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.
Saturation curves of activity versus concentration were determined for aspartate transcarbamylase from Escherichia coli (EC 2.1.3.2) for the substrate L-aspartate at saturating carbamyl phosphate (4.8 mM) in buffered solution at pH values from 6.0 to 12.0. Hill coefficients were obtained from the sigmoidal curves. At pH values from 7.8 to 9.1, where substrate inhibition causes difficulties in the Hill approximation, our kinetic scheme includes substrate inhibition and residual activity in the abortive enzyme-substrate complex. The plot of Hill coefficient versus pH has pKalpha values of 7.4 and 9.8 at the half-maximum positions of the curve which has a plateau from pH 8.1 to 9.1. These pKalpha values may be associated with functional groups involved in the allosteric transition which activates the enzyme. A plot of [S]0.5 versus pH shows a pKalpha of 8.5, which may belong to a residue either at or near the aspartate binding site. At 50 mM aspartate concentration the pH-rate profile shows maxima at pH values of 8.8 and 10.0 (cf. Weitzman, P.D.J., and Wilson, I.B.(1966)J. Biol. Chem. 2418 5481-5488, who used 100 mM aspartate). However, when the pH-dependent substrate inhibition is included, the calculated Vmax--H curve is bell-shaped like that of the isolated catalytic subunit.  相似文献   

2.
The amylolytic system of Streptomyces sp IMD 2679 is composed of three α-amylases, amylase I, II and III, with temperature maxima of 60, 60–65 and 65°C, respectively. Although each α-amylase displayed higher stability in the pH range 6.0–8.5 than at pH 5.0–5.5, differences in their thermostabilities were more evident as the pH increased from pH 6.0 to 8.5. There was a 14-min difference in half-lives between amylase III, the most thermostable enzyme and amylase II at pH 6.0, and a 46-min difference in the half-lives of amylase III and the least thermostable enzyme, amylase I at pH 6.5. In addition, the α-amylases underwent a pH-dependent monomer-dimer transformation. Increased thermostability of the α-amylases was reflected in the variable contents of amino acids (Arg, His, Ser) responsible for electrostatic interactions, and in the levels of aliphatic and bulky hydrophobic amino acids. There was a two-fold reduction in Cys levels in amylase III relative to amylase I and II. Received 31 August 1998/ Accepted in revised form 11 January 1999  相似文献   

3.
1. The ATP sites. Homotropic interactions between ATP sites have been studied in a very large range of Na+ and K+ concentrations. The ( Na+, K+)-activated ATPase displays Michaelis-Menten kinetics for ATP under standard concentration conditions of Na+ (100 mM) and K+ (10 mM). The steady-state kinetics behavior changes at very low concentrations of K+ where negative cooperativity is observed. The existence of a high affinity and a low affinity site for ATP was clearly demonstrated from the study of the ATP stimulated hydrolysis of p-nitrophenylphosphate in the presence of Na+ and K+. The ratio of apparent affinities of high and low affinity sites for ATP is 86 at pH 7.5. 2. The Na+ sites. The binding of Na+ to its specific stimulatory sites (internal sites) is characterized by positive cooperativity with a Hill coefficient n(H(Na+))=2.0. Homotropic interactions between Na+ sites are unaffected by variations of the K+ concentration. 3. The K+ sites. (a) Binding of K+ to the (external) stimulatory site of the ATPase has been analyzed by following the (Na+, K+)-ATPase activity as well as the p-nitrophenylphosphatase activity in the presence of Na+ and K+ (with or without ATP). Binding is characterized by a Hill coefficient of 1.0 and a K(0.5(K+))=0.1 to 0.8 mM. The absence of positive or negative cooperativity persists between 5 mM and 100 mM Na+. (b) The analysis of the p-nitrophenylphosphatase or of the 2, 4 dinitrophenylphosphatase activity in the presence of K+ alone indicates the existence of low affinity sites for K+ with positive homotropic interactions. The characteristics of stimulation in that case are, K(0.5)=5 mM, n(H)=1.9. The properties of this family of site(s) are the following: firstly, saturation of the low affinity site(s) by K+ prevents ATP binding to its high affinity internal site. Secondly, saturation of the low affinity sites for K+ prevents binding of Na+ to its internal sites. Thirdly, this family of sites disappears in the presence of ATP, p-nitrophenylphosphate or of both substrates, when Na+ binds to its internal sites. Na+ binding to its specific stimulatory sites provokes the formation of the high affinity type of site for K+. 4. Mg2+ stimulation of the (Na+, K+)-ATPase is characterized by a Hill coefficient n(H(Mg2+))=1.0 and a K(0.5(Mg2+))=1 mM stimulation is essentially a V effect. Heterotropic effects between binding of Mg2+ and substrate to their respective sites are small. Heterotropic interactions between the Ms2+, Na+ and K+ sites are also small. 5. The fluidity of membrane lipids also controls the (Na+, K+)-ATPase activity. Phase transitions or separations in the membrane hardly affect recognition properties of substrates, Na+, K+ and Mg2+ for their respective sites on both sides of the membrane. Only the rate of the catalytic transformation is affected.  相似文献   

