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.
The interactions of nucleotides at the allosteric and catalytic sites of phosphorylase kinase were examined. Binding of nucleoside triphosphates at the nucleoside diphosphate allosteric activation site inhibited enzymatic activity; this was observed with either ATP or GTP. Increasing concentrations of ADP caused a biphasic response: low concentrations activated and higher concentrations inhibited. Inhibition was due to the binding of ADP at the catalytic site, as opposed to an allosteric inhibitory site. GDP activated at low concentrations, but did not inhibit even at relatively high concentrations, and is therefore a specific probe for the allosteric site. Maximal activity of the nonactivated holoenzyme at pH 6.8 is achieved at an optimal ratio of ATP to ADP, such that the inhibitory actions of ATP at the allosteric site and of ADP at the catalytic site are balanced. Various potential molecular mechanisms to explain the allosteric activation by ADP were examined and ruled out, thus strengthening our previous conclusion that the activation is predominantly caused by a conformational transition in the beta subunits directly induced by the binding of ADP (Cheng, A., Fitzgerald, T. J., and Carlson, G. M. (1985) J. Biol. Chem. 260, 2535-2542; Trempe, M. R., and Carlson, G. M. (1987) J. Biol. Chem. 262, 4333-4340; Cheng, A., Fitzgerald, T. J., Bhatnager, D., Roskoski, R., Jr., and Carlson, G. M. (1988) J. Biol. Chem. 263, 5534-5542). The catalytic site exhibited high stereospecificity for inhibition by the Rp and Sp epimers of adenosine 5'-O-(1-thiodiphosphate), with the Rp epimer (Ki = 0.5 microM) being 136-fold more effective than its Sp counterpart. This can readily explain the inability of the Rp epimer to be an effective allosteric activator.  相似文献   

2.
A total of 26 different purine nucleotides with specific modifications in the base moiety and/or in the polyphosphate chain as well as various combinations of nucleotides were tested as allosteric effectors of beef liver glutamate dehydrogenase (L-glutamate : NAD(P)+ oxidoreductase (deaminating), EC 1.4.1.3). The capacity of these nucleotide analogs to activate or to inhibit the glutamate dehydrogenase activity is expressed quantitatively and scaled between the extreme effects of ADP and GTP, respectively. The significance of distinct structural elements for the enzyme-effector interaction is discussed. While the inhibitory GTP site is less specific, accepting many natural and most modified nucleoside triphosphates as inhibitors, the activating ADP site shows a much higher specificity for nucleotides as activators.  相似文献   

3.
The reduction of purine nucleoside diphosphates by murine ribonucleotide reductase requires catalytic (R1) and free radical-containing (R2) enzyme subunits and deoxynucleoside triphosphate allosteric effectors. A quantitative 16 species model is presented, in which all pertinent equilibrium constants are evaluated, that accounts for the effects of the purine substrates ADP and GDP, the deoxynucleoside triphosphate allosteric effectors dGTP and dTTP, and the dimeric murine R2 subunit on both the quaternary structure of murine R1 subunit and the dependence of holoenzyme (R1(2)R2(2)) activity on substrate and effector concentrations. R1, monomeric in the absence of ligands, dimerizes in the presence of substrate, effectors, or R2(2) because each of these ligands binds R1(2) with higher affinity than R1 monomer. This leads to apparent positive heterotropic cooperativity between substrate and allosteric effector binding that is not observed when binding to the dimeric protein itself is evaluated. Allosteric activation results from an increase in k(cat) for substrate reduction upon binding of the correct effector, rather than from heterotropic cooperativity between effector and substrate. Neither the allosteric site nor the active site displays nucleotide base specificity: dissociation constants for dGTP and dTTP are nearly equivalent and K(m) and k(cat) values for both ADP and GDP are similar. R2(2) binding to R1(2) shows negative heterotropic cooperativity vis-à-vis effectors but positive heterotropic cooperativity vis-à-vis substrates. Binding of allosteric effectors to the holoenzyme shows homotropic cooperativity, suggestive of a conformational change induced by activator binding. This is consistent with kinetic results indicating full dimer activation upon binding a single equivalent of effector per R1(2)R2(2).  相似文献   

