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.
The kinetics of ATP hydrolysis and tension responses were studied simultaneously in a permeabilized preparation of cardiac tissue of the guinea pig. This was achieved by combining laserflash photolysis of P3-1-(2-nitrophenyl)ethyladenosine 5'-triphosphate ("caged-ATP") and a rapid freezing technique. In the presence of calcium ions, tension increased following the photolytic production of ATP with a half-time of 0.3 s. The timecourse of ATP hydrolysis consisted of an initial rapid phase followed by a steady-state hydrolysis rate of 0.4 s-1, indicating that the rate-limiting step of the ATPase in isometric fibers is slower and subsequent to the nucleotide hydrolysis step: the isometric steady state intermediate is probably an actomyosin-ADP complex. In the absence of calcium ions, rigor tension decreased upon the photolytic production of ATP with a half-time of 0.45 s. The time course of ATP hydrolysis was biphasic with a rapid initial phase of ATP hydrolysis, followed by a steady-state hydrolysis rate which was too slow to measure over the time scale of these experiments (less than 0.04 s-1). A comparison of the results obtained in this study with those reported for rabbit skeletal muscle reveals qualitative similarities between cardiac and skeletal muscle and also quantitative differences in their physiological and kinetic behavior.  相似文献   

2.
S Sadis  L E Hightower 《Biochemistry》1992,31(39):9406-9412
The mammalian 70-kilodalton heat shock cognate protein (Hsc70) is an abundant, cytosolic molecular chaperone whose interactions with protein substrates are regulated by ATP hydrolysis. In vitro, purified Hsc70 was found to have a slow, intrinsic ATPase activity in the absence of protein substrates. The addition of an unfolded protein such as apocytochrome c stimulated ATP hydrolysis 2-3-fold. In contrast, the native holoprotein, cytochrome c, did not stimulate the ATPase rate, in accord with recent observations that 70-kilodalton heat shock proteins interact selectively with unfolded proteins. Stimulation of ATP hydrolysis by apocytochrome c was due to an increase in the Vmax, with no effect on the Km for ATP. Following hydrolysis of [3H]ATP, a relatively stable [3H]ADP.Hsc70 complex was formed. Release of [3H]ADP from Hsc70 was most efficient in the presence of other nucleotides such as ADP or ATP, suggesting that ADP release occurs as an ADP/ATP exchange reaction. The loss of radiolabeled ADP from Hsc70 in the presence of exogenous nucleotides followed first-order kinetics. In the presence of nucleotides, apocytochrome c induced a 2-fold increase in the rate of ADP release from Hsc70. Moreover, rate constants of the nucleotide exchange reaction measured in the absence and presence of apocytochrome c (0.16 and 0.34 min-1, respectively) closely matched the kcat values derived from ATP hydrolysis measurements (0.15 and 0.38 min-1, respectively). The results suggest that ADP release in a rate-limiting step in the Hsc70 ATPase reaction and that unfolded proteins stimulate ATP hydrolysis by accelerating the rate of ADP/ATP exchange.  相似文献   

3.
D Wu  P D Boyer 《Biochemistry》1986,25(11):3390-3396
When the heat-activated chloroplast F1 ATPase hydrolyzes [3H, gamma-32P]ATP, followed by the removal of medium ATP, ADP, and Pi, the enzyme has labeled ATP, ADP, and Pi bound to it in about equal amounts. The total of the bound [3H]ADP and [3H]ATP approaches 1 mol/mol of enzyme. Over a 30-min period, most of the bound [32P]Pi falls off, and the bound [3H]ATP is converted to bound [3H]ADP. Enzyme with such remaining tightly bound ADP will form bound ATP from relatively high concentrations of medium Pi with either Mg2+ or Ca2+ present. The tightly bound ADP is thus at a site that retains a catalytic capacity for slow single-site ATP hydrolysis (or synthesis) and is likely the site that participates in cooperative rapid net ATP hydrolysis. During hydrolysis of 50 microM [3H]ATP in the presence of either Mg2+ or Ca2+, the enzyme has a steady-state level of about one bound [3H]ADP per mole of enzyme. Because bound [3H]ATP is also present, the [3H]ADP is regarded as being present on two cooperating catalytic sites. The formation and levels of bound ATP, ADP, and Pi show that reversal of bound ATP hydrolysis can occur with either Ca2+ or Mg2+ present. They do not reveal why no phosphate oxygen exchange accompanies cleavage of low ATP concentrations with Ca2+ in contrast to Mg2+ with the heat-activated enzyme. Phosphate oxygen exchange does occur with either Mg2+ or Ca2+ present when low ATP concentrations are hydrolyzed with the octyl glucoside activated ATPase. Ligand binding properties of Ca2+ at the catalytic site rather than lack of reversible cleavage of bound ATP may underlie lack of oxygen exchange under some conditions.  相似文献   