4.
Long-chain fatty acid esters of CoA activate (Na+ + K+)-ATPase (the sodium pump) when ATP is suboptimal. To explore the nature of the interactions of these CoA derivatives with the pump, reversible effects of palmitoyl-CoA on the purified membrane-bound kidney enzyme were studied under conditions where interference from the irreversible membrane-damaging effect of the compound was ruled out. With 50 microM ATP, while saturating palmitoyl-CoA increased (Na+ + K+)-ATPase activity, it caused partial inhibition of Na+-ATPase activity without affecting the steady-state level of the phosphoenzyme. Palmitoyl-CoA did not change the K0.5 of ATP for Na+-ATPase, but it altered the complex Na+ activation curve to suggest the antagonism of the low-affinity, but not the high-affinity, Na+ sites. At a low ATP concentration (0.5 microM), K+ inhibited Na+-ATPase as expected. In the presence of palmitoyl-CoA and 0.5 microM ATP, however, K+ became an activator, as it is at high ATP concentrations. The activating effect of palmitoyl-CoA on (Na+ + K+)-ATPase activity was reduced with increasing pH (6.5-8.5), but its inhibitory effect on Na+-ATPase was not altered in this pH range. The data show two distinct actions of palmitoyl-CoA: 1) blockade of the extracellular "allosteric" Na+ sites whose exact role in the control of the pump is yet to be determined, and 2) activation of the pump through increased rate of K+ deocclusion. Since in their latter action the fatty acid esters of CoA are far more effective than ATP at a low-affinity regulatory site, we suggest that these CoA derivatives may be the physiological ligands of this regulatory site of the pump.  相似文献   

5.
P B O'Hara  S H Koenig 《Biochemistry》1986,25(6):1445-1450
A human serum transferrin complex was prepared in which Gd(III) was substituted for Fe(III) at the two metal-binding sites. Characteristic changes upon metal binding in both the UV absorption of ligated tyrosines and the solvent proton longitudinal magnetic relaxation rates demonstrated 2/1 metal stoichiometry and pH-dependent binding constants. Binding studies were complicated both by binding of Gd(III) to nonspecific sites on transferrin at pH less than or equal to 7 and by complexation of the Gd(III) by the requisite bicarbonate anion at pH greater than or equal to 6.0. A unique Gd(III) electron spin resonance spectrum, with a prominent signal at g = 4.96, was observed for the specific Gd(III)-transferrin complex. The major features of this spectrum were fit successfully by a model Hamiltonian which utilized crystal field parameters similar to those determined for Fe(III) in transferrin [Aasa, R. (1970) J. Chem. Phys. 52, 3919-3924]. The magnetic field dependence of the solvent proton relaxation rate was measured as a function of both pH and metal ion concentration. An observed biphasic dependence of the relaxation rate on metal concentration is attributed to either sequential metal binding to the two iron-binding sites with different relaxation properties or random binding to two sites that are similar but show conformationally induced changes in relaxation properties as the second metal is bound. The increase in the solvent proton relaxation rate with pH is consistent with a model in which a proton of a second coordination sphere water molecule is hydrogen bonded to a metal ligand which becomes deprotonated at pH 8.5.  相似文献   

