibitor constant for glucose (at 59.5°C) corresponding to the reciprocal of binding constant was evaluated to be 0.25 m as was expected.  相似文献   

2.
1. Diverse classes of compounds such as dicarboxylates, pyrophosphates, quinols and nitrophenols are known to activate mitochondrial succinate dehydrogenase (EC 1.3.99.1). Examples in each class -- malonate, pyrophosphate, ubiquinol and 2,4-dinitrophenol -- are selected for comparative studies on the kinetic constants and structural relationship. 2. The activated forms of the enzyme obtained on preincubating mitochondria with the effectors exhibited Michaelian kinetics and gave double-reciprocal plots which are nearly parallel to that of the basal form. On activation, Km for the substrate also increased along with V. The effectors activated the enzyme at low concentrations and inhibited, in a competitive fashion, at high concentrations. The binding constant for activation was lower than that for inhibition for each effector. 3. These compounds possess ionizable twin oxygens separated by a distance of 5.5 +/- 0.8 A and having fractional charges in the range of -0.26 to -0.74 e. The common twin-oxygen feature of the substrate and the effectors suggested the presence of corresponding counter charges in the binding domain. The competitive nature of effectors with the substrate for inhibition further indicated the close structural resemblance of the activation and catalytic sites.  相似文献   

3.
  • 1.1. Aldehyde dehydrogenase from rat testis cytosol has been purified to electrophoretic homogeneity. With an isoelectric point of 9.5, the enzyme appears a dimer with a subunit molecular weight of 52,500.
  • 2.2. The influence of pregnenolone and progesterone on the kinetic behaviour has been investigated using valeraldehyde as substrate.
  • 3.3. The kinetic data were fitted to a modified version of the Monod-Wyman-Changeux model and the fitting procedure resulted in a good correspondence between theoretical and experimental reaction rates over a wide range of valeraldehyde concentrations.
  • 4.4. According to the model, the dimeric enzyme is in equilibrium between two confonnational states R and T. The R state displays higher affinity for valeraldehyde, but lower catalytic power. In the absence of substrates and effectors the [T]/[R] ratio is near to 1.
  • 5.5. Pregnenolone and progesterone activate the enzyme by stabilizing the more active state T and by increasing the catalytic power of the R state. The increase of activity is counteracted by the inhibition exerted by both steroids on the T state.
  相似文献   

4.
The kinetic behavior of -chymotrypsin was studied in water–DMSO mixtures at concentrations of the organic solvent that do not cause irreversible denaturation of the enzyme. Various substrates (N-substituted derivatives of L-tyrosine) were found to display substantially different kinetic patterns of interaction with -chymotrypsin, which can be described by totally different kinetic schemes. The differences were ascribed to competition between the N-acyl group of the substrate and the DMSO molecule at the S 2 site of substrate binding to the active site of the enzyme.  相似文献   