4.
The H(+)-ATPase from chloroplasts, CF(0)F(1), was isolated and purified. The enzyme contained one endogenous ADP at a catalytic site, and two endogenous ATP at non-catalytic sites. Incubation with 2-azido-[alpha-(32)P]AD(T)P leads to a tight binding of the azido-nucleotides. Free nucleotides were removed by three consecutive passages through centrifugation columns, and after UV-irradiation, the label was covalently bound. The labelled enzyme was digested by trypsin, the peptides were separated by ion exchange chromatography into nitreno-AMP, nitreno-ADP and nitreno-ATP labelled peptides, and these were then separated by reversed phase chromatography. Amino acid sequence analysis was used to identify the type of the nucleotide binding site. After incubation with 2-azido-[alpha-(32)P]ADP, the covalently bound label was found exclusively at beta-Tyr-362, i.e. binding occurs only to catalytic sites. Incubation conditions with 2-azido-[alpha-(32)P]ADP were varied, and conditions were found which allow selective binding of the label to different catalytic sites, either to catalytic site 2 or to catalytic site 3. For measurements of the degree of inhibition by covalent modification, CF(0)F(1) was reconstituted into phosphatidylcholine liposomes, and the membranes were energised by an acid-base transition in the presence of a K(+)/valinomycin diffusion potential. The rate of ATP synthesis was 120 s(-1), and the rate of ATP hydrolysis was 20 s(-1), both measured under multi-site conditions. Covalent modification of either catalytic site 2 or catalytic site 3 inhibited both ATP synthesis and ATP hydrolysis, the degree of inhibition being proportional to the degree of modification. Extrapolation to complete inhibition indicates that modification of one catalytic site, either site 2 or site 3, is sufficient to completely block multi-site ATP synthesis and ATP hydrolysis. The rate of ATP synthesis and the rate of ATP hydrolysis were measured as a function of the substrate concentration from multi-site to uni-site conditions with covalently modified CF(0)F(1) and with non-modified CF(0)F(1). The result was that uni-site ATP synthesis and ATP hydrolysis were not inhibited by covalent modification of either catalytic site 2 or site 3. The results indicate cooperative interactions between catalytic nucleotide binding sites during multi-site catalysis, whereas neither uni-site ATP synthesis nor uni-site ATP hydrolysis require interaction with other sites.  相似文献   

5.
1. A formula is given that describes the appearance of [14C]ATPADP outside the mitochondria after the addition of [14C] 1atp during the steady-state uncoupler-induced hydrolysis of extramitochondrial ATP. If the transported adenine nucleotides equilibrate with the intramitochondrial pool, [14C]ADP0 would be expected to appear with a lag phase that corresponds with the time needed for the radioactive labelling of the intramitochondrial adenine nucleotide pool. 2. The rates of formation of [14C]ADP outside the mitochondria after addition of [14C]ATP during the steady-state uncoupler-induced ATP hydrolysis catalysed by rat-liver mitochondria at 0 degree C were measured. 3. In the presence of carbonyl cyanide m-chlorophenylhydrazone the time course of the [14]ADPo formation was the same as that predicted on the basis of the above assumption. 4. In the presence of the less effective uncoupler, 2,4-dinitrophenol, the time course of [14C]ADPo formation was not consistent with the theoretical predictions: no lag phase was present and the measured rate was higher than the maximal calculated rate. These results can be explained by assuming a functional interaction between the adenine nucleotide translocator and the mitochondrial ATPase (F1). 5. It is concluded that under phosphorylating as well as dephosphorylating conditions, the adenine nucleotide translocator and the mitochondrial ATPase can be functionally linked to catalyse phosphorylation or dephosphorylation of extramitochondrial ADP or ATP, without participation of the intramitochondrial adenine nucleotides.  相似文献   