6.
Hydrolysis of adenosine 5'-triphosphate (ATP) and p-nitrophenyl phosphate by the hydrogen ion-transporting potassium-stimulated adenosine triphosphatase (H,K-ATPase) was investigated. Hydrolysis of ATP was studied at pH 7.4 in vesicles treated with the ionophore nigericin. The kinetic analysis showed negative cooperativity with one high affinity (Km1 = 3 microM) and one low affinity (Km2 = 208 microM) site for ATP. The rate of hydrolysis decreased at 2000 microM ATP indicating a third site for ATP. When the pH was decreased to 6.5 the experimental results followed Michaelis-Menten enzyme kinetics with one low affinity site (Km = 116 microM). Higher concentrations than 750 microM ATP were inhibitory. Proton transport was measured as accumulation of acridine orange in vesicles equilibrated with 150 mM KCl. The transport at various concentrations of ATP in the pH interval from 6.0 to 8.0 correlated well with the Hill equation with a Hill coefficient between 1.5-1.9. The concentration of ATP resulting in half-maximal transport rate (S0.5) increased from 5 microM at pH 6.0 to 420 microM at pH 8.0. At acidic pH the rate of proton transport decreased at 1000 microM ATP. The K+-stimulated p-nitrophenylphosphatase (pNPPase) activity resulted in a Hill coefficient close to 2 indicating cooperative binding of substrate. The pNPPase was noncompetitively inhibited by ATP and ADP; half-maximal inhibition was obtained at 2 and 100 microM, respectively. Phospholipase C-treated vesicles lost 80% of the pNPPase activity, but the Hill coefficient did not change. These kinetic results are used for a further development of the reaction scheme of the H,K-ATPase.  相似文献   

7.
Allosteric interactions in the cupro-heme enzyme tryptophan oxygenase (EC 1.13.11.11) of Pseudomonas acidovorans are shown to be pH-dependent. Increasing the assay pH from 6.0 to 8.0 progressively desensitizes the enzyme from both homotropic and heterotropic ligand interactions. This pH-dependent reversible transition has a pK of 6.2. Hill coefficients for the substrate L-tryptophan of 2.0 and 1.4 were measured at pH 6.0 and pH 7.0, respectively. In attempting to identify the enzymatic residue (or residues) responsible for these pH-dependent effects, the enzyme was observed to be irreversibly inactivated by photoinduced oxidation in the presence of the sensitizer, methylene blue. The photoinactivated enzyme showed a loss of one-half its Soret (405 nm) absorption which accompanied the loss of one-half its heme and histidine contents. This first order photoinduced inactivation was pH-dependent and corresponded to a requirement for a protonated species with a pK of 6.2. These results suggest that histidine residues may be involved in the catalytic function and in mediating cooperative interactions of tryptophan oxygenase. Absolute and difference sedimentation velocity analyses indicate that the molecule undergoes a conformational transition when the pH is decreased from pH 8.0 to pH 6.0. This conformational alteration, measured as a 3.9% increase in S20, w can be regarded as an equivalent decrease in the frictional coefficient. If, a more or less spherical shape to the molecule is assumed, then, the 3.9% decrease in the frictional coefficient between pH 8.0 and 6.0 corresponds to a 12% decrease in apparent hydrodynamic volume of the enzyme. Thus, protonation of an enzymatic moiety, possibly histidine, determines both the conformational and functional interactions between enzymatic sites.  相似文献   

