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.
Inorganic pyrophosphatase must bind two phosphate molecules in order to catalyze pyrophosphate synthesis. In this report it is shown that Pi causes marked effect on the absorption spectrum of baker's yeast inorganic pyrophosphatase and this effect can be used to analyze Pi binding to this enzyme. A series of absorbance versus Pi concentration curves in the presence of 0.5-20 mM free Mg2+ were obtained at pH 7.2 and computer-fitted to 19 models. The dissociation constant of magnesium phosphate (8.5 +/- 0.4 mM) used in this analysis was measured with a Mg2+-sensitive electrode. The best model implies successive binding of two magnesium phosphate molecules or random-order binding of magnesium phosphate and free phosphate molecules. The first route predominates at physiological concentrations of Mg2+. The Pi-inhibition pattern of pyrophosphate hydrolysis confirmed that Pi adds to the active site and provided further evidence for the existence of an activating Pi-binding site. The possibility is raised that the pathways of pyrophosphate synthesis and hydrolysis by inorganic pyrophosphatase may differ in the sense that the binding of the fourth metal ion/subunit may facilitate the synthesis and inhibit the hydrolysis.  相似文献   

2.
A scheme of Mg2+ and Pi binding to yeast inorganic pyrophosphatase has been deduced from the concentration dependencies of the rate of oxygen exchange between Pi and water. The exchange reaction requires the binding of MgPi and free Pi (pathway I) or two MgPi (pathway II) in addition to two Mg2+ ions bound in the absence of Pi. Pathway II predominates above 0.16 mM Mg2+. The rate of formation of bound PPi from bound Pi for pathway II is three times as high as that for pathway I. The results suggest that the binding of the fourth Mg2+ ion to pyrophosphatase stimulates its synthetic vs its hydrolytic capability.  相似文献   

3.
1. An active monomeric form of inorganic pyrophosphatase from baker's yeast was prepared by maleylation of the protein at pH 10.5. 2. The dimeric and monomeric pyrophosphatase bound at non-catalytic sites 0.5 and 1.0 mol of slowly dissociating Pi per mol subunit, respectively. This stoichiometry was not affected on active site blockage with PPi. 3. Added Pi accelerated the dissociation of Pi from the dimeric but not monomeric enzyme. 4. Our results indicate a strong interaction to occur between the non-catalytic sites of two subunits of native pyrophosphatase which results in diminished stability of Pi binding to one of them.  相似文献   

4.
Conyers GB  Wu G  Bessman MJ  Mildvan AS 《Biochemistry》2000,39(9):2347-2354
Recombinant IalA protein from Bartonella bacilliformis is a monomeric adenosine 5'-tetraphospho-5'-adenosine (Ap4A) pyrophosphatase of 170 amino acids that catalyzes the hydrolysis of Ap4A, Ap5A, and Ap6A by attack at the delta-phosphorus, with the departure of ATP as the leaving group [Cartwright et al. (1999) Biochem. Biophys. Res. Commun. 256, 474-479]. When various divalent cations were tested over a 300-fold concentration range, Mg2+, Mn2+, and Zn2+ ions were found to activate the enzyme, while Ca2+ did not. Sigmoidal activation curves were observed with Mn2+ and Mg2+ with Hill coefficients of 3.0 and 1.6 and K0.5 values of 0.9 and 5.3 mM, respectively. The substrate M2+ x Ap4A showed hyperbolic kinetics with Km values of 0.34 mM for both Mn2+ x Ap4A and Mg2+ x Ap4A. Direct Mn2+ binding studies by electron paramagnetic resonance (EPR) and by the enhancement of the longitudinal relaxation rate of water protons revealed two Mn2+ binding sites per molecule of Ap4A pyrophosphatase with dissociation constants of 1.1 mM, comparable to the kinetically determined K0.5 value of Mn2+. The enhancement factor of the longitudinal relaxation rate of water protons due to bound Mn2+ (epsilon b) decreased with increasing site occupancy from a value of 12.9 with one site occupied to 3.3 when both are occupied, indicating site-site interaction between the two enzyme-bound Mn2+ ions. Assuming the decrease in epsilon(b) to result from cross-relaxation between the two bound Mn2+ ions yields an estimated distance of 5.9 +/- 0.4 A between them. The substrate Ap4A binds one Mn2+ (Kd = 0.43 mM) with an epsilon b value of 2.6, consistent with the molecular weight of the Mn2+ x Ap4A complex. Mg2+ binding studies, in competition with Mn2+, reveal two Mg2+ binding sites on the enzyme with Kd values of 8.6 mM and one Mg2+ binding site on Ap4A with a Kd of 3.9 mM, values that are comparable to the K0.5 for Mg2+. Hence, with both Mn2+ and Mg2+, a total of three metal binding sites were found-two on the enzyme and one on the substrate-with dissociation constants comparable to the kinetically determined K0.5 values, suggesting a role in catalysis for three bound divalent cations. Ca2+ does not activate Ap4A pyrophosphatase but inhibits the Mn2+-activated enzyme competitively with a Ki = 1.9 +/- 1.3 mM. Ca2+ binding studies, in competition with Mn2+, revealed two sites on the enzyme with dissociation constants (4.3 +/- 1.3 mM) and one on Ap4A with a dissociation constant of 2.1 mM. These values are similar to its Ki suggesting that inhibition by Ca2+ results from the complete displacement of Mn2+ from the active site. Unlike the homologous MutT pyrophosphohydrolase, which requires only one enzyme-bound divalent cation in an E x M2+ x NTP x M2+ complex for catalytic activity, Ap4A pyrophosphatase requires two enzyme-bound divalent cations that function in an active E x (M2+)2 x Ap4A x M2+ complex.  相似文献   