5.
ADP-glucose pyrophosphorylase catalyzes the synthesis of ADP-glucose (ADP-Glc) from Glc-1-phosphate (G-1-P) and ATP. Kinetic studies were performed to define the nature of the reaction, both in the presence and absence of allosteric effector molecules. When 3-phosphoglycerate (3-PGA), the putative physiological activator, was present at a saturating level, initial velocity studies were consistent with a Theorell-Chance BiBi mechanism and product inhibition data supported sequential binding of ATP and G-1-P, followed by ordered release of pyrophosphate and ADP-Glc. A sequential mechanism was also followed when 3-PGA was absent, but product inhibition patterns changed dramatically. In the presence of 3-PGA, ADP-Glc is a competitive inhibitor with respect to ATP. In the absence of 3-PGA—with or without 5.0 mm inorganic phosphate—ADP-Glc actually stimulated catalytic activity, acting as a feedback product activator. By contrast, the other product, pyrophosphate, is a potent inhibitor in the absence of 3-PGA. In the presence of subsaturating levels of allosteric effectors, G-1-P serves not only as a substrate but also as an activator. Finally, in the absence of 3-PGA, inorganic phosphate, a classic inhibitor or antiactivator of the enzyme, stimulates enzyme activity at low substrate by lowering the KM values for both substrates.Plant ADP-Glc pyrophosphorylase (AGPase) catalyzes an important, rate-limiting step in starch biosynthesis: the reversible formation of ADP-Glc from ATP and Glc-1-P (G-1-P). Most AGPases are regulated by effector molecules derived from the prevalent carbon metabolism pathway, with inorganic phosphate (Pi) and 3-phosphoglycerate (3-PGA) being the most studied effectors of higher plants. Interestingly, the barley (Hordeum vulgare) endosperm form of AGPase is unique among higher plant homologs in its insensitivity to both 3-PGA and Pi (Kleczkowski et al., 1993a). Heat lability (as often found for endosperm AGPases) and reductive activation (for those AGPases harboring an N-terminal Cys residue in the small subunit) are also important mechanisms by which AGPases are regulated (Fu et al., 1998; Tiessen et al., 2002).Transgenic plant studies emphasize the importance of allosteric effectors in controlling enzyme activity and, in turn, starch yield. For example, expressing an allosterically enhanced Escherichia coli AGPase resulted in a 35% increase in potato (Solanum tuberosum) tuber starch (Stark et al., 1992) and a 22% to 25% increase in maize (Zea mays) seed starch (Wang et al., 2007). Rice (Oryza sativa) seed weight was increased up to 11% by expression of a second E. coli-derived AGPase mutant with altered allosteric properties (Sakulsingharoj et al., 2004). In another example, expressing a maize AGPase variant with less sensitivity to Pi and enhanced heat stability led to a 38% increase in wheat (Triticum aestivum) yield (Smidansky et al., 2002), a 23% increase in rice yield (Smidansky et al., 2003), and up to a 68% increase in maize yield (L.C. Hannah, unpublished data). Increases in these cases were due to enhanced seed number. Finally, transgenic expression of an allosterically altered potato tuber AGPase enhanced Arabidopsis (Arabidopsis thaliana) leaf transitory starch turnover and improved growth characteristics (Obana et al., 2006) and enhanced the fresh weight of aerial parts of lettuce (Lactuca sativa) plants (Lee et al., 2009).In higher plants, AGPase is a heterotetramer, consisting of two large and two small subunits; by contrast, most bacterial AGPases are homotetramers. Crystal structures of a bacterial AGPase and a nonnative, small subunit homotetramer derived from the potato tuber enzyme have been described recently (Jin et al., 2005; Cupp-Vickery et al., 2008). Unfortunately, since both structures were determined in the presence of high sulfate concentrations, both enzymes are in inactive forms.While AGPase allosteric regulation has received a great deal of attention, the kinetic mechanism has been defined completely only for two cases: the homotetrameric form from the bacterium Rhodospirillum rubrum and the plant heterotetrameric enzyme from barley leaf (Paule and Preiss, 1971; Kleczkowski et al., 1993b). The kinetic mechanism is sequential in both cases, with ATP the first substrate bound and ADP-Glc the final product released. Despite this similarity, there are important differences, most notably the existence of an isomerization step following ADP-Glc release, so that this product and ATP bind to different forms of the barley enzyme. This isomerization step is absent from the bacterial enzyme. Interestingly, isoforms of the closely related nucleoside diphospho-Glc family exhibit fundamentally different kinetic mechanisms. Some UDP-Glc pyrophosphorylases catalyze a sequential BiBi mechanism (Elling, 1996), while others, such as dTDP-Glc and CDP-Glc pyrophosphorylases from Salmonella, employ a ping-pong mechanism (Lindqvist et al., 1993, 1994).Because of the mechanistic diversity exhibited by pyrophosphorylases in general and by the two well-characterized AGPases in particular, we investigated the kinetic mechanism of the recombinant maize endosperm AGPase. We were particularly interested in the roles played by allosteric effectors that appear to be critically important in catalytic efficiency and, thus, starch content. Surprisingly, patterns of initial velocity at varying substrate concentrations as well as product inhibition behavior were identical to those observed for the homotetrameric bacterial enzyme (Paule and Preiss, 1971) and differed significantly from the heterotetrameric barley leaf enzyme (Kleczkowski et al., 1993b). Moreover, we found that both G-1-P and ADP-Glc could stimulate AGPase catalytic activity beyond that expected for simple substrate effects. We also found that the classic inhibitor, Pi, actually enhanced AGPase activity at low substrate concentrations but inhibited activity at high substrate levels. A model is presented to account for this observation. Finally, we determined that 3-PGA only stimulates AGPase activity by 2.5-fold if care is taken to saturate with substrates during assays.  相似文献   