8.
Net K movements in reconstituted human red cell ghosts and the resealing of ghosts to cations after osmotic hemolysis of red cells have been studied as functions of the free Ca ion concentration. The Ca-dependent specific increase in K permeability was shown to be mediated by a site close to the internal surface of the membrane with an apparent dissociation constant ap pH 7.2 for Ca (K'p1) of 3-5 X 10(-7) M, for Sr of 7 X 10(-6) M. Ba and Mg did not increase the K-permeability of the membrane but inhibited the Ca-mediated permeability changes. K'D1 decreased in a nonlinear fashion when the pH was increased from 6.0 to 8.5. Two different pK' values of this membrane site were found at pH 8.3 and 6.3. The Ca-activated net K efflux into a K-free medium was almost completely inhibited by an increase in intracellular Na from 4 to 70mM. Extracellular K antagonized this Na effect. Changes in the extracellular Na (0.1-140 mM) or K (0.1-6 mM) concentrations had little effect and did not change K'p1. The Ca-stimulated recovery of a low cation permeability in ghost cells appeared to be mediated by a second membrane site which was accessible to divalent cations only during the process of hemolysis in media of low ionic strength. The apparent dissociation constant for Ca at this site (K'p2) varied between 6 X 10(-7) and 4 X 10(-6) M at pH 7.2 Mg, Sr, and Ba could replace Ca functionally. The selectivity sequence was Ca greater than Sr greater than Ba greater than Mg. K'p2 was independent on the pH value in the range between 6.0 and 8.0 Hill coefficients of 2 were observed for the interaction of Ca with both membrane sites suggesting that more than one Ca ion is bound per site. The Hill cofficients were affected neither by the ion composition nor by the Ph values of the intra-and extracellular media. It is concluded that two different pathways for the permeation of cations across the membrane are controlled by membrane sites with high affinities for Ca: One specific for K, one unspecific with respect to cations. The K-specific "channel" has properties similar to the K channel in excitable tissues.  相似文献   

9.
Cobalt ion inhibits the Ca2+ + Mg2(+)-ATPase activity of sealed sarcoplasmic reticulum vesicles, of solubilized membranes and of the purified enzyme. To use Co2+ appropriately as a spectroscopic ruler to map functional sites of the Ca2+ + Mg2(+)-ATPase, we have carried out studies to obtain the kinetic parameters needed to define the experimental conditions to conduct the fluorimetric studies. 1. The apparent K0.5 values of inhibition of this ATPase are 1.4 mM, 4.8 mM and 9.5 mM total Co2+ at pH 8.0, 7.0 and 6.0, respectively. The inhibition by Co2+ is likely to be due to free Co2+ binding to the enzyme. Millimolar Ca2+ can fully reverse this inhibition, and also reverses the quenching of the fluorescence of fluorescein-labeled sarcoplasmic reticulum membranes due to Co2+ binding to the Ca2+ + Mg2(+)-ATPase. Therefore, we conclude that Co2+ interacts with Ca2+ binding sites. 2. Co2+.ATP can be used as a substrate by this enzyme with Vmax of 2.4 +/- 0.2 mumol ATP hydrolyzed min-1 (mg protein)-1 at 20-22 degrees C and pH 8.0, and with a K0.5 of 0.4-0.5 mM. 3. Co2+ partially quenches, about 10 +/- 2%, the fluorescence of fluorescein-labeled sarcoplasmic reticulum Ca2+ + Mg2(+)-ATPase upon binding to this enzyme at pH 8.0. From the fluorescence data we have estimated an average distance between Co2+ and fluorescein in the ATPase of 1.1-1.8 nm or 1.3-2.1 nm for one or two equidistant Co2+ binding sites, respectively. 4. Co2+.ATP quenches about 20-25% of the fluorescence of fluorescein-labeled Ca2+ + Mg2(+)-ATPase, from which we obtain a distance of 1.1-1.9 nm between Co2+ and fluorescein located at neighbouring catalytic sites.  相似文献   