4.
Here we report the first three-dimensional structure of a phosphoribosylpyrophosphate (PRPP) synthetase. PRPP is an essential intermediate in several biosynthetic pathways. Structures of the Bacillus subtilis PRPP synthetase in complex with analogs of the activator phosphate and the allosteric inhibitor ADP show that the functional form of the enzyme is a hexamer. The individual subunits fold into two domains, both of which resemble the type I phosphoribosyltransfereases. The active site is located between the two domains and includes residues from two subunits. Phosphate and ADP bind to the same regulatory site consisting of residues from three subunits of the hexamer. In addition to identifying residues important for binding substrates and effectors, the structures suggest a novel mode of allosteric regulation.  相似文献   

5.
The catalytic and allosteric sites of proton translocating adenosine triphosphatase (ATPase) were studied by measuring the binding of nucleotides to the ATPase, and its alpha and beta subunits purified from thermophilic bacterium PS3, with a circular dichroic spectrometer. In contrast to mesophilic ATPases, this thermophilic enzmye contained no tightly bound nucleotides, and its subunits were stable after their purification. These properties were advantageous for analyzing both catalytic and allosteric sites. The former site showed rapid and loose binding, but the latter slow (t 1/2 = 1 h, for ADP) and tight binding. When a nucleotide was bound, the beta subunits showed a negative ellipticity at 275 nm corresponding to a tyrosyl residue, while the alpha subunits showed an ellipticity change corresponding to the absorption curve of the bound nucleotide. This difference enabled us to distinguish the binding sites in ATPase. At a low concentration, ADP selectively bound to alpha subunits in the ATPase, while at a high concentration, it bound to both subunits. This finding suggests that the tight binding sites are located in the alpha subunits. Although ADP and ATP bound to both the purified alpha and beta subunits, CTP did not bind to beta but only to alpha subunits, and ITP bound to beta but hardly to alpha. These nucleotide specificities also supported the idea that the catalytic sites are located in the beta subunits and the allosteric sites are located in the alpha subunits.  相似文献   

6.
The binding of the allosteric activator, AMP, and the inhibitor, ATP, to glycogen phosphorylase b has been studied in the crystal at 3 Å resolution. The nucleotides bind to two sites on the enzyme which are identified as site N, the allosteric effector site which is close to the subunit-subunit interface, and site I, a nucleoside inhibitor site which blocks the entrance to the active site crevasse. AMP when bound at the allosteric effector site makes several defined interactions with the enzyme in agreement with the results of solution studies. The contacts involve the N-10 position of the base, the 2′ hydroxyl of the ribose and the phosphate. IMP, analysed at 4 Å resolution, appears to bind in an identical conformation to AMP. At 3 Å resolution no well defined conformational changes are observed on binding AMP, although there are indications of a disturbance of the crystal lattice. It is concluded that the forces which stabilise the crystal lattice prevent the allosteric response of the enzyme in the crystal.  相似文献   