6.
Gating of the cystic fibrosis transmembrane conductance regulator (CFTR) involves a coordinated action of ATP on two nucleotide binding domains (NBD1 and NBD2). Previous studies using nonhydrolyzable ATP analogues and NBD mutant CFTR have suggested that nucleotide hydrolysis at NBD1 is required for opening of the channel, while hydrolysis of nucleotides at NBD2 controls channel closing. We studied ATP-dependent gating of CFTR in excised inside-out patches from stably transfected NIH3T3 cells. Single channel kinetics of CFTR gating at different [ATP] were analyzed. The closed time constant (tauc) decreased with increasing [ATP] to a minimum value of approximately 0.43 s at [ATP] >1.00 mM. The open time constant (tauo) increased with increasing [ATP] with a minimal tauo of approximately 260 ms. Kinetic analysis of K1250A-CFTR, a mutant that abolishes ATP hydrolysis at NBD2, reveals the presence of two open states. A short open state with a time constant of approximately 250 ms is dominant at low ATP concentrations (10 microM) and a much longer open state with a time constant of approximately 3 min is present at millimolar ATP. These data suggest that nucleotide binding and hydrolysis at NBD1 is coupled to channel opening and that the channel can close without nucleotide interaction with NBD2. A quantitative cyclic gating scheme with microscopic irreversibility was constructed based on the kinetic parameters derived from single-channel analysis. The estimated values of the kinetic parameters suggest that NBD1 and NBD2 are neither functionally nor biochemically equivalent.  相似文献   

7.
M F Carlier  D Pantaloni 《Biochemistry》1986,25(24):7789-7792
The sequence of reactions involved in the polymerization of ATP-actin and accompanying hydrolysis of ATP has been investigated by using a new glass-fiber filter assay. The assay allows the rapid separation of filaments from monomeric actin, and therefore the straightforward identification of the nucleotide bound to F-actin in the time course of polymerization, using double-labeled [gamma-32P,3H]ATP. The data bring a direct confirmation of the existence of the previously proposed ATP-F-actin intermediate in the time course of polymerization. Moreover, comparison of the hydrolyzed ATP (i.e., acid-labile [32P]Pi) and of 32P bound to F-actin provides direct evidence for the second intermediate ADP-Pi-F-actin in the polymerization process. This latter species is the major transient in the polymerization of ATP-actin, its lifetime being of the order of minutes.  相似文献   

8.
ATP binding to the first and second NBDs (nucleotide-binding domains) of CFTR (cystic fibrosis transmembrane conductance regulator) are bivalent-cation-independent and -dependent steps respectively [Aleksandrov, Aleksandrov, Chang and Riordan (2002) J. Biol. Chem. 277, 15419-15425]. Subsequent to the initial binding, Mg(2+) drives rapid hydrolysis at the second site, while promoting non-exchangeable trapping of the nucleotide at the first site. This occlusion at the first site of functional wild-type CFTR is somewhat similar to that which occurs when the catalytic glutamate residues in both of the hydrolytic sites of P-glycoprotein are mutated, which has been proposed to be the result of dimerization of the two NBDs and represents a transient intermediate formed during ATP hydrolysis [Tombline and Senior (2005) J. Bioenerg. Biomembr. 37, 497-500]. To test the possible relevance of this interpretation to CFTR, we have now characterized the process by which NBD1 occludes [(32)P]N(3)ATP (8-azido-ATP) and [(32)P]N(3)ADP (8-azido-ADP). Only N(3)ATP, but not N(3)ADP, can be bound initially at NBD1 in the absence of Mg(2+). Despite the lack of a requirement for Mg(2+) for ATP binding, retention of the NTP at 37 degrees C was dependent on the cation. However, at reduced temperature (4 degrees C), N(3)ATP remains locked in the binding pocket with virtually no reduction over a 1 h period, even in the absence of Mg(2+). Occlusion occurred identically in a DeltaNBD2 construct, but not in purified recombinant NBD1, indicating that the process is dependent on the influence of regions of CFTR in addition to NBD1, but not NBD2.  相似文献   