10.
We report three experiments which show that the hydrolysis of 4-nitrophenyl acetate catalyzed by carbonic anhydrase III from bovine skeletal muscle occurs at a site on the enzyme different than the active site for CO2 hydration. This is in contrast with isozymes I and II of carbonic anhydrase for which the sites of 4-nitrophenyl acetate hydrolysis and CO2 hydration are the same. The pH profile of kcat/Km for hydrolysis of 4-nitrophenyl acetate was roughly described by the ionization of a group with pKa 6.5, whereas kcat/Km for CO2 hydration catalyzed by isozyme III was independent of pH in the range of pH 6.0-8.5. The apoenzyme of carbonic anhydrase III, which is inactive in the catalytic hydration of CO2, was found to be as active in the hydrolysis of 4-nitrophenyl acetate as native isozyme III. Concentrations of N-3 and OCN- and the sulfonamides methazolamide and chlorzolamide which inhibited CO2 hydration did not affect catalytic hydrolysis of 4-nitrophenyl acetate by carbonic anhydrase III.  相似文献   

11.
Modulation of gastric H+,K+-transporting ATPase function by sodium   总被引:3,自引:0,他引:3  
T K Ray  J Nandi 《FEBS letters》1985,185(1):24-28
Gastric H+,K+-ATPase activity is not affected by Na+ at pH 7.0 but is significantly stimulated by Na+ at pH 8.5. For the stimulation at the latter pH, the presence of both Na+ and K+ were essential. Contrary the H+,K+-ATPase, the associated K+-pNPPase was inhibited by Na+ at both pH values. Sodium competes with K+ for the K+-pNPPase reaction. Also, unlike the H+, K+-ATPase activity the ATPase-mediated transport of H+ within the gastric microsomal vesicles was inhibited by Na+. For the latter event only the extravesicular and not the intravesicular Na+ was effective. The data suggest that the K+-pNPPase activity does not represent the phosphatase step of the H+,K+-ATPase reaction. In addition, the observed inhibition of vesicular H+ uptake by Na+ appears to be due to the displacement by Na+ of a cytosolic (extravesicular) H+ site responsible for the vectorial translocation of H+.  相似文献   

12.
Calcium bound to the sarcoplasmic reticulum Ca2+ -ATPase was removed by chelating free calcium ion with EGTA. The kinetic calcium binding reaction to the calcium-unbound ATPase was studied by varying the pH (6.0-8.8) and temperature (0-20 degrees C) at a saturating concentration of 50-100 microM [Ca2+]. At pH 6.0 and 0 degrees C, calcium sites of the enzyme at a rate of t1/2 approximately 10 s. By increasing the pH from 6.0 to 8.8, about half of the total calcium sites were converted from a slow binding state to a rapid binding state (less than 2s). The maximum level was reached at about pH 7.4, and the midpoint of the conversion was observed at about pH 6.7. On the other hand, the slow binding reaction to the other sites was not significantly affected by the pH increase. At pH 7.0 and 20 degrees C, about 90% of the total calcium sites rapidly (less than 2s) bound calcium. The present results suggest that pH and temperature resolve the kinetics of two pools of calcium bound to the Ca2+-ATPase.  相似文献   