7.
Mammalian NAD-dependent isocitrate dehydrogenase is an allosteric enzyme, activated by ADP and composed of 3 distinct subunits in the ratio 2alpha:1beta:1gamma. Based on the crystal structure of NADP-dependent isocitrate dehydrogenases from Escherichia coli, Bacillus subtilis, and pig heart, and a comparison of their amino acid sequences, alpha-Arg88, beta-Arg99, and gamma-Arg97 of human NAD-dependent isocitrate dehydrogenase were chosen as candidates for mutagenesis to test their roles in catalytic activity and ADP activation. A plasmid harboring cDNA that encodes alpha, beta, and gamma subunits of the human isocitrate dehydrogenase (Kim, Y. O., Koh, H. J., Kim, S. H., Jo, S. H., Huh, J. W., Jeong, K. S., Lee, I. J., Song, B. J., and Huh, T. L. (1999) J. Biol. Chem. 274, 36866-36875) was used to express the enzyme in isocitrate dehydrogenase-deficient E. coli. Wild type (WT) and mutant enzymes (each containing 2 normal subunits plus a mutant subunit with alpha-R88Q, beta-R99Q, or gamma-R97Q) were purified to homogeneity yielding enzymes with 2alpha:1beta:1gamma subunit composition and a native molecular mass of 315 kDa. Specific activities of 22, 14, and 2 micromol of NADH/min/mg were measured, respectively, for WT, beta-R99Q, and gamma-R97Q enzymes. In contrast, mutant enzymes with normal beta and gamma subunits and alpha-R88Q mutant subunit has no detectable activity, demonstrating that, although beta-Arg99 and gamma-Arg97 contribute to activity, alpha-Arg88 is essential for catalysis. For WT enzyme, the Km for isocitrate is 2.2 mm, decreasing to 0.3 mm with added ADP. In contrast, for beta-R99Q and gamma-R97Q enzymes, the Km for isocitrate is the same in the absence or presence of ADP, although all the enzymes bind ADP. These results suggest that beta-Arg99 and gamma-Arg97 are needed for normal ADP activation. In addition, the gamma-R97Q enzyme has a Km for NAD 10 times that of WT enzyme. This study indicates that a normal alpha subunit is required for catalytic activity and alpha-Arg88 likely participates in the isocitrate site, whereas the beta and gamma subunits have roles in the nucleotide functions of this allosteric enzyme.  相似文献   

8.
The beta2 adrenoreceptor (beta2AR) is a prototypical G protein-coupled receptor (GPCR) activated by catecholamines. Agonist activation of GPCRs leads to sequential interactions with heterotrimeric G proteins, which activate cellular signaling cascades, and with GPCR kinases and arrestins, which attenuate GPCR-mediated signaling. We used fluorescence spectroscopy to monitor catecholamine-induced conformational changes in purified beta2AR. Here we show that upon catecholamine binding, beta2ARs undergo transitions to two kinetically distinguishable conformational states. Using a panel of chemically related catechol derivatives, we identified the specific chemical groups on the agonist responsible for the rapid and slow conformational changes in the receptor. The conformational changes observed in our biophysical assay were correlated with biologic responses in cellular assays. Dopamine, which induces only a rapid conformational change, is efficient at activating Gs but not receptor internalization. In contrast, norepinephrine and epinephrine, which induce both rapid and slow conformational changes, are efficient at activating Gs and receptor internalization. These results support a mechanistic model for GPCR activation where contacts between the receptor and structural determinants of the agonist stabilize a succession of conformational states with distinct cellular functions.  相似文献   

9.
Extensively purified rat liver cytosolic 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase kinase was used to examine the role of ADP in inactivation of HMG-CoA reductase (EC 1.1.1.34). Solubilized HMG-CoA reductase was a suitable substrate for HMG-CoA reductase kinase. At sufficiently high concentrations of solubilized HMG-CoA reductase, reductase kinase activity approached that measured using microsomal HMG-CoA reductase as substrate. Inactivation of solubilized HMG-CoA reductase by HMG-CoA reductase kinase required both MgATP and ADP. Other nucleoside diphosphates, including alpha, beta-methylene-ADP, could replace ADP. HMG-CoA reductase kinase catalyzed phosphorylation of bovine serum albumin fraction V by [gamma-32P]ATP. This process also required a nucleoside diphosphate (e.g. alpha, beta-methylene-ADP). Nucleoside diphosphates thus act on HMG-CoA reductase kinase, not on HMG-CoA reductase. For inactivation of HMG-CoA reductase, the ability of nucleoside triphosphates to replace ATP decreased in the order ATP greater than dATP greater than GTP greater than ITP, UTP. TTP and CTP did not replace ATP. Both for inactivation of HMG-CoA reductase and for phosphorylation of bovine serum albumin protein, the ability of nucleoside diphosphates to replace ADP decreased in the order ADP greater than CDP, dADP greater than UDP. GDP did not replace ADP. Nucleoside di- and triphosphates thus appear to bind to different sites on HMG-CoA reductase kinase. Nucleoside diphosphates act as allosteric activators of HMG-CoA reductase kinase. For inactivation of HMG-CoA reductase by HMG-CoA reductase kinase, Km for ATP was 140 microM and the activation constant, Ka, for ADP was 1.4 mM. The concentration of ADP required to modulate reductase kinase activity in vitro falls within the physiological range. Modulation of HMG-CoA reductase kinase activity, and hence of HMG-CoA reductase activity, by changes in intracellular ADP concentrations thus may represent a control mechanism of potential physiological significance.  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.
Levels of nucleotides and sugar nucleotides in embryos of Bufo arenarum at the stages of morula, gastrula, and neurula have been measured. The total amounts of purine nucleoside diphosphates decreased from morula to gastrula, but increased sharply from gastrula to neurula. The levels of ADP followed this pattern, but those of GDP did not change significantly through the three stages. Purine nucleoside triphosphate levels, which had increased immediately after fertilization, remained almost constant through morula, gastrula, and neurula. As with the purine nucleoside diphosphates, the adenine nucleotide decreased from morula to gastrula, and increased from gastrula to neurula. In contrast, the level of GTP showed a sharp maximum at gastrula. The total pyrimidine nucleoside triphosphate did not change significantly from morula through neurula. As in previous stages of development, only uridine sugar nucleotides were detected. A sharp increase of the galactosyl ester of nucleotides was found at gastrula.  相似文献   