6.
The kinetic mechanism of NADP-dependent malic enzyme purified from maize leaves was studied in the physiological direction. Product inhibition and substrate analogues studies with 3 aminopyridine dinucleotide phosphate and tartrate indicate that the enzyme reaction follows a sequential ordered Bi-Ter kinetic mechanism. NADP is the leading substrate followed by l-malate and the products are released in the order of CO2, pyruvate and NADPH. The enzyme also catalyzes a slow, magnesium-dependent decarboxylation of oxaloacetate and reduction of pyruvate and oxaloacetate in the presence of NADPH to produce l-lactate and l-malate, respectively.  相似文献   

7.
Most investigations of the allosteric properties of the regulatory enzyme aspartate transcarbamoylase (ATCase) from Escherichia coli are based on the sigmoidal dependence of enzyme activity on substrate concentration and the effects of the inhibitor, CTP, and the activator, ATP, on the saturation curves. Interpretations of these effects in terms of molecular models are complicated by the inability to distinguish between changes in substrate binding and catalytic turnover accompanying the allosteric transition. In an effort to eliminate this ambiguity, the binding of the 3H-labeled bisubstrate analog N-(phosphonacetyl)-L-aspartate (PALA) to aspartate transcarbamoylase in the absence and presence of the allosteric effectors ATP and CTP has been measured directly by equilibrium dialysis at pH 7 in phosphate buffer. PALA binds with marked cooperativity to the holoenzyme with an average dissociation constant of 110 nM. ATP and CTP alter both the average affinity of ATCase for PALA and the degree of cooperativity in the binding process in a manner analogous to their effects on the kinetic properties of the enzyme; the average dissociation constant of PALA decreases to 65 nM in the presence of ATP and increases to 266 nM in the presence of CTP while the Hill coefficient, which is 1.95 in the absence of effectors, becomes 1.35 and 2.27 in the presence of ATP and CTP, respectively. The isolated catalytic subunit of ATCase, which lacks the cooperative kinetic properties of the holoenzyme, exhibits only a very slight degree of cooperativity in binding PALA. The dissociation constant of PALA from the catalytic subunit is 95 nM. Interpretation of these results in terms of a thermodynamic scheme linking PALA binding to the assembly of ATCase from catalytic and regulatory subunits demonstrates that saturation of the enzyme with PALA shifts the equilibrium between holoenzyme and subunits slightly toward dissociation. Ligation of the regulatory subunits by either of the allosteric effectors leads to a change in the effect of PALA on the association-dissociation equilibrium.  相似文献   

