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 546 毫秒
1.
The active form of Chlorella fusca nitrate reductase can be reversibly converted into its inactive form by reduction with NADH in the presence of ADP. Under the experimental conditions used, no inactivation occurs when nitrate is simultaneously present or when the nucleotides act isolately, the inactivating effect being maximal at a concentration of ADP (0.3 mM) equimolecular with that of NADH. The inactive enzyme thus attained can be completely reactivated by reoxidation with ferricyanide. The redox state of the pyridine nucleotide and the phosphorylation degree of the adenine nucleotide are critical for the inactivation process to ensue, since neither NAD+ nor AMP or ATP do exert any effect. ADP is also a powerful, although rather unspecific, protector against thermal inactivation of the NADH-diaphorase moiety of the NADH-nitrate reductase complex.  相似文献   

2.
Various modulating influences of negative and positive membrane charges on binding and transport properties of the reconstituted ADP/ATP carrier from mitochondria were investigated. The results are interpreted in terms of functional and structural asymmetries of the adenine nucleotide carrier embedded in the liposomal membrane. The surface potential of liposomes was measured directly either by potential-dependent adsorption of the fluorescent dye 2-p-toluidinylnaphthalene 6-sulfonate (TNS) or by the pK shift of the lipophilic pH indicator pentadecylumbelliferone. These results were correlated with the following observations. (1) Negative surface potentials increase the apparent dissociation constant, Kd, for binding of the negatively charged inhbitor carboxyatractylate to the reconstituted carrier protein. (2) Surface potentials modulate the apparent transport affinity, Km, of the reconstituted adenine nucleotide carrier for ADP and ATP. The interaction of surface charges with the transport function was investigated with carrier proteins oriented both right-side-out and inside-out. Thus the influence of the surface potential on the function of the ADP/ATP carrier could be determined for the internal and external active sites of the translocator on the outer side of the membrane. Large discrepancies were observed not only between the potentials measured directly (fluorescent dyes) and those measured indirectly (binding and transport affinities), but also between the different surface potentials determined from the influence on the alternatively oriented carrier proteins. The effect of surface charges was rather weak on the cytosolic side of the translocator, whereas there was a strong influence of surface charges on the active site at the matrix side. The most obvious explanation, i.e., screening of negative membrane charges by positively charged amino acid residues at the protein surface, could be ruled out. Besides the modulation of binding affinities for substrates and inhibitors, an additional side-specific effect of surface charges on the transport velocity was observed. Again, the influence on the internal active site of the ADP/ATP carrier was found to be much higher than that on the cytosolic site. The observed effects can be explained by a definite structural asymmetry of the carrier embedded in the liposomal membrane. That site which is physiologically exposed to the cytosol is located at a considerable distance from the plane of the membrane, whereas the opposite site seems to be in close proximity to the membrane surface. Moreover, a spatial equivalence of carboxyatractylate binding site and nucleotide binding site at the external side of the carrier protein was concluded.  相似文献   

3.
The 2′,3′-dialdehyde of ADP, obtained by periodate oxidation of ADP, inhibited the hydrolytic activity of the purified Ca2+, Mg2+-activated ATPase of Escherichiacoli. In the initial stages of the reaction inhibition was due to the reaction of 1 mol inhibitor/active site. When non-specific labelling of amino groups by the dialdehyde was lowered by the simultaneous presence of 15 mM ATP in the reaction mixture, 3 mol “ATP-protectable” binding sites/mol ATPase were found. “ATP-protectable” binding of the dialdehyde was not observed when the hydrolytically inactive ATPase of an unc A mutant of E.coli was used although binding of the inhibitor to non-protected amino groups still occurred. This suggests that the mutant ATPase is unable to bind ATP or that the amino groups with which the dialdehyde reacts in the native enzyme are absent or masked.  相似文献   