5.
A comparative study of the orthophosphate-pyrophosphate exchange reaction catalyzed by the soluble pyrophosphatase from baker's yeast and by the membrane-bound pyrophosphatase of Rhodospirillum rubrum chromatophores was performed. In both systems the rate of exchange increased when the pH of the medium was raised from 6.0 to 7.8 and when the MgCl2 concentration was raised from 0.1 mM to 20 mM. For the yeast pyrophosphatase the exchange rates measured at different pH values and in the presence of 6.7 to 8.8 mM free Mg2+ superimposed as a single curve when plotted as a function of the concentrations of either HPO4(2-) or MgHPO4. This was not observed with the use of R. rubrum chromatophores. With yeast pyrophosphatase, the Km for Pi was higher than 10 mM and could not be measured when the free Mg2+ concentration in the medium was lower than 0.5 mM. There was a decrease in the Km for Pi when the free Mg2+ concentration was raised to 6.7-8.8 mM or when, in the presence of low free Mg2+, the organic solvents dimethylsulfoxide (20% v/v) or ethyleneglycol (40% v/v) were included in the assay medium. In the presence of 6.7-8.8 mM free Mg2+ the Km for total Pi was 7 mM at pH 7.0 and 12 mM at pH 7.8. For the ionic species HPO4(2-) and MgHPO4, the Km values were 5.8 mM and 4.2 mM respectively. In the presence of 0.24-0.42 mM free Mg2+ and either 20% (v/v) dimethylsulfoxide or 40% (v/v) ethyleneglycol the Km values for total Pi, HPO4(2-) and MgHPO4 were 7.6, 3.5 and 0.5 mM respectively. With R. rubrum chromatophores, the Km for Pi in the presence of 5.5-7.5 mM free Mg2+ was very high and could not be measured. In the presence of 0.24-0.45 mM free Mg2+ the ratio between the velocities of hydrolysis and synthesis of pyrophosphate measured at pH 7.8 with yeast pyrophosphatase and chromatophores of R. rubrum were practically the same. When the free Mg2+ concentration was raised to 5.5-8.8 mM this ratio decreased from 1028 to 540 when the yeast pyrophosphatase was used and from 754 to 46 when chromatophores were used.  相似文献   