8.
In pathogenic trypanosomes, trypanothione synthetase (TryS) catalyzes the synthesis of both glutathionylspermidine (Gsp) and trypanothione (bis(glutathionyl)spermidine (T(SH)2)). Here we present a thorough kinetic analysis of Trypanosoma brucei TryS in a newly developed phosphate buffer system at pH 7.0 and 37 °C, mimicking the physiological environment of the enzyme in the cytosol of bloodstream parasites. Under these conditions, TryS displays Km values for GSH, ATP, spermidine, and Gsp of 34, 18, 687, and 32 μm, respectively, as well as Ki values for GSH and T(SH)2 of 1 mm and 360 μm, respectively. As Gsp hydrolysis has a Km value of 5.6 mm, the in vivo amidase activity is probably negligible. To obtain deeper insight in the molecular mechanism of TryS, we have formulated alternative kinetic models, with elementary reaction steps represented by linear kinetic equations. The model parameters were fitted to the extensive matrix of steady-state data obtained for different substrate/product combinations under the in vivo-like conditions. The best model describes the full kinetic profile and is able to predict time course data that were not used for fitting. This system''s biology approach to enzyme kinetics led us to conclude that (i) TryS follows a ter-reactant mechanism, (ii) the intermediate Gsp dissociates from the enzyme between the two catalytic steps, and (iii) T(SH)2 inhibits the enzyme by remaining bound at its product site and, as does the inhibitory GSH, by binding to the activated enzyme complex. The newly detected concerted substrate and product inhibition suggests that TryS activity is tightly regulated.  相似文献   

9.
In this study, rate equations that predict the regulatory kinetic behavior of homocitrate synthase were derived, and simulation of the predicted behavior was carried out over a range of values for the kinetic parameters. The data obtained allow application of the resulting expressions to enzyme systems that exhibit activation and inhibition as a result of the interaction of effectors at multiple sites in the free enzyme. Homocitrate synthase was used as an example in terms of its activation by Na+ binding to the active enzyme conformer at an allosteric site, inhibition by binding to the active site, and inhibition by lysine binding to the less active enzyme conformer.  相似文献   

10.
11.
We have previously described simple models for active transport and have derived steady state equations for the unidirectional flux of substrate in terms of a minimal set of kinetic parameters. Here we consider how to maximize the pumping rate of a carrier-enzyme through the optimal utilization of the ATP hydrolysis reaction. The equations for net flux contain rate constants and dissociation constants and these determine the maximum velocities and affinities measured in transport kinetic analysis. It is assumed that the rate constants can evolve to the diffusion limited rate of substrate binding as has apparently occurred in the enzyme triosephosphate isomerase (Knowles & Albery, 1977). The dissociation constants of the rate limiting intermediates fit the affinities for substrates on different sides of the membrane and are dependent on the basic free energy levels (Hill, 1976) of the carrier substrate system. From our analysis it is clear that there are three ways to design a system with optimal affinities and that the choice is linked to the sequence of substrate binding. It is possible to use free energy differences of isomerization (Boyer, 1975) or ligand-ligand interactions (Weber, 1975) both of which have been described previously, but which are incorporated here into a unified treatment. A third possibility is to couple the binding step of a transported ligand to the progress of a chemical reaction as might occur, for example, if Na+ must be bound before the carrier can be phosphorylated. In this way the free energy of hydrolysis can be used not only to drive the overall pumping reaction, but also to fix differentially the affinity for substrate on either side of the membrane, as required for rapid pumping.  相似文献   

12.

Background

Fructose-1,6-bisphosphatase, a major enzyme of gluconeogenesis, is inhibited by AMP, Fru-2,6-P2 and by high concentrations of its substrate Fru-1,6-P2. The mechanism that produces substrate inhibition continues to be obscure.

Methods

Four types of experiments were used to shed light on this: (1) kinetic measurements over a very wide range of substrate concentrations, subjected to detailed statistical analysis; (2) fluorescence studies of mutants in which phenylalanine residues were replaced by tryptophan; (3) effect of Fru-2,6-P2 and Fru-1,6-P2 on the exchange of subunits between wild-type and Glu-tagged oligomers; and (4) kinetic studies of hybrid forms of the enzyme containing subunits mutated at the active site residue tyrosine-244.

Results

The kinetic experiments with the wild-type enzyme indicate that the binding of Fru-1,6-P2 induces the appearance of catalytic sites with lower affinity for substrate and lower catalytic activity. Binding of substrate to the high-affinity sites, but not to the low-affinity sites, enhances the fluorescence emission of the Phe219Trp mutant; the inhibitor, Fru-2,6-P2, competes with the substrate for the high-affinity sites. Binding of substrate to the low-affinity sites acts as a “stapler” that prevents dissociation of the tetramer and hence exchange of subunits, and results in substrate inhibition.