4.
Abstract

Bovine liver adenosine kinase displays a characteristic intrinsic fluorescence due to 3 tryptophans/mol. This fluorescence is very sensitive to ligand binding and was used to characterize the ligand binding sites of the enzyme. ADP or ATP showed a monophasic saturation curve consistent with the existence of one binding site. In contrast, adenosine and AMP gave biphasic saturation curves suggesting the existence of at least two binding sites with high and low affinity. These binding sites were further characterized by studying the complexation of adenosine kinase with O-(N-methylanthraniloyl)adenosine nucleoside or nucleotide analogues.  相似文献   

5.
The binding of the fluorescent analog of adenosine diphosphate (ADP)1, 1,N6-ethenoadenosine diphosphate (εADP) to myosin and its subfragments, heavy meromyosin (HMM) and subfragment one (S1), has been studied under analagous conditions to those previously used in comparable studies on the binding of ADP to these molecules. The results indicate that there are two binding sites for εADP on myosin and HMM, and one site on S1. The dissociation constants for all had an identical value, within experimental error, of 2.0 (± .5) × 10?5 M?1. This is identical to the values found by Young (J. Biol. Chem., 242, 2790 (1967)) for ADP. In addition, the kinetics of hydrolysis of εATP versus ATP by S1 were studied. Values of Vmax and Km were 25 μM phosphate sec?1 (gm protein)?1 and 5 × 10?5 M?1 for ATP, and 80 μN phosphate sec?1 (gm protein)?1 and 45 × 10?5 M?1 for εATP. The results indicate that the increased Vmax that occurs when εATP is used as a substitute for ATP is not due to either an increased binding affinity of ATP for myosin and its subfragments, nor due to a decreased binding affinity of εATP versus ADP. This in turn suggests that the increase in Vmax may be due to an increased hydrolytic rate of εATP vs ATP in the enzyme substrate complex.  相似文献   

6.
The ATP/ADP exchange is shown to be a partial reaction of the (H+ + K+)-ATPase by the absence of measurable nucleoside diphosphokinase activity and the insensitivity of the reaction to P1, P5 -di(adenosine-5′) pentaphosphate, a myokinase inhibitor. The exchange demonstrates an absolute requirement for Mg2+ and is optimal at an ADP/ATP ratio of 2. The high ATP concentration (K0.5 = 116 μM) required for maximal exchange is interpreted as evidence for the involvement of a low affinity form of nucleotide site. The ATP/ADP exchange is regarded as evidence for an ADP-sensitive form of the phosphoenzyme. In native enzyme, pre-steady state kinetics show that the formation of the phosphoenzyme is partially sensitive to ADP while modification of the enzyme by pretreatment with 5,5′-dithiobis(2-nitrobenzoic acid) (DTNB) in the absence of Mg2+ results in a steady-state phosphoenzyme population, a component of which is ADP sensitive. The ATP/ADP exchange reaction can be either stimulated or inhibited by the presence of K+ as a function of pH and Mg2+.  相似文献   

7.
We have investigated the suitability of 5′-p-fluorosulfonylbenzoyladenosine (FSBA) as an ATP site affinity probe for the canine kidney Na+,K+-ATPase. The purified enzyme is slowly inactivated by this compound in suitable buffers, losing about half of its activity over a two-hour period. The rate of inactivation is more rapid in 0.1 M KCl than in 0.1 M NaCl. Low concentrations of ATP protect the enzyme against inactivation, with half-maximal effects at 4 μM ATP in 0.1 M NaCl and 350 μM ATP in 0.1 M KCl. ADP also protects against FSBA inhibition, but AMP is ineffective when present at 100 μM levels. This pattern is consistent with the previously described nucleotide specificity of the Na+,K+-ATPase. Addition of protective amounts of ATP after inactivation has occurred does not restore enzyme activity, indicating that inhibition is irreversible. Measurement of the concentration-dependence of FSBA inactivation suggests an apparent Kd for binding of this compound well above 1 mM, the solubility limit of the analog. This finding is reinforced by the failure of 1 mM FSBA to compete effectively with ATP for the high-affinity ATP site of the enzyme. Nevertheless, attachment of the analog to this site is indicated by its ability to prevent [3H]-ADP binding in proportion to the number of sites it has inactivated. Studies with [3H]-FSBA show that about 1 mole of the analog attaches specifically to the α subunit per mole of enzyme inactivated. A similar amount of nonspecific labeling also occurs with negligible effect on enzyme activity. These findings suggest that FSBA may be useful in probing the topography of the high-affinity ATP binding site of the Na+,K+-ATPase and related enzymes.  相似文献   