6.
We have developed two methods for quantitatively measuring inorganic pyrophosphate (PPi) in the presence of 10(3)--10(4) molar excesses of inorganic phosphate (Pi) and used them to measure the extent of enzyme-bound pyrophosphate (EPPi) formation in solutions of yeast inorganic pyrophosphatase and Pi. We have also measured the rate of enzyme-catalyzed H2O--phosphate oxygen exchange. We find both processes to have essentially identical dependence on Mg2+ and Pi concentrations, thus providing important confirmation for the recent proposal by Janson et al. (1979) that oxygen exchange proceeds via EPPi formation. Our results are consistent with a model in which three Mg2+ per active site are required for EPPi formation but inconsistent with a model requiring only two Mg2+ per active site and permit the formulation of an overall scheme for inorganic pyrophosphatase catalysis of PPi--Pi equilibration as well as the evaluation of equilibrium and rate constants in this scheme. The major results and conclusions of our work are the following: (a) the equilibrium constant for PPi (enzyme-bound) in equilibrium with 2Pi (enzyme-bound) is 4.8; (b) following PPi hydrolysis, the first released Pi contains an oxygen from solvent water; (c) the steps for PPi hydrolysis on the enzyme and for release of both product Pi's are all partially rate determining in overall enzyme-catalyzed PPi hydrolysis; (d) PPi formation on the enzyme is rate determining for H2O--Pi oxygen exchange; (e) PPi dissociation from the enzyme is very slow and is the rate-determining step in Pi--PPi exchange (Cohn, 1958; Janson et al., 1979). This also accounts for the observation that the calculated dissociation constant for MgPPi complex binding to enzyme is considerably lower than the measured Km for enzyme-catalyzed MgPPi hydrolysis.  相似文献   

7.
Inhibition of (Na+ + K+)-dependent adenosine triphosphatase phosphatase by vanadate is thought to occur through the tight binding of vanadate to the same site from which Pi is released. To see if ATP binds to [48V] vanadate-enzyme complex, just as it does to the phosphoenzyme, the effects of Na+, K+, and ATP on the dissociation rate of the complex at 10 degrees C were studied. The rate constant was increased by Na+, and this increase was blocked by K+, indicating that either Na+ or K+ binds to the complex. ATP alone, or in combination with K+, had no effect on the rate constant. In the presence of Na+, however, ATP caused a further increase in the rate constant. The value of K0.5 of Na+ was the same in the presence or absence of ATP; K0.5 of ATP (0.2 mM) did not seem to change significantly when Na+ concentration was varied, and K0.5 of K+, at a constant Na+ concentration, was the same in the presence or absence of ATP. The data indicate that ATP binds to the enzyme-vanadate complex regardless of the presence or absence of Na+ or K+, but it affects the dissociation rate only when Na+ is bound simultaneously. The value of K0.5 of Na+ decreased as pH was increased in the range of 6.5-7.8, but K0.5 of ATP was independent of pH. Demonstration of ATP binding to the enzyme-vanadate complex provides further support for the suggestion that the oligomeric enzyme contains a low-affinity regulatory site for ATP that is distinct from the interacting high-affinity catalytic sites.  相似文献   

8.
Hiromasa Y  Roche TE 《Biochemistry》2008,47(8):2298-2311
In the complete absence of K+ and phosphate (Pi), pyruvate dehydrogenase kinase isoform 2 (PDHK2) was catalytically very active but with an elevated Km for ATP, and this activity is insensitive to effector regulation. We find that K+ or 5-fold lower levels of NH4+ markedly enhanced quenching of Trp383 fluorescence of PDHK2 by ADP and ATP. K+ binding caused an approximately 40-fold decrease in the equilibrium dissociation constants (Kd) for ATP from approximately 120 to 3.0 microM and an approximately 25-fold decrease in Kd for ADP from approximately 950 to 38 microM. Linked reductions in Kd of PDHK2 for K+ were from approximately 30 to approximately 0.75 mM with ATP bound and from approximately 40 to approximately 1.7 mM with ADP bound. Without K+, there was little effect of ADP on pyruvate binding, but with 100 mM K+ and 100 microM ADP, the L0.5 of PDHK2 for pyruvate was reduced by approximately 14 fold. In the absence of K+, Pi had small effects on ligand binding. With 100 mM K+, 20 mM Pi modestly enhanced binding of ADP and hindered pyruvate binding but markedly enhanced the binding of pyruvate with ADP; the L0.5 for pyruvate was specifically decreased approximately 125-fold with 100 microM ADP. Pi effects were minimal when NH4+ replaced K+. We have quantified coupled binding of K+ with ATP and ADP and elucidated how linked K+ and Pi binding are required for the potent inhibition of PDHK2 by ADP and pyruvate.  相似文献   