13.
The enzyme ketol acid reductoisomerase catalyzes the second common reaction in the biosynthesis of the branched chain amino acids. The reaction is complex as an alkyl migration and a ketone reduction apparently occur as separate steps during the conversion of acetolactate to 2,3-dihydroxy-3-methylbutyrate. This paper reports on the pH dependence of the kinetic parameters of the enzyme. The pH variation of log(V/K) for acetolactate was fit to an equation describing a bell-shaped curve, indicating an acid and a base catalyst for the reaction. In the reverse direction, V/K for 2,3-dihydroxy-3-methylbutyrate is constant over the pH range 8 to 10 and decreases below pH 8 with the ionization of two catalytic groups. The pH dependence of the V/K values for reduction of the kinetically competent intermediate and analogs of this intermediate are also described by a bell-shaped curve. The pH dependence of the V/K for alkyl migration of this intermediate indicates a single base catalyst for this reaction. We observe no deuterium kinetic isotope effect on V or V/K for the reaction of acetolactate at any pH. We observe a pH-dependent kinetic isotope effect on V/K for the reduction of the intermediate, the magnitude of which is metal ion dependent. Larger KIE's are observed in the presence of Mn2+ as opposed to Mg2+. In the reverse reaction there is a pH-dependent kinetic isotope effect on V/K. Based on the pH dependence of the kinetic parameters and the kinetic isotope effects, we propose a base-catalyzed proton shuttle mechanism for the alkyl migration reaction followed by an acid-assisted ketone reduction by NADPH.  相似文献   

14.
The aim of this work was to study the hydroxylation of N, N-dimethyltyramine (DMTA) by tyrosinase in the presence of hydrogen peroxide, a reaction that does not take place without the addition of the hydrogen peroxide. Some properties of this hydroxylating activity are analyzed. The kinetic parameters of mushroom tyrosinase toward hydrogen peroxide (K(m) = 0.5 mM, V(m) = 11 microM/min, V(m)/K(m) = 2.2 x 10(-2) min(-1)) and toward DMTA (K(m) = 0.3 mM, V(m) = 4.8 microM/min, V(m)/K(m) = 16 x 10(-2) min(-1)) were evaluated. There was a lag period, which was similar to the characteristic lag of monophenolase activity at the expense of molecular oxygen. The length of this lag phase decreased with increasing hydrogen peroxide concentration, and disappeared at approximately 0.5 mM H(2)O(2). However, the lag was longer with higher DMTA concentrations. The pH optimum range for this hydroxylating activity was 6.0 to 7.0. The lag also varied with pH, increasing at pH values higher than 6.7. The presence of hydrogen peroxide is necessary for the oxidation of DMTA, as is the presence of active enzyme since the reaction was completely inhibited when selective tyrosinase inhibitors were added.  相似文献   

15.
The urea and heat-induced unfolding-refolding behaviours of chicken egg white ovomucoid and its four fragments representing domains I, II + III, I + II and III were systematically investigated in 0.06 M sodium phosphate buffer (pH 7.0) by difference spectral measurements. The effect of temperature on ovomucoid and its fragments was also studied in 0.05 M sodium acetate buffer (pH 5.0) and in presence of 2 M urea at pH 7.0. Intrinsic viscosity data showed that ovomucoid and its different fragments did not lose any significant amount of their structure under mild acidic conditions (pH 4.6). Difference spectral results showed extensive disruption of the native structure by urea or temperature. Isothermal transitions showed single-step for domain I, domain I + II and domain III, and two-step having one stable intermediate, for ovomucoid and its fragment representing domain II + III. However, the presence of intermediate was not detected when the transitions were studied with temperature at pH 7.0. Strikingly, the single-step thermal transitions of ovomucoid and its fragment representing domain II + III, became two-step when measured either at pH 5.0 or in presence of 2 M urea at pH 7.0. Analysis of the equilibrium data on urea and heat denaturation showed that the second transition observed with ovomucoid or domain II + III represent the unfolding of domain III. The kinetic results of ovomucoid and its fragments indicate that the protein unfolds with three kinetic phases. A comparison of three rate constants for the unfolding of intact ovomucoid with that of its various fragments revealed that domain I, II and III of the protein correspond to the three kinetic phases having rate constants 0.456, 0.120 and 0.054 min-1, respectively. These data have led us to conclude: (i) the unusual stability of ovomucoid towards various denaturants, including temperature, is due to its domain III, (ii) initiation of the folding of the ovomucoid molecule starts from its NH2-terminal region which probably provides the nucleation site for the formation of the subsequent structure and (iii) domains I and II have greater mutual recognition between them as compared to the recognition either of them have with domain III.  相似文献   