8.
(1) Eosin bound to the (Na+ + K+)-ATPase in the presence of K+ has practically the same fluorescence as eosin without enzyme while in the presence of Na+ the fluorescence is higher, the excitation maximum is shifted from 518 to 524 nm, the emission maximum from 538 to 542 nm, and a shoulder appears at about 490 nm on the excitation curve. (2) The amount of eosin bound increases with the K+ concentration but with a low affinity. With equal concentrations of Na+ and K+ more is bound in the presence of Na+, and the difference between 150 mM Na+ and 150 mM K+ shows one high-affinity eosin binding site per 32P-labelling site (KD 0.45 μM). With lower concentrations of the cations there are between one and two Na+-dependent high-affinity eosin binding sites per 32P-labelling site. (3) ATP (and ADP) prevents the hig-affinity Na+-dependent eosin binding and there is competition between eosin and ATP for the hydrolysis in the presence of Na+ (+Mg2+). (4) Eosin, like ATP, increases the Na+ relative to K+ affinity (Na+ + K+ = 150 mM) for Na+ activation of hydrolysis and for Na+ protection against inactivation by N-ethylmaleimide. (5) The results suggest that the high affinity eosin binding site is an ATP binding site and that it is located on the enzyme in an environment with a low polarity, i.e., the conformational change induced by Na+ opens a high-affinity site for ATP while K+ closes the site (or decreases the affinity to a low level). The experiments suggest, furthermore, that the ATP which increases the Na+ relative to K+ affinity of the internal sites is not the ATP which is hydrolyzed, i.e., in a turnover cycle in the presence of Na+ + K+ the system reacts with two different ATP molecules.  相似文献   

9.
Three types of assays were used to characterize adenine nucleotide binding sites on the Ca2+, Mg2+-activated ATPase of normal Escherichia coli and its unc A 401 and unc D 412 mutants. ADP was bound mainly at a single site in normal and mutant ATPase. In the absence of divalent cations ATP was bound at a single high-affinity and three low-affinity sites in normal and unc D ATPases. The 2′,3′-dialdehyde (oADP) obtained by periodate oxidation of ADP reacted with both low- and high-affinity sites whereas oATP was bound primarily at a low-affinity site. Two types of adenine nucleotide binding sites, a high-affinity site reacting with ATP and ADP and a low-affinity site for ATP, were detected by the effects of these nucleotides on the fluorescence of the aurovertin D-ATPase complex. This high-affinity site(s) was present in normal and mutant ATPases. However, the fluorescence response at both high- and low-affinity sites was modified in the unc D ATPase as a consequence of the abnormal β subunit in this enzyme. Normal fluorescence responses were not induced by the binding of oADP or oATP to the ATPases. ATP was bound at a single site on isolated α subunits of the enzyme. Since this site was not detected in the unc A ATPase, it is unlikely to be the high-affinity site detected in the intact enzyme or the binding site for the endogenous tightly bound adenine nucleotides found in the purified ATPase. It is more probable that the site detected on the isolated α subunit from the normal enzyme is that which binds oADP since this site was absent in the unc A ATPase. Pretreatment of the normal ATPase with either N, N′-dicyclohexyl-carbodiimide (DCCD) or with 4-chloro-7-nitrobenzofurazan (NbfCl), reagents which inhibit ATPase activity by reacting with a β subunit, affected binding of oADP to α subunit(s) but had less effect with oATP. Inhibition of oADP binding could be due to conformational changes induced in the α subunit by the reaction of DCCD and NbfCl with a β subunit, or to steric reasons. If the latter hypothesis is correct, the active site of the ATPase would be at the interface between α and β subunits of the enzyme.  相似文献   