Conclusions

Binding of the first substrate molecule, in one dimer of the enzyme, produces a conformational change at the other dimer, reducing the substrate affinity and catalytic activity of its subunits.

General significance

Mimics of the substrate inhibition of fructose-1,6-bisphosphatase may provide a future option for combatting both postprandial and fasting hyperglycemia.  相似文献   

13.
The tight binding of Meisenheimer intermediate with octopus digestive gland glutathione transferase was analyzed with 1,3,5-trinitrobenzene, which forms a trapped Meisenheimer complex with glutathione because there is no leaving group at the ipso carbon. By steady-state enzyme kinetic analysis, an inhibition constant of 1.89 ± 0.17 M was found for the transient formed, S-(2,4,6-trinitrophenyl) glutathione. The above inhibition constant is 407-fold smaller than the K m value for the substrate (2,4-dinitrochlorobenzene). Thus, S-(2,4,6-trinitrophenyl) glutathione is considered to be a transition-state analog. The tight binding of this inhibitor to the enzyme provides an explanation for the involvement of the biological binding effect on the rate enhancement in the glutathione transferase-catalyzed SNAr mechanism.  相似文献   

14.
Summary The mitochondrial succinate dehydrogenase (E.C. 1.3.3.99) is subjected to apparently complicated regulatory mechanism. Yet, systematic analysis of the mechanism reveals the simplicity of the control. There are two stable forms of the enzyme; the non-active form stabilized as 1:1 complex with oxaloacetate and the active form stabilized by binding of activating ligands. This model quantitatively describes either the equilibrium level of active enzyme or the kinetics of activation-deactivation, in the presence of various concentrations of opposing effectors. The site where the regulatory ligands interact with the enzyme is not the substrate bonding site. The marked differences of dissociation constants of the same ligand from the two sites clearly distinguish between them.This model is fully developed for simple cases where the activating ligands are dicarboxylic acids or monovalent anions. On the other hand with activators such as ATP or CoQH2, quantitation is still not at hand. This stems from the difficulties in maintaining determined, measurable, concentrations of the ligand in equilibrium with the membranal enzyme.While in active form the histidyl flavin moity of the enzyme is reduced by physiological substrate (succinate; CoQH2). The non-active form is not reduced by these compounds, only strong reductants with low redox potential reduce the non-active enzyme. It is suggested that deactivation is a simple modulation of the redox potential of the flavin form E 0 mV in the active enzyme to E < –190 mV. The switch from one state to another might be achieved by distortion of the planar form of oxidized flavin to the bend configuration of the reduced flavin. Thus, in the active enzyme such distortion will destabilize the oxidized state of the flavin, shifting the redox potential to the higher value. The binding of oxaloacetate to the regulatory sites releases the distorting forces by relaxing the conformation of the enzyme. Consequently, the flavin assumes its planar form with the low redox potential. This assumption is supported by the spectral shifts of the flavin associated with the activation deactivation transition.The suicidal oxidation of malate to oxaloacetate, carried by the succinate dehydrogenase, plays an important role in modulating the enzyme activity in the mitochondria. This mechanism might supply oxaloacetate for deactivation in spite of the negligible concentration of free oxaloacetate in the matrix. The oxidation of malate by the enzyme is controlled by the redox potential at the immediate vicinity of the enzyme, and is imposed by the redox level of the membranal quinone.Finally, the modulation of succinate dehydrogenase activity is closely associated with regulation of NADH oxidation through the mutual inhibition between oxidases (Gutman, M. in Bioenergetics of Membranes, L. Packer et al., ed. Elsevier 1977, p. 165). The consequence of these interactions is the selection for the main electron donnor for the respiratory chain, during mixed substrate respiration, according to the metabolic demands from the mitochondria.Abbreviations SDH succinate dehydrogenase (succinate: acceptor oxidoreductase (E.C. 1.3.99.1)); - OAA oxaloacetate - Act activator - EA, EA active and non active forms of the enzyme, respectively - K'eq apparent equilibrium constant - K'd apparent dissociation constant - KAct, KOAA dissociation constant of the respective ligand from the enzyme - K'a, k'd the apparent rate constants of activation and deactivation, respectively - ka, kd the true rate constant of activation and deactivation respectively - ETP, ETPII non phosphorylating and phosphorylating submitochondrial particles - PMS phenazine methosulfate - DCIP dichlorophenol indophenol - CoQ ubiquinone - TIFA Thenotriflouvoacetone - NEM N methyl Maleimide  相似文献   