16.
Catalytic oxidation of p-cresol by ascorbate peroxidase   总被引:2,自引:0,他引:2  
Transient and steady state kinetics, together with a range of chromatographic and spectroscopic techniques, have been used to establish the mechanism and the products of the H(2)O(2)-dependent oxidation of p-cresol by ascorbate peroxidase (APX). HPLC, GC-MS, and NMR analyses are consistent with the formation of 2, 2'-dihydroxy-5,5'-dimethylbiphenyl (II) and 4alpha,9beta-dihydro-8, 9beta-dimethyl-3(4H)-dibenzofuranone (Pummerer's ketone, III) as the major products of the reaction. In the presence of cumene hydroperoxide, two additional products were observed which, from GC and MS analyses, were shown to be 1,1-dimethylbenzylalcohol (IV) and bis-(1-methyl-1-phenyl-ethyl)-peroxide (V). The product ratio II:III was dependent on enzyme concentration: at low concentrations Pummerer's ketone (III) predominates and at high concentrations formation of the biphenyl compound (II) is favored. Steady-state data showed a sigmoidal dependence on [p-cresol] that was consistent with the presence of 2.01 +/- 0.15 binding sites for the substrate (25.0 degrees C, sodium phosphate, pH 7.0, mu = 2.2 mM) and independent of ionic strength in the range 2.2-500 mM. Single turnover kinetic experiments (pH 7.0, 5.0 degrees C, mu = 0.10 M) yielded second-order rate constants for Compound I reduction by p-cresol, k(2), of 5.42 +/- 0.10 x 10(5) M(-1) s(-1), respectively. Rate-limiting reduction of Compound II by p-cresol, k(3), showed saturation kinetics, giving values for K(d) = 1.54 +/- 0.12 x 10(-3) M and k(3) = 18.5 +/- 0.7 s(-1). The results are discussed in the more general context of APX-catalyzed aromatic oxidations.  相似文献   

17.
Twelve acid hydrolases, 4 near-neutral hydrolases, and alkaline phosphatase were demonstrated in 0.34 M sucrose homogenates of Trypanosoma cruzi strain Y: p-nitrophenylphosphatase and alpha-naphthylphosphatase, with optimum pH at approximately 6.0; alpha=ga;actpsodase. beta=ga;actpsodase. beta=g;icpsodase, N-acetyl-beta-glucosaminidase, cathepsin A and peptidase I and III, with optimum pH between 5.0 and 6.0; and arylsulfatase, cathepsin D, alpha-arabinase and alpha-mannosidase with optimum pH at approximately 4.0. alpha-Glucosidase, glucose-6-phosphatase and peptidase II had optimum pH at approximately 7.0. beta-Glycerophosphatase had a broad pH-activity curve from 4,0 to 7.4, with maximum activity at pH 7.0. The main kinetic characteristics of these enzymes and their quantitative assay methods were studied. No activity was detected for alpha-fucosidase, beta-xylosidase, beta-glucuronidase, elaidate esterase, acid lipase, and alkaline phosphodiesterase.  相似文献   