9.
P-glycoprotein (Pgp) is an ATP-dependent drug efflux pump whose overexpression confers multidrug resistance to cancer cells. Pgp exhibits a robust drug substrate-stimulable ATPase activity, and vanadate (Vi) blocks this activity effectively by trapping Pgp nucleotide in a non-covalent stable transition state conformation. In this study we compare Vi-induced [alpha-(32)P]8-azido-ADP trapping into Pgp in the presence of [alpha-(32)P]8-azido-ATP (with ATP hydrolysis) or [alpha-(32)P]8-azido-ADP (without ATP hydrolysis). Vi mimics P(i) to trap the nucleotide tenaciously in the Pgp.[alpha-(32)P]8-azido-ADP.Vi conformation in either condition. Thus, by using [alpha-(32)P]8-azido-ADP we show that the Vi-induced transition state of Pgp can be generated even in the absence of ATP hydrolysis. Furthermore, half-maximal trapping of nucleotide into Pgp in the presence of Vi occurs at similar concentrations of [alpha-(32)P]8-azido-ATP or [alpha-(32)P]8-azido-ADP. The trapped [alpha-(32)P]8-azido-ADP is almost equally distributed between the N- and the C-terminal ATP sites of Pgp in both conditions. Additionally, point mutations in the Walker B domain of either the N- (D555N) or C (D1200N)-terminal ATP sites that arrest ATP hydrolysis and Vi-induced trapping also show abrogation of [alpha-(32)P]8-azido-ADP trapping into Pgp in the absence of hydrolysis. These data suggest that both ATP sites are dependent on each other for function and that each site exhibits similar affinity for 8-azido-ATP (ATP) or 8-azido-ADP (ADP). Similarly, Pgp in the transition state conformation generated with either ADP or ATP exhibits drastically reduced affinity for the binding of analogues of drug substrate ([(125)I]iodoarylazidoprazosin) as well as nucleotide (2'(3')-O-(2,4,6-trinitrophenyl)adenosine 5'-triphosphate). Analyses of Arrhenius plots show that trapping of Pgp with [alpha-(32)P]8-azido-ADP (in the absence of hydrolysis) displays an approximately 2.5-fold higher energy of activation (152 kJ/mol) compared with that observed when the transition state intermediate is generated through hydrolysis of [alpha-(32)P]8-azido-ATP (62 kJ/mol). In aggregate, these results demonstrate that the Pgp.[alpha-(32)P]8-azido-ADP (or ADP).Vi transition state complexes generated either in the absence of or accompanying [alpha-(32)P]8-azido-ATP hydrolysis are functionally indistinguishable.  相似文献   

10.
1. A synaptic vesicle fraction isolated from bovine caudatolenticular nuclei showed enzymic activities of tyrosine hydroxylase [EC 1.14.16.2] and dopamine beta-hydroxylase [EC 1.14.17.1]. Tyrosine hydroxylase, whose subcellular localization is uncertain, appeared to be associated with the synaptic vesicles. 2. The vesicle fraction took up [3H]dopamine increasingly with time without the aid of ATP (6.3 pmol of [3H]dopamine per mg of vesicle proteins, or 3-4% of the added dopamine, at 30 min). In the presence of ATP, a transient accumulation of the amine was observed, reaching the highest level at 7-10 min (5.6 pmol dopamine/mg protein), and then a rapid release of the amine took place, obeying first-order kinetics with respect to the amine concentration. The amine uptake was strongly inhibited with NEM, regardless of the presence or absence of ATP. 3. The vesicle fraction also exhibited a weak ability to translocate protons inward and ATP-dependently, as monitored by an increase in the fluorescence intensity of ANS. The fluorescence enhancement persisted for at least 30 min and this time-dependent change was not consistent with that of the transient accumulation of [3H]dopamine mentioned above. Since the present vesicle preparation contained a small amount of mitochondrial ATPase (18-20% of the total activity), the proton translocating ability could be attributable to contaminating submitochondrial particles. 4. Therefore these results made it impossible to conclude that the transient uptake of [3H]dopamine by the synaptic vesicles was coupled to ATP hydrolysis.  相似文献   

11.
The adenine nucleotide stores of cultured adrenal medullary cells were radiolabeled by incubating the cells with 32Pi and [3H]adenosine and the turnover, subcellular distribution, and secretion of the nucleotides were examined. ATP represented 84-88% of the labeled adenine nucleotides, ADP 11-13%, and AMP 1-3%. The turnover of 32P-adenine nucleotides and 3H-nucleotides was biphasic and virtually identical; there was an initial fast phase with a t1/2 of 3.5-4.5 h and a slow phase with a half-life varying from 7 to 17 days, depending upon the particular cell preparation. The t1/2 of the slow phase for labeled adenine nucleotides was the same as that for the turnover of labeled catecholamines. The subcellular distribution of labeled adenine nucleotides provides evidence that there are at least two pools of adenine nucleotides which make up the component with the long half-life. One pool, which contains the bulk of endogenous nucleotides (75% of the total), is present within the chromaffin vesicles; the subcellular localization of the second pool has not been identified. The studies also show that [3H]ATP and [32P]ATP are distributed differently within the cell; 3 days after labeling 75% of the [32P]ATP was present in chromaffin vesicles while only 35% of the [3H]ATP was present in chromaffin vesicles. Evidence for two pools of ATP with long half-lives and for the differential distribution of [32P]ATP and [3H]ATP was also obtained from secretion studies. Stimulation of cell cultures with nicotine or scorpion venom 24 h after labeling with [3H]adenosine and 32Pi released relatively twice as much catecholamine as 32P-labeled compounds and relatively three times as much catecholamine as 3H-labeled compounds.  相似文献   