10.
D. Bar-Zvi  N. Shavit 《BBA》1982,681(3):451-458
Inactivation of the chloroplast ATPase upon tight nucleotide binding was studied with several adenine nucleotide analogs. Compared with ADP, the other nucleoside diphosphates were less effective in the follwing order: IDP >?-ADP > 1-oxido-ADP > GDP. The nucleotide analogs compete with ADP for binding to the tight nucleotide-binding site(s) on the ATPase and also prevent further inactivation by ADP. AdoPP[NH]P also causes inactivation but has a lower affinity than ADP. [3H]GDP binds tightly to the ATPase, but the resulting enzyme-GDP complex is more readily dissociable than the enzyme-ADP complex. Although both nucleotides appear to bind to the same site, the catalytic and binding properties of the coresponding nucletide-enzyme complexes differ. Binding of GDP also decreases the rate and extent of the sontaneous decay of the activated enzyme. PPi decreases the rate of inacivation caused by ADP and also the level of tigthly buond ADP. Based on these results, we suggest that two different confomations of the ATPase exist which contain tigthly bound ADP. The active conformation is conveted to the inactive conformation in the absence of a continued supply of energy by illumination or ATP hydrolysis.  相似文献   

11.
H+-FOF1-ATP synthase (F-ATPase, F-type ATPase, FOF1 complex) catalyzes ATP synthesis from ADP and inorganic phosphate in eubacteria, mitochondria, chloroplasts, and some archaea. ATP synthesis is powered by the transmembrane proton transport driven by the proton motive force (PMF) generated by the respiratory or photosynthetic electron transport chains. When the PMF is decreased or absent, ATP synthase catalyzes the reverse reaction, working as an ATP-dependent proton pump. The ATPase activity of the enzyme is regulated by several mechanisms, of which the most conserved is the non-competitive inhibition by the MgADP complex (ADP-inhibition). When ADP binds to the catalytic site without phosphate, the enzyme may undergo conformational changes that lock bound ADP, resulting in enzyme inactivation. PMF can induce release of inhibitory ADP and reactivate ATP synthase; the threshold PMF value required for enzyme reactivation might exceed the PMF for ATP synthesis. Moreover, membrane energization increases the catalytic site affinity to phosphate, thereby reducing the probability of ADP binding without phosphate and preventing enzyme transition to the ADP-inhibited state. Besides phosphate, oxyanions (e.g., sulfite and bicarbonate), alcohols, lauryldimethylamine oxide, and a number of other detergents can weaken ADP-inhibition and increase ATPase activity of the enzyme. In this paper, we review the data on ADP-inhibition of ATP synthases from different organisms and discuss the in vivo role of this phenomenon and its relationship with other regulatory mechanisms, such as ATPase activity inhibition by subunit ε and nucleotide binding in the noncatalytic sites of the enzyme. It should be noted that in Escherichia coli enzyme, ADP-inhibition is relatively weak and rather enhanced than prevented by phosphate.  相似文献   