18.
Adding 15 mM free Mg2+ decreased Vmax of the Na+/K(+)-ATPase reaction. Mg2+ also decreased the K0.5 for K+ activation, as a mixed inhibitor, but the increased inhibition at higher K+ concentrations diminished as the Na+ concentration was raised. Inhibition was greater with Rb+ but less with Li+ when these cations substituted for K+ at pH 7.5, while at pH 8.5 inhibition was generally less and essentially the same with all three cations: implying an association between inhibition and ion occlusion. On the other hand, Mg2+ increased the K0.5 for Na(+)-activation of the Na+/K(+)-ATPase and Na(+)-ATPase reactions, as a mixed inhibitor. Changing incubation pH or temperature, or adding dimethylsulfoxide affected inhibition by Mg2+ and K0.5 for Na+ diversely. Presteady-state kinetic studies on enzyme phosphorylation, however, showed competition between Mg2+ and Na+. In the K(+)-phosphatase reaction catalyzed by this enzyme Mg2+ was a (near) competitor toward K+. Adding Na+ with K+ inhibited phosphatase activity, but under these conditions 15 mM Mg2+ stimulated rather than inhibited; still higher Mg2+ concentrations then inhibited with K+ plus Na+. Similar stimulation and inhibition occurred when Mn2+ was substituted for Mg2+, although the concentrations required were an order of magnitude less. In all these experiments no ionic substitutions were made to maintain ionic strength, since alternative cations, such as choline, produced various specific effects themselves. Kinetic analyses, in terms of product inhibition by Mg2+, require Mg2+ release at multiple steps. The data are accommodated by a scheme for the Na+/K(+)-ATPase with three alternative points for release: before MgATP binding, before K+ release and before Na+ binding. The latter alternatives necessitate two Mg2+ ions bound simultaneously to the enzyme, presumably to divalent cation-sites associated with the phosphate and the nucleotide domains of the active site.  相似文献   

19.
The Eu(III)7F0-->5D0 excitation spectra of the parvalbumins are highly pH-dependent. Below pH 6.0, they exhibit a sharp, partially resolved doublet centered near 5,795 A. However, as the pH is raised, the spectrum becomes increasingly dominated by a much broader signal near 5,784 A. This behavior has been traced to the Eu(III) ion bound at the CD site, but the identity of the moiety undergoing deprotonation remains uncertain. Site-specific mutagenesis studies on the parvalbumin-like protein known as oncomodulin now suggest that the species in question is a liganding serine hydroxyl group. Specifically, replacement of serine-55 by aspartate (the residue present at the corresponding position in the EF site) affords a protein that retains two functional lanthanide binding sites, but fails to undergo the pH-dependent spectral alteration. By contrast, replacement of aspartate-59 by glycine (the corresponding EF site residue) fails to abolish the pH-dependent behavior.  相似文献   

20.
Dihydrofolate reductase from Mycobacterium tuberculosis (MtDHFR) catalyzes the NAD(P)-dependent reduction of dihydrofolate, yielding NAD(P)(+) and tetrahydrofolate, the primary one-carbon unit carrier in biology. Tetrahydrofolate needs to be recycled so that reactions involved in dTMP synthesis and purine metabolism are maintained. In this work, we report the kinetic characterization of the MtDHFR. This enzyme has a sequential steady-state random kinetic mechanism, probably with a preferred pathway with NADPH binding first. A pK(a) value for an enzymic acid of approximately 7.0 was identified from the pH dependence of V, and the analysis of the primary kinetic isotope effects revealed that the hydride transfer step is at least partly rate-limiting throughout the pH range analyzed. Additionally, solvent and multiple kinetic isotope effects were determined and analyzed, and equilibrium isotope effects were measured on the equilibrium constant. (D(2)O)V and (D(2)O)V/K([4R-4-(2)H]NADH) were slightly inverse at pH 6.0, and inverse values for (D(2)O)V([4R-4-(2)H]NADH) and (D(2)O)V/K([4R-4-(2)H]NADH) suggested that a pre-equilibrium protonation is occurring before the hydride transfer step, indicating a stepwise mechanism for proton and hydride transfer. The same value was obtained for (D)k(H) at pH 5.5 and 7.5, reaffirming the rate-limiting nature of the hydride transfer step. A chemical mechanism is proposed on the basis of the results obtained here.  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