12.
Cooperative interactions between nucleotide binding sites on beef heart mitochondrial F1-ATPase have been studied by measuring substrate-promoted release of 5'adenylyl-beta,gamma-imidodiphosphate (AMP-PNP) from a single high affinity site. The site is initially loaded by incubating F1 with an equimolar amount of the nonhydrolyzable ATP analog. When unbound [3H]AMP-PNP is removed and the complex diluted to a concentration below the Kd, release of ligand shows an apparent absolute requirement for medium ADP. Release is biphasic with the extent of release during the initial rapid phase dependent on the concentration of medium ADP. Although phosphate alone has no effect, it enhances the rapid phase of ADP-promoted release over 2-fold with a half-maximal effect at 60 micrometers P1. The binding of efrapeptin (A23871) to the F1.AMP-PNP complex completely prevents ADP-promoted dissociation. Although AMP-PNP release also occurs in the presence of medium ATP, the F1.AMP-PNP complex does not dissociate if an ATP-regenerating system of sufficient capacity to prevent accumulation of medium ADP is added. Consistent with an inability of nucleoside triphosphate to promote release is the failure of medium, nonradioactive AMP-PNP to affect retention of the 3H-labeled ligand. The stability of F1.AMP-PNP complex in the absence of medium nucleotide and the highly specific ability of ADP plus P1 to promote rapid release of the ATP analog are interpreted as support for an ATP synthesis mechanism that requires substrate binding at one catalytic site for product release from an adjacent interacting site.  相似文献   