12.
It was found that nucleoside 5'-diphosphates could serve as effectors of ribonucleotide reductase. ADP was an activator of CDP reduction; ADP reduction was activated by dGDP; GDP reduction was activated by dTDP. Conversely, dADP inhibited the reduction of CDP, UDP, GDP, and ADP; dGDP inhibited UDP and GDP reductions; and dTDP inhibited UDP reduction. The inhibition of UDP reduction by dADP, dTDP, and dGDP was at least equal to that observed for dATP, dTTP, and dGTP, respectively. In these experiments with the nucleoside diphosphates as effectors, high-pressure liquid chromatography analysis of the reaction mixtures showed that no nucleoside 5'-triphosphates were found during the reaction period which could account for the effects seen with the nucleoside diphosphates as effectors. Further experiments were carried out in which adenyl-5'-yl imidodiphosphate was used as the positive effector of CDP and UDP reductions in place of ATP. Under these conditions, CDP and UDP reductions were inhibited by dADP, dTDP, and dGDP to the same extent observed in the presence of ATP. ADP served not only as a substrate for ribonucleotide reductase but also as an activator of CDP and UDP reductions. The direct products (dNDPs) also served as positive and negative effectors. Dixon plots indicated that the dNDPs were acting as noncompetitive inhibitors with respect to the substrate. ADP increased the sedimentation velocity of the ribonucleotide reductase in a manner similar to ATP. These data are consistent with the allosteric effects seen with the nucleoside 5'-triphosphates. Additionally, from the thorough study of the role of effectors on UDP reduction, it is clear that UDP reduction was most sensitive to the negative effectors dATP, dADP, dTTP, dTDP, dGTP, and dGDP.  相似文献   

13.
The beta subunit isolated from the chloroplast ATP synthase F1 (CF1) has a single dissociable nucleotide binding site, consistent with the proposed function of this subunit in nucleotide binding and catalysis. The beta subunit bound the nucleotide analogs trinitrophenyl-ATP (TNP-ATP) or trinitrophenyl-ADP (TNP-ADP) with nearly equal affinities (Kd = 1-2 microM) but did not bind trinitrophenyl-AMP. Both ATP and ADP effectively competed with TNP-ATP for binding. Other nucleoside triphosphates were also able to compete with TNP-ATP for binding to beta; their order of effectiveness (ATP greater than GTP, ITP greater than CTP) mimicked the normal substrate specificity of CF1. The single nucleotide binding site on the isolated beta subunit very closely resembles the low affinity catalytic site (site 3) of CF1 (Bruist, M.F., and Hammes, G. G. (1981) Biochemistry 20, 6298-6305), suggesting that tight nucleotide binding by other sites on the enzyme involves other CF1 subunits in addition to the beta subunit. The results are inconsistent with an earlier report (Frasch, W.D., Green, J., Caguial, J., and Mejia, A. (1989) J. Biol. Chem. 264, 5064-5069), which suggested more than one nucleotide binding site per beta subunit. Binding of nucleotides to the isolated beta subunit was eliminated by a brief heat treatment (40 degrees C for 10 min) of the protein. A small change in the circular dichroism spectrum of beta accompanied the heat treatment indicating that a localized (rather than global) change in the folding of beta, involving at least part of the nucleotide binding domain, had occurred. Also accompanying the loss of nucleotide binding was a loss of the reconstitutive capacity of the beta subunit. ATP protected against the effects of the heat treatment.  相似文献   