9.
The rate of spontaneous efflux of Ca from liver mitochondria incubated in the absence of ATP and Mg increases with time and is associated with a synchronous collapse of membrane potential and with Pi efflux. In the presence of Mg and ATP the ruthenium-red-induced Ca efflux does not change with time. The activity of the Ca efflux pathway in Pi-depleted mitochondria is 15-fold greater than in mitochondria equilibrated with 3.3 mM Pi. 50% inhibition is caused by 0.3 mM Pi. The membrane potential is not affected by changes in Pi concentration, although the steady-state extra-mitochondrial free Ca concentration reflects the alterations in efflux rate. In the presence of Pi, the ruthenium-red-induced efflux rate is independent of the total matrix Ca content; however in Pi-depleted mitochondria, with acetate substituting as permeant anion, the efflux rate increases with total matrix Ca content. The lowered efflux rate in the presence of Pi is not due to a limitation in the rate of dissociation of the matrix Ca-phosphate complex. The efflux pathway is activated by a lowered membrane potential, but the relative effect of Pi is retained. Under the present conditions Na slightly inhibits the efflux rate. The lack of an effect of total matrix Ca content on the efflux rate in the presence of Pi is used as the basis of a highly accurate determination of the activity of the Ca uniporter as a function of external free Ca concentration.  相似文献   

10.
Carbamoyl phosphate synthethase I synthesizes carbamoyl phosphate from ammonia, HCO3- and two molecules of ATP, one of which, ATPA, yields Pi while the other, ATPB, yields the phosphoryl group of carbamoyl phosphate. Pulse-chase experiments with [gamma-32P]ATP without added HCO3- demonstrate separate binding sites for ATPA and ATPB. Bound ATPA dissociates readily from its site (t1/2 approximately 1--2 s) and the Kd is 0.2--0.7 mM. For the ATPB binding site the t1/2 for dissociation is 5--12 s and the Kd approximately 10 mM. Kd for ATPA seems to increase with enzyme concentration whereas Kd for ATPB does not change. HClO4 releases the ATP unchanged from the enzyme . ATPB and enzyme . ATPB . ATPA complexes. In the presence of HCO3-, ATP and N-acetylglutamate, an enzyme . ATPB . HCO3- . ATPA complex is formed. Its formation by the addition of HCO3- to the enzyme . ATPB . ATPA complex appears to involve an initial bimolecular addition reaction followed by an isomerization. Treatment with HClO4 releases Pi from ATPA but ATPB is released unchanged. Spontaneous hydrolysis of ATPA is responsible for the ATPase activity of the enzyme. Thus, a covalent bond may form between HCO3- and ATPA. However, ATPA can dissociate rapidly (t1/2 less than 10 s). The Kd for ATPA is approximately 0.2 mM. ATPB appears unable to dissociate from the enzyme . ATPB . HCO3- . ATPA complex since the t1/2 for dissociation of ATPB from the enzyme is lengthened about five times in the presence of 19 mM HCO3- and at 1 mM ATP. ATPA may also hydrolyse in this complex and be replaced by another molecule of ATP in the absence of exchange of ATPB. However, the ATPA binding site must be occupied to prevent ATPB release. ATPB may be bound in a pocket which becomes inaccessible to the solution when HCO3- and ATPA also bind. In contrast, HCO3- does not inhibit the binding of ATPB to the enzyme. Various intermediate steps in the formation of the enzyme . ATPb . HCO3- . ATPA complex are discussed. Additional evidence is presented that the ATPB binding site is only periodically accessible to ATP in solution and that ATPB in the steady-state reaction binds when the products leave. Since greater than 1.3 mol ATPB and greater than 1.8 mol ATPA bind/mol enzyme dimer, the enzyme monomer may be an active species.  相似文献   

11.
Na+,K+-ATPase has been purified from lamb kidney and consists of two polypeptide peaks on polyacrylamide gel electrophoresis with an enzyme activity of 1,000 mumole Pi/mg pro per hr. A scheme depicting the interaction of cardiac glycoside with the enzyme and ligand effects on binding has been constructed. Under all ligand conditions, ouabain binding tends to reach the same maximum if sufficient ouabain is present. Initial rates vary with ligand conditions. Using a chase method, the rate of dissociation of the glycoside from the enzyme is not influenced by the ligands present, although with separation of the enzyme-glycoside complex from the binding medium, differences are noted. The effect of ouabain on Na binding demonstrated two classes of sites, KD = 0.2 mM and KD = 18 mM. Denaturation decreased the high affinity sites. There was also a good correlation between ouabain binding and inhibition of Na binding. Clearly, ligands are critical in regulating cardiac glycoside interaction with the enzyme.  相似文献   