13.
Incubation of human platelets with myo-[3H]inositol in a low-glucose Tyrode's solution containing MnCl2 enhanced the labelling of phosphoinositides about sevenfold and greatly facilitated the measurement of [3H]inositol phosphates formed by the activation of phospholipase C. Labelled platelets were permeabilized by high-voltage electric discharges and equilibrated at 0 degree C with ATP, Ca2+ buffers and guanine nucleotides, before incubation in the absence or presence of thrombin. Incubation of these platelets with ATP in the presence or absence of Ca2+ ions led to the conversion of [3H]phosphatidylinositol to [3H]phosphatidylinositol 4-phosphate and [3H]phosphatidylinositol 4,5-bisphosphate ([3H]PtdInsP2). At a pCa of 6, addition of 100 microM GTP[gamma S] both prevented this accumulation of [3H]PtdInsP2 and stimulated its breakdown; the formation of [3H]inositol phosphates was increased ninefold. After 5 min these comprised 70% [3H]inositol monophosphate ([3H]InsP), 28% [3H]inositol bisphosphate ([3H]InsP2) and 2% [3H]inositol trisphosphate ([3H]InsP3). In shorter incubations higher percentages of [3H]InsP2 and [3H]InsP3 were found. In the absence of added Ca2+, the formation of [3H]inositol phosphates was decreased by over 90%. Incubation of permeabilized platelets with GTP[gamma S] in the presence of 10 mM Li+ decreased the accumulation of [3H]InsP and increased that of [3H]InsP2, without affecting [3H]InsP3 levels. Addition of unlabelled InsP3 decreased the intracellular hydrolysis of exogenous [32P]InsP3 but did not trap additional [3H]InsP3. These results and the time course of [3H]inositol phosphate formation suggest that GTP[gamma S] stimulated the action of phospholipase C on a pool of [3H]phosphatidylinositol 4-phosphate that was otherwise converted to [3H]PtdInsP2 and that much less hydrolysis of [3H]phosphatidylinositol to [3H]InsP or of [3H]PtdInsP2 to [3H]InsP3 occurred. At a pCa of 6, addition of thrombin (2 units/ml) to permeabilized platelets caused small increases in the formation of [3H]InsP and [3H]InsP2. This action of thrombin was enhanced twofold by 10-100 microM GTP and much more potently by 4-40 microM GTP[gamma S]. In the presence of the latter, thrombin also increased [3H]InsP3. The total formation of [3H]inositol phosphates by permeabilized platelets incubated with thrombin and GTP[gamma S] was comparable with that observed on addition of thrombin alone to intact platelets. However, HPLC of the [3H]inositol phosphates formed indicated that about 75% of the [3H]InsP accumulating in permeabilized platelets was the 4-phosphate, whereas in intact platelets stimulated by thrombin, up to 80% was the 1-phosphate.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
The guanine nucleotides guanosine 5'[beta, gamma-imido]triphosphate (Gpp[NH]p), guanosine 5'-[gamma-thio]-triphosphate (GTP gamma S), GMP, GDP and GTP stimulated the hydrolysis of inositol phospholipids by a phosphodiesterase in rat cerebral cortical membranes. Addition of 100 microM-Gpp[NH]p to prelabelled membranes caused a rapid accumulation of [3H )inositol phosphates (less than 30 s) for up to 2 min. GTP gamma S and Gpp [NH]p caused a concentration-dependent stimulation of phosphoinositide phosphodiesterase with a maximal stimulation of 2.5-3-fold over control at concentrations of 100 microM. GMP was as effective as the nonhydrolysable analogues, but much less potent (EC50 380 microM). GTP and GDP caused a 50% stimulation of the phospholipase C at 100 microM and at higher concentrations were inhibitory. The adenine nucleotides App[NH]p and ATP also caused small stimulatory effects (64% and 29%). The guanine nucleotide stimulation of inositide hydrolysis in cortical membranes was selective for inositol phospholipids over choline-containing phospholipids. Gpp[NH]p stimulated the production of inositol trisphosphate and inositol bisphosphate as well as inositol monophosphate, indicating that phosphoinositides are substrates for the phosphodiesterase. EGTA (33 microM) did not prevent the guanine nucleotide stimulation of inositide hydrolysis. Calcium addition by itself caused inositide phosphodiesterase activation from 3 to 100 microM which was additive with the Gpp[NH]p stimulation. These data suggest that guanine nucleotides may play a regulatory role in the modulation of the activity of phosphoinositide phosphodiesterase in rat cortical membranes.  相似文献   

15.
Dephosphorylation of [32P]phosphoenzyme of bovine brain Na+,K+-stimulated ATP phosphohydrolase (EC 3.6.1.3), labelled by [gamma-32P]ATP, was investigated at 21 degrees C by means of a rapid-mixing technique. On addition of a high concentration of KCl (10 mM) to [32P]phosphoenzyme at steady state in the presence of Mg2+ and Na+, very rapid dephosphorylation was obtained. Simultaneously, the amount of [32P]orthophosphate increased at about the same rate. It was concluded that this K+-stimulated dephosphorylation and liberation of [32P]orthophosphate from the [32P]phosphoenzyme was rapid enough to participate in the Na+,K+-stimulated hydrolysis of ATP. In order to study the dephosphorylation in absence of continuing 32P-labelling, excess unlabelled ATP or a chelator of Mg2+ was added. Simultaneous addition of a high concentration of KCl to the [32P]phosphoenzyme formed in the presence of Mg2+ and Na+ but in the absence of K+, resulted in an initial very rapid phase and a subsequent slower phase of dephosphorylation. With KCl also initially present in the incubation medium, only the slow phase was observed. The slow phase of dephosphorylation also seemed to be sufficiently rapid to participate in the Na+, K+-stimulated ATPase reaction. On addition of a high concentration of ADP (5 mM) to [32P]phosphoenzyme formed in the presence of Mg2+ and Na+, an initial comparatively rapid, and later slow phase of dephosphorylation were detected. This gave further support for different forms of phosphoenzyme. Approximate concentrations of these forms, in the absence and presence of KCl, were estimated by extrapolation and the turnover of these forms was calculated. The nature of the kinetically different components of phosphoenzyme and their role in the Na+, K+-stimulated ATPase reaction is discussed.  相似文献   