12.
13.
The 13S coupling factor of oxidative phosphorylation from Alcaligenes faecalis forms an unusually stable complex with ADP which can be isolated by simple gel filtration. Most preparations of enzyme exhibit an apparent binding ratio of 1 mol of ADP per mol of enzyme with a dissociation constant of approximately 15 μm. One mol of adenylyl imidodiphosphate (AMP-PNP) also binds, with a dissociation constant of about 3 μm. A constant could not be obtained from ATP binding studies because this nucleotide is hydrolyzed by the enzyme. Competition studies suggest that both ADP and AMP-PNP bind to the same site. Bound nucleotides are in a very slow equilibrium with free nucleotides, with a turnover time of 1–2 h. The rate of radionucleotide dissociation from the isolated enzyme-nucleotide complex increases when unlabeled nucleotide is added, suggesting that binding of nucleotide to one site on the enzyme allosterically promotes dissociation of nucleotide from another site. A nucleotide-induced “flip-flop” type of oscillation of the properties of the nucleotide binding sites on the coupling factor is proposed. From a comparison of the kinetic parameters of the intrinsic adenosinetriphosphatase activity and the nucleotide binding parameters of the enzyme population in toto, it is suggested that the enzyme exhibits functional polymorphism.  相似文献   

14.
Cold labile extramitochondrial acetyl-CoA hydrolase (dimeric form) purified from rat liver was activated by various nucleoside triphosphates and inhibited by various nucleoside diphosphates. Activation of acetyl-CoA hydrolase by ATP was inhibited by a low concentration of ADP (Ki congruent to 6.8 microM) or a high concentration of AMP (Ki congruent to 2.3 mM). ADP and AMP were competitive inhibitors of ATP. A Scatchard plot of the binding of ATP to acetyl-CoA hydrolase (dimer) at room temperature gave a value of 25 microM for the dissociation constant with at least 2 binding sites/mol of dimer. Cold-treated monomeric enzyme also associated with ATP-agarose, suggesting that the monomeric form of the enzyme also has a nucleotide binding site(s), probably at least 1 binding site/mol of monomer. Phenylglyoxal or 2,3-butanedione, both of which modify arginyl residues of protein, inactivated acetyl-CoA hydrolase. ATP (an activator) greatly protected acetyl-CoA hydrolase from inactivation by these reagents, while ADP (an inhibitor) greatly (a substratelike, competitive inhibitor), and CoASH (a product) were less effective. However, addition of ADP plus valeryl-CoA (or CoASH) effectively prevented the inactivation by 2,3-butanedione, but that is not the case for phenylglyoxal. These results suggest that one or more arginyl residues are involved in the nucleotide binding site of extramitochondrial acetyl-CoA hydrolase and that their nucleotide binding sites locate near the substrate binding site.  相似文献   

15.
Luit Slooten  Adriaan Nuyten 《BBA》1984,766(1):88-97
(1) Rates of ATP synthesis and ADP-arsenate synthesis catalyzed by Rhodospirillum rubrum chromatophores were determined with the firefly luciferase method and by a coupled enzyme assay involving hexokinase and glucose-6-phosphate dehydrogenase. (2) Vm for ADP-arsenate synthesis was about 2-times lower than Vm for ATP-synthesis. With saturating [ADP], K(Asi) was about 20% higher than K(Pi). With saturating [anion], K(ADP) was during arsenylation about 20% lower than during phosphorylation. (3) Plots of 1v vs. 1[substrate] were non-linear at low concentrations of the fixed substrate. The non-linearity was such as to suggest a positive cooperativity between sites binding the variable substrate, resulting in an increased VmKm ratio. High concentrations of the fixed substrate cause a similar increase in VmKm, but abolish the cooperativity of the sites binding the variable substrate. (4) Low concentrations of inorganic arsenate (Asi) stimulate ATP synthesis supported by low concentrations of Pi and ADP about 2-fold. (5) At high ADP concentrations, the apparent Ki of Asi for inhibition of ATP-synthesis was 2–3-times higher than the apparent Km of Asi for arsenylation; the apparent Ki of Pi for inhibition of ADP-arsenate synthesis was about 40% lower than the apparent Km of Pi for ATP synthesis. (6) The results are discussed in terms of a model in which Pi and Asi compete for binding to a catalytic as well as an allosteric site. The interaction between these sites is modulated by the ADP concentration. At high ADP concentrations, interaction between these sites occurs only when they are occupied with different species of anion.  相似文献   