12.
The binding of mouse epidermal growth factor-urogastrone (EGF-URO) to membranes from term human placenta is peptide-specific, saturable (about 20 pmol of EGF-URO bound maximally/mg of protein), reversible, and of high affinity (KD about 400 pM). Optimal binding is observed at pH 7.6. At low pH (3.5 to 5.0). EGF-URO can be reversibly dissociated from the receptor; however, exposure to pH < 3 irreversibly inactivates the receptor. The binding, which does not exhibit ligand cooperativity, exhibits an association rate constant of 6.1 x 10(-4) s-1 and a dissociation rate constant of 6.1 x 10(-4) s-1. The dissociation constant determined from the rate constants, 240 pM, is in reasonable agreement with the constant estimated by equilibrium methods. Both monovalent and divalent cations augment EGF-URO binding 2- to 3-fold. Although in general, divalent cations enhance binding at lower concentrations (optimum, 5 mM) than do monovalent cations (optimum, approximately 80 mM), there is no cation-specific effect. Neither guanine nor adenine nucleotides affect EGF-URO binding. Whereas the proteolytic enzymes (trypsin, chymotrypsin, papain, and pepsin) inactivate the receptor, neuraminidase and phospholipases A2, C, and D augment EGF-URO binding. Neuraminidase increases the number of available sites without affecting ligand affinity. Wheat germ agglutinin, concanavalin A, and phytohemagglutinin all compete for the binding of EGF-URO. The data complement previous observations of EGF-URO binding obtained in intact cells and provide a basis for the solubilization, characterization, and isolation of this receptor from a rich tissue source.  相似文献   

13.
D J Hamm  B S Cooperman 《Biochemistry》1978,17(19):4033-4040
Yeast inorganic pyrophosphatase is a dimer of identical subunits. Previous work (Rapoport, T.A., et al. (1973) Eur. J. Biochem. 33, 341) indicated the presence of two different Mn2+ binding sites per subunit. In the present work, the binding of inorganic phosphate to the Mn2+-inorganic pyrophosphatase complex has been studied by 1H and 31P nuclear magnetic resonance. Two distinct phosphate sites have been found, having dissociation constants of 0.24 mM and 18 mM. The Mn2+-31P distance from tightly bound Mn2+ to phosphate bound in the low affinity site (6.2 A) is consistent with outer sphere binding. Binding to both phosphate sites can be simultaneously inhibited by the pyrophosphate analogue, hydroxymethanebisphosphonate, providing evidence for the physical proximity of these two sites. The weaker Mn2+ site is apparently far from both phosphate sites. From the magnitudes of the dissociation constants found for both phosphate and analogue binding and the recent work of P.D. Boyer and his co-workers (private communication) on enzyme-catalyzed phosphate-water exchange, it appears unlikely that the hydrolysis of enzyme-bound pyrophosphate is the rate-determining step in the overall enzymatic catalysis of pyrophosphate hydrolysis, at least when Mn2+ is the required divalent metal ion cofactor.  相似文献   