16.
Under conditions of molar excess of enzyme, isolated F1-ATPase from beef heart mitochondria catalyses ATP hydrolysis biphasically. The rate constants for product release are approximately 10(-1) and 10(-4)-10(-3) s-1, respectively. The slow phase of ATP hydrolysis is insensitive to EDTA. [gamma-32P]ATP splitting in the slow phase cannot be chased from the enzyme during several catalytic turnovers. It follows from these results that the slow single-site hydrolysis of ATP (kcat approximately 10(-4) s-1), initially observed by Grubmeyer et al. [(1982) J. Biol. Chem. 257, 12092-12100], is not carried out by the normal catalytic site. In contrast, the phase of rapid ATP hydrolysis (kcat approximately 10(-1) s-1) is completely prevented by EDTA and is believed to be the normal function of the normal catalytic site of F1-ATPase.  相似文献   

17.
A soluble porcine H,K-ATPase preparation was obtained with the nonionic detergent, C12E8. ATP hydrolysis by the soluble H,K-ATPase was stimulated with respect to the native preparation at pH 6.1, while the K(+)-phosphatase activity was comparable to the native enzyme. The soluble enzyme demonstrated characteristic ligand-dependent effects on ATP hydrolysis, including ATP activation of K(+)-stimulated hydrolysis with a K0.5 of 28 +/- 4 microM ATP, and inhibition with an IC50 of 2.1 mM ATP. The activation and inhibition of ATP hydrolysis by K+ was also observed with a K0.5 for activation of 2.8 +/- 0.4 mM KCl at 2.0 mM ATP (pH 6.1) and inhibition with an IC50 of 135 mM KCl at 0.05 mM ATP. 2-Methyl-8-(phenylmethoxy)imidazo[1,2a]pyridine-3-acetonitrile (SCH 28080), a specific inhibitor of the native H,K-ATPase, competitively inhibited the K(+)-stimulated activity with a Ki of 0.035 microM. The soluble enzyme was stable with a t0.5 for ATPase activity of 6 h between 4 and 11 degrees C. The demonstration of these related ligand responses in the catalytic reactions of the soluble preparation indicates that it is an appropriate medium for investigation of the subunit associations of the functional H,K-ATPase. Subunit associations of the active soluble enzyme were assessed following treatment with the crosslinking reagent, glutaraldehyde. The distribution of crosslinked particles was independent of the soluble protein concentration in the crosslinking buffer within the protein range 0.3 to 2.0 mg/ml or the detergent to protein ratio varied from 1 to 15 (w/w). The crosslinked pattern was unaffected by the presence or absence of K during crosslinking or nucleotide concentration. These observations suggest that crosslinking occurs in associated subunits that do not undergo rapid associations dependent upon enzyme turnover. Phosphorylation of the soluble enzyme with 0.1 mM MgATP produced a phosphoprotein at 94 kDa. A phosphoprotein obtained after glutaraldehyde treatment exhibited identical electrophoretic mobility to the crosslinked particle identified by silver stain. Glutaraldehyde treatment of soluble protein fractions resolved on a linear 10-35% glycerol gradient revealed several smaller peptides partially resolved from the crosslinked pump particle, but no active fraction enriched in the monomeric H,K-ATPase. This data indicates that the functional porcine gastric H,K-ATPase is organized as a structural dimer.  相似文献   

18.
The mechanism of accumulation of radioactive label from fNle-Leu-[3H]Phe by guinea pig alveolar macrophages was investigated. The binding of fNle-Leu-[3H]Phe to macrophages reached equilibrium within 5 min at 4 degrees C, but equilibrium could not be achieved at temperatures where fNle-Leu-Phe stimulated superoxide anion production is observed (e.g., 21-23 degrees C). At this temperature a rapid phase of initial binding of fNle-Leu-[3H]Phe to its receptor was followed by continued accumulation of cell-associated radioactivity which was linear and was dependent on the extracellular pH, i.e., the rate increased as the pH was lowered from pH 8 to pH 6. Examination for possible intracellular hydrolysis of fNle-Leu-[3H]Phe revealed the presence of extensive amounts of [3H]phenylalanine, both cell-associated and in the medium. The increases in cell-associated [3H]phenylalanine correlated in time and pH with cell-associated radioactivity that was accumulated after stimulation with fNle-Leu-[3H]Phe. The addition of 1 mM unlabelled phenylalanine blocked the long term accumulation of label from fNle-Leu-[3H]Phe by macrophages. 1 mM phenylalanine had no measureable effect on fNle-Leu-Phe stimulated O2- production, fNle-Leu-[3H]Phe hydrolysis or on fNle-Leu-[3H]Phe binding to its receptor. These results indicated that the long term accumulation of radioactivity by alveolar macrophages was due to extracellular hydrolysis of fNle-Leu-[3H]Phe followed by transport of liberated [3H]phenylalanine into the cells. A high affinity (Km = 3.56 X 10(-8) M) transport system for phenylalanine was measured in alveolar macrophages, which was not stimulated by the addition of fNle-Leu-Phe. The extracellular hydrolysis of fNle-Leu-[3H]Phe could not be attributed to release of macrophage enzymes into the medium. The responsible proteinase appears to be membrane bound and has a Km for the hydrolysis of fNle-Leu-[3H]Phe of 2.6 X 10(-7) M which is similar to the Kd (1.5 X 10(-7) M) for fNle-Leu-Phe binding. Taken together, these data suggest that for the alveolar macrophage: (1) formyl peptides are not internalized by a receptor-mediated process; (2) a surface proteinase can catalyze the hydrolysis of formyl peptides; and (3) [3H]phenylalanine formed by fNle-Leu-[3H]Phe hydrolysis is transported into the interior of the macrophage.  相似文献   