16.
17.
The adenine nucleotide translocase, the transport protein for ADP and ATP, located in the inner mitochondrial membrane is an important site for the regulation of cell metabolism. Inhibition of the adenine nucleotide translocase by long chain fatty acyl CoA esters demonstrated invitro may also occur invivo when the complete oxidation of fatty acids by the myocardium has been compromised during ischemia. Reversal of this biochemical lesion may be of benefit in the preservation of the ischemic myocardium.  相似文献   

18.
K+ induces an apparent heterogeneity among an otherwise homogeneous population of nucleotide-binding sites in (Na+ + K+)-ATPase preparations from pig kidney. With the help of ouabain we show that this heterogeneity cannot be due to a mixture of different and independent sites and conclude that each enzyme molecule must contain two nucleotide site-containing units that show interaction. Na+ induces an apparent heterogeneity among an otherwise homogeneous population of ouabain-binding sites. The argument is, therefore, extended to include one ouabain site on each of the structural units that bind nucleotide. All these structural units are shown to hydrolyse substrate at identical rates. Using the presently available molecular weight data, it is concluded that the enzyme is composed of two subunits each possessing one nucleotide-binding site, one ouabain-binding site, one α-peptide and the capacity for hydrolysing ATP and p-nitrophenyl phosphate.  相似文献   

19.
The total adenine nucleotide content of rat liver mitochondria was varied in vitro over a wide range in order to investigate a possible relationship between net changes in the total matrix ATP + ADP + AMP content and the overall rate of citrulline synthesis. Isolated mitochondria were specifically depleted of matrix adenine nucleotides by incubating with inorganic pyrophosphate (G. K. Asimakis and J. R. Aprille, 1980, Arch. Biochem. Biophys.203, 307–316); alternatively, matrix adenine nucleotides were increased by incubating mitochondria with 1 mm ATP at 30 °C. No exogenous ATP or ADP was included in the subsequent incubations for the determination of citrulline synthesis. Rates varied from 0.1 to 1.6 μmol citrulline/mg protein/h as a linear function of total adenine nucleotide content in the range 2–15 nmol (ATP + ADP + AMP)/mg protein. Further increases in the matrix ATP + ADP + AMP content caused no further increase in citrulline synthesis rates. Changes in the total adenine nucleotide content were reflected in proportional changes in both the ATP and ADP content of the matrix. The ATPADP ratio did not change significantly. Therefore, the variations in citrulline synthesis were most simply explained as the effect of different concentrations of ATP on the activity of carbamoyl-phosphate synthetase. It was concluded that net changes in the total adenine nucleotide content can contribute to the control of citrulline synthesis. These findings are significant in the context of recent evidence which shows that the matrix adenine nucleotide pool size is under hormonal control.  相似文献   

20.
D. discoideum contains kinetically distinguishable cell surface cAMP binding sites. One class, S, is slowly dissociating and has high affinity for cAMP (Kd = 15 nM, t12 = 15 s). A second class is fast dissociating (t12 about 1 s) and is composed of high affinity binding sites H (Kd ≈ 60 nM), and low affinity binding sites L (Kd = ≈ 450 nM) which interconvert during the binding reaction. Guanine nucleotides affect these three binding types in membranes prepared by shearing D.discoideum cells through Nucleopore filters. The affinity of S for cAMP is reduced by guanine nucleotides from 13 nM to 25 nM, and the number of S-sites is reduced about 50%. The number of fast dissociating sites is not altered by guanine nucleotides, but these sites are mainly in the low affinity state. Half-maximal effects are obtained at about 1 μM GTP, 2 μM GDP and 10 μM Gpp(NH)p(guanyl-5′-yl-imidodiphosphate); ATP and ADP are without effect up to 1 mM. These results indicate that D.discoideum cells have a functionally active guanine nucleotide binding protein involved in the transduction of extracellular cAMP signals via cell surface cAMP receptors.  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