14.
L de Meis  G Inesi 《Biochemistry》1985,24(4):922-925
Sarcoplasmic reticulum ATPase is phosphorylated by ATP in the presence of calcium, with a consequent reduction of the affinity of the binding sites for calcium and dissociation of the divalent cation from the enzyme. ATPase phosphorylation with Pi, on the other hand, requires prior removal of calcium from the enzyme, indicating that the energy requirement for phosphorylation of the enzyme-calcium complex can be met by ATP but not by Pi. We find that when the energy yield of the Pi reaction with the enzyme is increased by the addition of dimethyl sulfoxide to the medium, ATPase phosphorylation with Pi occurs even in the presence of calcium, and the binding sites undergo a reduction in affinity with consequent dissociation of Ca2+ from the enzyme, in analogy to the effect of ATP. It is thereby demonstrated experimentally that an essential step in the coupling of catalytic and transport activities is an interdependence and mutual ligand exclusion of the phosphorylation and calcium sites, in which ATP does not play a direct role. An important difference between the effects of ATP and Pi is that the former produces dissociation of Ca2+ inside the vesicles as the result of advancement of the catalytic cycle in the forward direction, while Pi produces dissociation of calcium into the outer medium as a consequence of equilibration of enzyme states producing a shift in the reverse direction of the enzyme cycle. These observations demonstrate how equilibration of intermediate enzyme states determines extent and direction of overall reaction flow.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Binding of 21 S dynein ATPase isolated from Tetrahymena cilia to B subfibers of microtubule doublets was used as a model system to study dynein-tubulin interactions and their relationship to the microtubule-based sliding filament mechanism. Binding of 21 S dynein to both A and B microtubule subfibers is supported by monovalent as well as divalent ions. Monovalent cation chlorides support dynein binding to B subfibers with the specificity Li greater than Na congruent to K congruent to Rb congruent to Cs congruent to choline. The corresponding sodium or potassium halides follow the order F greater than Cl greater than Br greater than I. However, an optimal binding concentration of 40 mM KCl supports only 55% of the protein binding which takes place in 3 mM MgSO4 and does not stabilize dynein cross-bridges when whole axonemes are fixed for electron microscopy. Divalent metal ion chlorides (MgCl2, CaCl2, SrCl2, and BaCl2) have nearly equivalent effects at a concentration of 6 mM; all support about 140% of the binding observed in 6 mM MgSO4. The binding data suggest negative cooperativity or the presence of more than one class of dynein binding sites on the microtubule lattice. Low concentrations of MgATP2- induce dissociation of dynein bound to B subfibers in either 6 mM MgSO4 or 40 mM KCl. ADP, Pi, PPi, and AMP-PCH2P are unable to induce dynein dissociation, while AMP-PNHP and ATP4- both cause dynein release from B subfiber sites. The half-maximal sensitivities of the tubulin-dynein complex to MgATP2-, ATP4-, and adenylyl-imidodiphosphate (AMP.PNP) are 1.3 X 10(-8) M, 3.6 X 10(-5) M, and 4.7 X 10(-4) M respectively. Incubation of doublets or 21 S dynein in N-ethylmaleimide (NEM), which can inhibit active sliding, has no effect on either association of dynein with the B subfiber or on dissociation of the resulting dynein-B subfiber complex by MgATP2-.  相似文献   

16.
The oxygen exchange parameters for the hydrolysis of ATP by the F1-ATPase have been determined over a 140,000-fold range of ATP concentrations and a 5,000-fold range of reaction velocity. The average number of water oxygens incorporated into each Pi product ranges from a limit of about 1.02 at saturating ATP concentrations to a limit of about 3.97 at very low ATP concentrations. The latter value represents 400 reversals of hydrolysis of bound ATP prior to Pi dissociation. In accord with the binding change mechanism, this means that ATP binding at one catalytic site increases the off constant of Pi and ADP from another catalytic site by at least 20,000-fold, equivalent to the use of 6 kcal mol-1 of ATP binding energy to promote product release. The estimated rate of reversal of hydrolysis of F1-ATPase-bound ATP to bound ADP + Pi varies only about 5-fold with ATP concentration. The rate is similar that observed previously for reversal of bound ATP hydrolysis or synthesis with the membrane-bound enzyme and is greater than the rate of net ATP formation during oxidative phosphorylation. This adds to evidence that energy input or membrane components are not required for bound ATP synthesis.  相似文献   

17.
Our previous work (Carlier, M.-F., and Pantaloni, D. (1986) Biochemistry 25, 7789-7792) had shown that F-ADP-Pi-actin is a major intermediate in ATP-actin polymerization, due to the slow rate of Pi release following ATP cleavage on filaments. To understand the mechanism of ATP-actin polymerization, we have prepared F-ADP-Pi-actin and characterized its kinetic parameters. 32Pi binds to F-ADP-actin with a stoichiometry of 1 mol/mol of F-actin subunit and an equilibrium dissociation constant Kpi of 1.5 mM at pH 7.0 Kpi increases with pH, indicating that the H2PO-4 species binds to F-actin. ADP-Pi-actin subunits dissociate much more slowly from filament ends than ADP-actin subunits; therefore, the stability of filaments in ATP is due to terminal ADP-Pi subunits. The slow rate of dissociation of ADP-Pi-actin also explains the decrease in critical concentration of ADP-actin in the presence of Pi reported by Rickard and Sheterline (Richard, J. E., and Sheterline, P. (1986) J. Mol. Biol. 191, 273-280). The effect of Pi on the rate of actin dissociation from filaments is much more pronounced at the barbed end than at the pointed end. Using gelsolin to block the barbed end, we have shown that the two ends are energetically different in the presence of ATP and saturating Pi, but less different than in the absence of Pi. The results are interpreted within a new model for actin polymerization. It is possible that phosphate binding to F-actin can regulate motile events in muscle and nonmuscle cells.  相似文献   