14.
A variety of proteases have been evaluated as potential structural and conformational probes of nonphosphorylated and phosphorylated phosphorylase kinase. In general, the enzyme's alpha subunit is rapidly degraded, followed in most cases by hydrolysis of the beta subunit; the gamma subunit is resistant to most proteases. Trypsin clearly distinguishes between the nonactivated and activated conformers of phosphorylase kinase, in that the beta subunit in phosphorylated enzyme, as opposed to nonphosphorylated enzyme, is markedly protected from tryptic attack. In contrast, only a small difference in the rates of proteolysis of the alpha subunit in phosphorylated and nonphosphorylated enzyme is seen, even when a protease is used that is highly selective for the alpha subunit, such as chymotrypsin or endoproteinase Arg C. Incubation of nonphosphorylated phosphorylase kinase with either Mg2+ or Ca2+, which are activating cations, also protects the beta subunit from tryptic hydrolysis, whereas Mn2+, which inhibits the kinase activity, has little effect on proteolysis. The allosteric activator ADP also causes the beta subunit to become refractory to trypsin and mimics the effects of phosphorylation. Similar effector-induced conformational changes in the beta subunit are also observed with enzyme in which the alpha subunit has previously been selectively destroyed. These data indicate that activation of phosphorylase kinase by dissimilar mechanisms is associated with a conformational change in the enzyme's beta subunit that is detectable by trypsin and confirm earlier studies from this laboratory employing a chemical cross-linker as a conformational probe for activated and nonactivated conformers of the enzyme (Fitzgerald, T. J., and Carlson, G. M. (1984) J. Biol. Chem. 259, 3266-3274).  相似文献   

15.
Microsomal 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase kinase activity is enhanced about 5 fold by 2 mM of either AMP or ADP. Activation constants, Ka, for AMP and ADP are 17 microM and 430 microM respectively, showing that AMP is a more potent activator than ADP. This property is expressed by increasing not only the rate of reductase inactivation but also the rate of reductase phosphorylation from [gamma-32P]ATP. GTP can replace ATP as substrate of reductase kinase but GMP and GDP cannot replace AMP as activators. Kinetic studies show that ATP can only act as a substrate. Nucleoside mono or diphosphates and nucleoside triphosphates, thus, appear to bind to different sites on microsomal HMG-CoA reductase kinase. Nucleoside mono or diphosphates act as allosteric activators of reductase kinase. The adenosyl moiety and the unaltered phosphate ester at the 5' position are two essential features of the activator molecule. Phosphorylation of reductase either by microsomal or cytosolic AMP-activated reductase kinase produces an 80% inactivation, with a concomitant incorporation of 0.8 mol of 32P per mol of reductase (Mr 55,000). In both cases exhaustive tryptic digestion of 32P-labeled HMG-CoA reductase, which had been denatured in 2M urea, yields two major phosphopeptides, the phosphoryl group being bound to serine residues.  相似文献   

16.
Comparison of AMP and NADH binding to glycogen phosphorylase b   总被引:3,自引:0,他引:3  
The binding sites for the allosteric activator, AMP, to glycogen phosphorylase b are described in detail utilizing the more precise knowledge of the native structure obtained from crystallographic restrained least-squares refinement than has hitherto been available. Localized conformational changes are seen at the allosteric effector site that include shifts of between 1 and 2 A for residues Tyr75 and Arg309 and very small shifts for the region of residues 42 to 44 from the symmetry-related subunit. Kinetic studies demonstrate that NADH inhibits the AMP activation of glycogen phosphorylase b. Crystallographic binding studies at 3.5 A resolution show that NADH binds to the same sites on the enzyme as AMP, i.e. the allosteric effector site N, which is close to the subunit-subunit interface, and the nucleoside inhibitor site I, which is some 12 A from the catalytic site. The conformations of NADH at the two sites are different but both conformations are "folded" so that the nicotinamide ring is close (approx. 6 A) to the adenine ring. These conformations are compared with those suggested from solution studies and with the extended conformations observed in the single crystal structure of NAD+ and for NAD bound to dehydrogenases. Possible mechanisms for NADH inhibition of phosphorylase activation are discussed.  相似文献   