15.
C F Hawkins  A S Bagnara 《Biochemistry》1987,26(7):1982-1987
The reaction catalyzed by adenosine kinase purified from human erythrocytes proceeds via a classical ordered sequential mechanism in which adenosine is the first substrate to bind to and AMP is the last product to dissociate from the enzyme. However, the interpretation of the steady-state kinetic data is complicated by the finding that while AMP acts as a classical product inhibitor at concentrations greater than 5 mM, at lower concentrations AMP can act as an apparent activator of the enzyme under certain conditions. This apparent activation by AMP is proposed to be due to AMP allowing the enzyme mechanism to proceed via an alternative reaction pathway that avoids substrate inhibition by adenosine. Quantitative studies of the protection of the enzyme afforded by adenosine against both spontaneous and 5,5'-dithiobis(2-nitrobenzoic acid)-mediated oxidation of thiol groups yielded "protection" constants (equivalent to enzyme-adenosine dissociation constant) of 12.8 microM and 12.6 microM, respectively, values that are more than an order of magnitude greater than the dissociation constant (Kia = 0.53 microM) for the "catalytic" enzyme-adenosine complex. These results suggest that adenosine kinase has at least two adenosine binding sites, one at the catalytic center and another quite distinct site at which binding of adenosine protects the reactive thiol group(s). This "protection" site appears to be separate from the nucleoside triphosphate binding site, and it also appears to be the site that is responsible for the substrate inhibition caused by adenosine.  相似文献   

16.
  • 1.1. The kinetic and regulatory properties of phenylalanine ammonia-lyase from Citrus sinensis fruit tissue were investigated. The substrate specificity of the enzyme was determined as well as the effects of pH and temperature on the catalytic activity.
  • 2.2. The enzyme exhibits negative homotropic effects between the substrate binding centra.
  • 3.3. Binding of l-phenylalanine to the enzyme is characterized by two Km-values; KmL = 13 μM and KmH = 52 μM; with a Hill-interaction coefficient of 0.75.
  • 4.4. The enzyme is subject to product inhibition by trans-cinnamate, but the effects of allosteric effectors and inhibitors seem to be of much greater importance in the short-term regulation of phenylpropanoid metabolism in Citrus sinensis.
  • 5.5. The enzyme activity was found to be modulated by end-products of diverging metabolic pathways, viz. umbelliferone, scopoletin, naringenin, quercetin, kaempferol, benzoic acid and gallic acid.
  相似文献   

17.
Biological membranes contain many domains enriched in phospholipid lipids and there is not yet clear explanation about how these domains can control the activity of phospholipid metabolizing enzymes. Here we used the surface dilution kinetic theory to derive general equations describing how complex substrate distributions affect the activity of enzymes following either the phospholipid binding kinetic model (which assumes that the enzyme molecules directly bind the phospholipid substrate molecules), or the surface-binding kinetic model (which assumes that the enzyme molecules bind to the membrane before binding the phospholipid substrate). Our results strongly suggest that, if the enzyme follows the phospholipid binding kinetic model, any substrate redistribution would increase the enzyme activity over than observed for a homogeneous distribution of substrate. Besides, enzymes following the surface-binding model would be independent of the substrate distribution. Given that the distribution of substrate in a population of micelles (each of them a lipid domain) should follow a Poisson law, we demonstrate that the general equations give an excellent fit to experimental data of lipases acting on micelles, providing reasonable values for kinetic parameters—without invoking special effects such as cooperative phenomena. Our theory will allow a better understanding of the cellular-metabolism control in membranes, as well as a more simple analysis of the mechanisms of membrane acting enzymes.  相似文献   