19.
(1) Incubation of the beef heart mitochondrial ATPase, F1 with Mg-ATP was required for the binding of the natural inhibitor, IF1, to F1 to form the inactive F1-IF1 complex. When F1 was incubated in the presence of [14C]ATP and MgCl2, about 2 mol 14C-labeled adenine nucleotides were found to bind per mol of F1; the bound 14C-labeled nucleotides consisted of [14C]ADP arising from [14C]ATP hydrolysis and [14C]ATP. The 14C- labeled nucleotide binding was not prevented by IF1. These data are in agreement with the idea that the formation of the F1-IF1 complex requires an appropriate conformation of F1. (2) The 14C-labeled adenine nucleotides bound to F1 following preincubation of F1 with Mg-[14C] ATP could be exchanged with added [3H]ADP or [3H]ATP. No exchange occurred between added [3H]ADP or [3H]ATP and the 14 C-labeled adenine nucleotides bound to the F1-IF1 complex. These data suggest that the conformation of F1 in the isolated F1-IF1 complex is further modified in such a way that the bound 14C-labeled nucleotides are no longer available for exchange. (3) 32Pi was able to bind to isolated F1 with a stoichiometry of about 1 mol of Pi per mol of F1 (Penefsky, H.S. (1977) J. Biol. Chem. 252, 2891-2899). There was no binding of 32Pi to the F1-IF1 complex. Thus, not only the nucleotides sites, but also the Pi site, are masked from interaction with external ligands in the isolated F1-IF1 complex.  相似文献   

20.
We have investigated effects of pH on the catalytic and allosteric properties of the cGMP-stimulated cyclic nucleotide phosphodiesterase purified from calf liver. In the "activated" state, i.e., with 0.5 microM [3H]cAMP plus 1 microM cGMP or at saturating substrate concentrations (250 microM [3H]cAMP or [3H]cGMP), hydrolysis was maximal at pH 7.5-8.0 in assays of different pH. Hydrolysis of concentrations of substrate not sufficient to saturate regulatory sites and below the apparent Michaelis constant (Kmapp), i.e., 0.5 microM [3H]cAMP or 0.01 microM [3H]cGMP, was maximal at pH 9.5. Although hydrolysis of 0.5 microM [3H]cAMP increased with pH from 7.5 to 9.5, cGMP stimulation of cAMP hydrolysis decreased. As pH increased or decreased from 7.5, Hill coefficients (napp) and Vmax for cAMP decreased. Thus, assay pH affects both catalytic (Vmax) and allosteric (napp) properties. Enzyme was therefore incubated for 5 min at 30 degrees C in the presence of MgCl2 at various pHs before assay at pH 7.5. Prior exposure to different pHs from pH 6.5 to 10.0 did not alter the Vmax or cGMP-stimulated activity (assayed at pH 7.5). Incubation at high (9.0-10.0) pH did, in assays at pH 7.5, markedly increase hydrolysis of 0.5 microM [3H]cAMP and reduce Kmapp and napp. After incubation at pH 10, hydrolysis of 0.5 microM [3H]cAMP was maximally increased and was similar in the presence or absence of cGMP. Thus, after incubation at high pH, the phosphodiesterase acquires characteristics of the cGMP-stimulated form. Activation at high pH occurs at 30 degrees C but not 5 degrees C, requires MgCl2, and is prevented but not reversed by ethylenediaminetetraacetic acid.(ABSTRACT TRUNCATED AT 250 WORDS)  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