17.
UDP-glucose is an R-state inhibitor of glycogen phosphorylase b, competitive with the substrate, glucose 1-phosphate and noncompetitive with the allosteric activator, AMP. Diffusion of 100 mM UDP-glucose into crystals of phosphorylase b resulted in a difference Fourier synthesis at 0.3-nm resolution that showed two peaks: (a) binding at the allosteric site and (b) binding at the catalytic site. At the allosteric site the whole of the UDP-glucose molecule can be located. It is in a well defined folded conformation with its uracil portion in a similar position to that observed for the adenine of AMP. The uracil and the glucose moieties stack against the aromatic side chains of Tyr-75 and Phe-196, respectively. The phosphates of the pyrophosphate component interact with Arg-242, Arg-309 and Arg-310. At the catalytic site, the glucose-1-P component of UDP-glucose is firmly bound in a position similar to that observed for glucose 1-phosphate. The pyrophosphate is also well located with the glucose phosphate interacting with the main-chain NH groups at the start of the glycine-loop alpha helix and the uridine phosphate interacting through a water molecule with the 5'-phosphate of the cofactor pyridoxal phosphate and with the side chains of residues Tyr-573, Lys-574 and probably Arg-569. However the position of the uridine cannot be located although analysis by thin-layer chromatography showed that no degradation had taken place. Binding of UDP-glucose to the catalytic site promotes extensive conformational changes. The loop 279-288 which links the catalytic site to the nucleoside inhibitor site is displaced and becomes mobile. Concomitant movements of residues His-571, Arg-569, and the loop 378-383, together with the major loop displacement, result in an open channel to the catalytic site. Comparison with other structural results shows that these changes form an essential feature of the T to R transition. They allow formation of the phosphate recognition site at the catalytic site and destroy the nucleoside inhibitor site. Kinetic experiments demonstrate that UDP-glucose activates the enzyme in the presence of high concentrations of the weak activator IMP, because of its ability to decrease the affinity of IMP for the inhibitor site.  相似文献   

18.
The binding of divalent cations and nucleotide to bovine brain glutamine synthetase and their effects on the activity of the enzyme were investigated. In ADP-supported gamma-glutamyl transfer at pH 7.2, kinetic analyses of saturation functions gave [S]0.5 values of approximately 1 microM for Mn2+, approximately 2 mM for Mg2+, 19 nM for ADP.Mn, and 7.2 microM for ADP.Mg. The method of continuous variation applied to the Mn2+-supported reaction indicated that all subunits of the purified enzyme express activity when 1.0 equiv of ADP is bound per subunit. Measurements of equilibrium binding of Mn2+ to the enzyme in the absence and presence of ADP were consistent with each subunit binding free Mn2+ (KA approximately equal to 1.5 X 10(5) M-1) before binding the Mn.ADP complex (KA' approximately equal to 1.1 X 10(6) M-1). The binding of the first Mn2+ or Mg2+ to each subunit produces structural perturbations in the octameric enzyme, as evidenced by UV spectral and tryptophanyl residue fluorescence changes. The enzyme, therefore, has one structural site per subunit for Mn2+ or Mg2+ and a second site per subunit for the metal ion-nucleotide complex, both of which must be filled for activity expression. Chloride binding (KA' approximately equal to 10(4) M-1) to the enzyme was found to have a specific effect on the protein conformation, producing a substantial (30%) quench of tryptophanyl fluorescence and increasing the affinity of the enzyme 2-4-fold for Mg2+ or Mn2+. Arsenate, which activates the gamma-glutamyl transfer activity by binding to an allosteric site, and L-glutamate also cause conformational changes similar to those produced by Cl- binding. Anion binding to allosteric sites and divalent metal ion binding at active sites both produce tryptophanyl residue exposure and tyrosyl residue burial without changing the quaternary enzyme structure.  相似文献   