18.
AMP-deaminase from human term placenta was chromatographed on a phosphocellulose column and physico-chemical and immunological properties of the purified enzyme were investigated. At physiological pH7.0, in the absence of regulatory ligands (control conditions) studied AMP-deaminase manifested sigmoid-shaped substrate saturation kinetics, with half-saturation parameter (S0.5) value of about 7 mM. Addition of important allosteric effectors (ATP, ADP or orthophosphate) modified kinetic properties of studied AMP-deaminase, influencing mainly the value of S0.5 parameter. Micromolar concentrations of stearylo-CoA inhibited potently the enzyme making it no longer sensitive towards 1 mM ATP-induced activation. SDS-PAGE electrophoresis of the purified enzyme revealed presence of 68 kDa protein fragment, reacting with anti-(human) liver AMP-deaminase antibodies. Experimental results presented indicate that liver type of AMP-deaminase is an enzyme form present in human term placenta.  相似文献   

19.
A study of the effect of monovalent cations on the steady-state kinetic parameters for the hydrolysis of the synthetic substrate N alpha-benzoyl-L-arginine-p-nitroanilide by activated bovine plasma protein C (APC) has been undertaken. The enzyme displayed a strict requirement for monovalent cations in its expression of amidolytic activity toward this substrate. Analysis of the variation in initial hydrolytic reaction rates, as a function of metal ion concentrations, suggested that at least two cation sites, or classes of sites, were necessary for catalysis to occur. After examination of the rate equations consequential to many different enzymic mechanisms that could account for these kinetic data, a mechanism was developed that fit the great majority of the experimental observations. In this mechanism it is postulated that cations bind to the enzyme in pairs, with a kinetically observable single binding constant, either preceded by or followed by binding of substrate. Catalysis occurs only after the enzyme-(metal cation)2-substrate complex is assembled. Some physical support for this mechanism was obtained upon the discovery that the binding (dissociation) constant for a competitive inhibitor of APC, p-aminobenzamidine, as determined by kinetic methodology, was independent of the concentration of Na+ and Cs+.  相似文献   

20.
The reduction of purine nucleoside diphosphates by murine ribonucleotide reductase requires catalytic (R1) and free radical-containing (R2) enzyme subunits and deoxynucleoside triphosphate allosteric effectors. A quantitative 16 species model is presented, in which all pertinent equilibrium constants are evaluated, that accounts for the effects of the purine substrates ADP and GDP, the deoxynucleoside triphosphate allosteric effectors dGTP and dTTP, and the dimeric murine R2 subunit on both the quaternary structure of murine R1 subunit and the dependence of holoenzyme (R1(2)R2(2)) activity on substrate and effector concentrations. R1, monomeric in the absence of ligands, dimerizes in the presence of substrate, effectors, or R2(2) because each of these ligands binds R1(2) with higher affinity than R1 monomer. This leads to apparent positive heterotropic cooperativity between substrate and allosteric effector binding that is not observed when binding to the dimeric protein itself is evaluated. Allosteric activation results from an increase in k(cat) for substrate reduction upon binding of the correct effector, rather than from heterotropic cooperativity between effector and substrate. Neither the allosteric site nor the active site displays nucleotide base specificity: dissociation constants for dGTP and dTTP are nearly equivalent and K(m) and k(cat) values for both ADP and GDP are similar. R2(2) binding to R1(2) shows negative heterotropic cooperativity vis-à-vis effectors but positive heterotropic cooperativity vis-à-vis substrates. Binding of allosteric effectors to the holoenzyme shows homotropic cooperativity, suggestive of a conformational change induced by activator binding. This is consistent with kinetic results indicating full dimer activation upon binding a single equivalent of effector per R1(2)R2(2).  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