18.
The thermodynamics of the drug-inhibitors acyclovir, ganciclovir, and 9-benzylguanine binding to human purine nucleoside phosphorylase (hsPNP) were determined from isothermal titration calorimetry as a function of the substrate phosphate ion (Pi) concentration from 0 to 0.125 M and temperature from 15 °C to 35 °C. At 25 °C and with an increase in the Pi concentration from 0 to 50 mM, acyclovir binding becomes more entropically-driven and ganciclovir binding becomes more enthalpically-driven. At 25 °C, the tighter 9-benzylguanine binding reaction goes from an enthalpically-driven reaction in the absence of Pi to an entropically-driven reaction at 10 mM Pi, and the enthalpically-driven nature of the binding reaction is restored at 75 mM Pi. Since the dependencies of the driving-nature of the binding reactions on Pi concentration can be simulated by Pi binding to its catalytic site, it is believed that bound Pi affects the interactions of the side-chains with the ribose catalytic site. However, the binding constants are unaffected by change in the bound Pi concentration because of enthalpy-entropy compensation. The enzymatic activity of hsPNP was determined by an ITC-based assay employing 7-methylguanosine and Pi as the substrates. The heat of reaction determined from the assay increased by 7.5 kJ mol−1 with increase in Pi concentration from 50 to 100 mM and is attributed to weak binding of the Pi to a secondary regulatory site. Although the binding constants of acyclovir and ganciclovir at 20 μM hsPNP were in agreement with the inverse inhibition constants determined from the ITC enzyme inhibition assays at 60 nM, the binding constant of 9-benzylguanine, which interacts with Phe159 from an adjacent subunit, decreased from 5.62 × 105 M−1 to 1.14 × 105 M−1. This reduction in the 9-benzylguanine binding affinity along with a 7-fold increase in the specific activity of hsPNP at 14.5 nM results from partial dissociation of the hsPNP trimer into monomers below the 60 nM level.  相似文献   

19.
Alpha-D-Glucopyranosyl fluoride was found to inhibit strongly the action of alpha-glucan phosphorylase b[EC 2.4.1.1] from rabbit muscle, and that of the enzyme from potato tubers rather weakly. The inhibition is highly specific, being competitive with respect to glucose 1-phosphate and noncompetitive with respect to polysaccharide, during polysaccharide synthesis. In the reverse process, it is competitive with respect to Pi. These results have been explained by assuming that the inhibitor binds to the glucose 1-phosphate site of the enzyme, occupying both subsites which normally bind the glucosyl and phosphate moities of the substrate, but does not directly interact with the polysaccharide site. Based on this assumption, the dissociation constants of the enzyme-inhibitor and enzyme-polysaccharide-inhibitor complexes have been evaluated (0.43 and 0.20 mM for the muscle enzyme, respectively; 24 and 23 mM for the potato enzyme, respectively). Glucosyl fluoride also acts as a noncompetitive inhibitor with respect to AMP. A high concentration of AMP causes an inhibitory effect on the action of the muscle enzyme, the effect being menifested in the presence of glucosyl fluoride.  相似文献   

20.
cGMP-dependent protein kinase binds 4 mol cGMP/mol enzyme to two different sites. Binding to site 1 (apparent Kd 17 nM) shows positive cooperativity and is inhibited by Mg . ATP, whereas binding to site 2 (apparent Kd 100-150 nM) is non-cooperative and not affected by Mg . ATP. Autophosphorylation of the enzyme abolishes the cooperative binding to site 1 and the inhibitory effect of Mg . ATP. The association (K1) and dissociation (K-1) rate constant for site 2 and K1 for site 1 are not affected significantly by Mg . ATP or autophosphorylation. The dissociation rate from site 1 measured in the presence of 1 mM unlabelled cGMP is decreased threefold and over tenfold by Mg . ATP and autophosphorylation, respectively. In contrast, the dissociation rate from site 1 measured after a 500-fold dilution of the enzyme-ligand complex is 100-fold faster than that determined in the presence of 1 mM cGMP and is only slightly influenced by Mg . ATP or autophosphorylation. Only Kd values calculated with the latter K-1 values are similar to the Kd values obtained by equilibrium binding. These results suggest that autophosphorylation of cGMP-dependent protein kinase affects mainly the binding characteristics of site 1.  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