19.
S P Ong  L S Nelson  H P Hogenkamp 《Biochemistry》1992,31(45):11210-11215
Two nucleoside diphosphate analogs, 3'-C-methyl-ADP and 3'-C-methyl-UDP, have been tested as substrate and/or allosteric effectors using the adenosylcobalamin-dependent ribonucleoside diphosphate reductase of Corynebacterium nephridii. Neither analog was a substrate for the reductase. However, they did function as allosteric effectors and as inhibitors of the reduction of ADP and UDP, respectively. The nucleotide analogs did not stimulate the hydrogen exchange reaction between [5'-3H2]adenosylcobalamin and the solvent, indicating that the cleavage of the 3'-carbon-hydrogen bond is a prerequisite for the exchange reaction. A reinvestigation of the requirements for the exchange reaction revealed that the deoxyribonucleoside diphosphate products are very effective promoters of this reaction. Indeed, the deoxyribonucleoside diphosphates were found to be more effective in promoting the exchange reaction than the ribonucleoside diphosphate substrates. In contrast, the deoxyribonucleoside triphosphate effectors, dATP, dUTP, and dTTP, were only marginally effective as promoters of this reaction.  相似文献   

20.
Various nucleoside di- and triphosphates have been compared with respect to their ability to protect rat brain hexokinase (ATP: d-hexose 6-phosphotransferase, EC 2.7.1.1) activity against inactivation by chymotrypsin, glutaraldehyde, heat, and 5,5′-dithiobis(2-nitrobenzoic) acid. ATP could be distinguished from other nucleoside triphosphates in these comparisons, which may be related to the specificity with which ATP is utilized as a substrate. All nucleoside derivatives examined provided substantial protection against two or more of the above inactivating agents, indicating relatively nonspecific binding of nucleotides by brain hexokinase, consistent with a similar lack of specificity in the inhibition of this enzyme by nucleoside derivatives. The fluorescence of 2-p-toluidinylnaphthalene-6-sulfonate (TNS) and of tetraiodofluorescein (TIF) was enhanced by binding to brain hexokinase. TNS binding was not affected by the presence of various relevant metabolites (Glc, glucose 6-phosphate, ATP), nor did TNS inhibit the enzyme. In contrast, substantial (approximately 70%) decreases in the fluorescence of bound TIF resulted from the addition of various nucleoside derivatives, and TIF served as a competitive inhibitor of brain hexokinase. These observations are consistent with the view that TIF binds to a nucleotide binding site of the enzyme. The inability of nucleotides to totally displace TIF was taken to indicate the existence of an additional TIF binding site (or sites) discrete from the catalytic site, and probably identical to the site(s) at which TNS binds with no effect on catalytic activity. The effects of saturating levels of ATP and ADP were not additive indicating that both compounds were displacing TIF from the same site i.e., a common nucleotide binding site. Glc, mannose, and 2-deoxyglucose greatly enhanced the ability of nucleotides to displace TIF, while fructose, galactose, and N-acetylglucosamine did not, indicating the existence of interactions between hexose and nucleotide binding sites; the hexoses themselves were not effective at displacing TIF. The enhanced binding of nucleotides in the presence of the first three hexoses but not the latter three can be directly correlated with the relative ability of these hexoses to induce specific conformational changes in the enzyme. The hexoses themselves were not effective at displacing TIF. Glucose 6-phosphate and 1,5-anhydroglucitol 6-phosphate could also displace TIF, and as with the nucleotides, a maximum of approximately 70% decrease in fluorescence was observed and the effectiveness of glucose 6-phosphate was enhanced in the presence of Glc. Other hexose 6-phosphates tested were not effective at displacing TIF. The specificity with which hexose 6-phosphates displaced TIF could be correlated with their ability to induce specific conformational change in the enzyme. The results are discussed as they relate to the kinetic mechanism and allosteric regulation by nucleotides that have been proposed for this enzyme.  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