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.
Numerous hydrophobic and amphipathic compounds including several detergents are known to inhibit the ubiquinone reductase reaction of respiratory chain complex I (proton pumping NADH:ubiquinone oxidoreductase). Guided by the X-ray structure of the peripheral arm of complex I from Thermus thermophilus we have generated a large collection of site-directed mutants in the yeast Yarrowia lipolytica targeting the proposed ubiquinone and inhibitor binding pocket of this huge multiprotein complex at the interface of the 49-kDa and PSST subunits. We could identify a number of residues where mutations changed I(50) values for representatives from all three groups of hydrophobic inhibitors. Many mutations around the domain of the 49-kDa subunit that is homologous to the [NiFe] centre binding region of hydrogenase conferred resistance to DQA (class I/type A) and rotenone (class II/type B) indicating a wider overlap of the binding sites for these two types of inhibitors. In contrast, a region near iron-sulfur cluster N2, where the binding of the n-alkyl-polyoxyethylene-ether detergent C(12)E(8) (type C) was exclusively affected, appeared comparably well separated. Taken together, our data provide structure-based support for the presence of distinct but overlapping binding sites for hydrophobic inhibitors possibly extending into the ubiquinone reduction site of mitochondrial complex I.  相似文献   

2.
The binding characteristics of inhibitors of the mitochondrial cytochrome c reductase were studied by fluorescence quench titration. Based on the standard binding equation, the applied numerical method allowed the online recorded titration curves to be interpreted by fitting the Kd, the number of binding sites, and the specific fluorescence of the free and the bound inhibitor. For the Qi center, 2-n-nonyl-4-hydroxyquinoline N-oxide and for the Qo center (E)-beta-methoxyacrylate-stilbene (MOA-stilbene) were used as fluorescing inhibitors. The experiments could be extended to other, non-fluorescing inhibitors by competition analysis. Using this method we were able to compare the binding behaviour of Qi and Qo center inhibitors under different redox states of the enzyme using the same experimental set up. We studied the competition between inhibitors of the cytochrome c reductase representative for all subgroups and demonstrated that at least three inhibitor binding sites exist, two located in the Qo center, one located in the Qi center. Determination of the dissociation constants of the oxidized, the partially reduced and the fully reduced enzyme showed that inhibitor binding at the Qi center is not redox-dependent. In contrast, the binding of MOA-stilbene to the Qo center is decreased after reduction of the iron-sulfur center and cytochrome c1, whereas this redox change increases the affinity for a Qo center inhibitor of the hydroxynaphthoquinone type, 3-n-undecyl-2-hydroxynaphthoquinone. From these results, aware of the fact that the inhibitory mechanism at the Qo center is a non-competitive one, we made the hypothesis of a 'catalytic switch' to explain both the bifurcation of electron flow and the inhibition at the Qo center. A steric blockage of one of two conformational states could serve as a cogent explanation for the great structural variability of the inhibitors and differential effects on the redox centers exerted by the inhibitors. Moreover, the proposed 'switch' gives some insight into other experimental results which are difficult to explain with the ubiquinone cycle as currently formulated.  相似文献   

3.
We have synthesized Deltalac-acetogenins that are new acetogenin mimics possessing two n-alkyl tails without an alpha,beta-unsaturated gamma-lactone ring and suggested that their inhibition mechanism may be different from that of common acetogenins [Hamada et al. (2004) Biochemistry 43, 3651-3658]. To elucidate the inhibition mechanism of Deltalac-acetogenins in more detail, we carried out wide structural modifications of original Deltalac-acetogenins and characterized the inhibitory action with bovine heart mitochondrial complex I. In contrast to common acetogenins, both the presence of adjacent bis-THF rings and the stereochemistry around the hydroxylated bis-THF rings are important structural factors required for potent inhibition. The inhibitory potency of a derivative possessing an n-butylphenyl ether structure (compound 7) appeared to be superior to that of the original Deltalac-acetogenins and equivalent to that of bullatacin, one of the most potent natural acetogenins. Double-inhibitor titration of steady-state complex I activity showed that the extent of inhibition of compound 7 and bullatacin is not additive, suggesting that the binding sites of the two inhibitors are not identical. Competition tests using a fluorescent ligand indicated that the binding site of compound 7 does not overlap with that of other complex I inhibitors. The effects of compound 7 on superoxide production from complex I are also different from those of other complex I inhibitors. Our results clearly demonstrate that Deltalac-acetogenins are a novel type of inhibitor acting at the terminal electron-transfer step of bovine complex I.  相似文献   

4.
Murai M  Ishihara A  Nishioka T  Yagi T  Miyoshi H 《Biochemistry》2007,46(21):6409-6416
The inhibitor binding domain in bovine complex I is believed to be constructed by multisubunits, but it remains to be learned how the binding positions of chemically diverse inhibitors relate to each other. To get insight into the inhibitor binding domain in complex I, we synthesized a photoreactive acetogenin [[125I](trifluoromethyl)phenyldiazirinylacetogenin, [125I]TDA], in which an aryldiazirine group serves as both a photoreactive group and a substitute for the gamma-lactone ring that is a common toxophore of numerous natural acetogenins, and carried out photoaffinity labeling to identify the labeled subunit using bovine heart submitochondrial particles (SMP). When SMP were UV-irradiated in the presence of [125I]TDA, radioactivity was predominantly incorporated into an approximately 30 kDa band on a SDS gel. Blue native gel electrophoresis of the [125I]TDA-labeled SMP revealed that the majority of radioactivity was observed in complex I. Analysis of complex I on a SDS gel showed a predominant peak of radioactivity at approximately 30 kDa. Immnoprecipitation of the [125I]TDA-labeled complex I with anti-bovine ND1 antibody indicated that the labeled protein is the ND1 subunit. A variety of complex I inhibitors such as piericidin A and rotenone efficiently suppressed the specific binding of [125I]TDA to ND1, indicating that they share a common binding domain. However, the suppression efficiency of Deltalac-acetogenin, a new type of complex I inhibitor synthesized in our laboratory, was much lower than that of the traditional inhibitors. Our results unequivocally reveal that the ND1 subunit constructs the inhibitor binding domain, though the contribution of this subunit has been challenged. Further, the present study corroborates our previous proposition that the inhibition site of Deltalac-acetogenins differs from that of traditional inhibitors.  相似文献   

5.
F1-ATPase inhibitor was purified from yeast, Saccharomyces cerevisiae. The purified inhibitor blocked ATPase activity in the presence of ATP and Mg2+ by forming a latent equimolar enzyme-inhibitor complex with ATP and ADP newly bound to loose sites on the enzyme. A small portion of externally added ATP was hydrolyzed before the latent complex was formed but the hydrolysis was not directly related to the complex formation. Newly bound ATP tended to be converted to ADP when the ATP concentration of the medium was low. ATP tightly bound to the enzyme was not directly involved in formation of the complex. The complex was fairly stable in the presence of excess inhibitor and ATP but at a high concentration of the enzyme (10(-5) M), the inhibition was not complete, although only about 0.03% of the original activity remained unblocked.  相似文献   

6.
Imino- and azasugar glycosidase inhibitors display pH dependant inhibition reflecting that both the inhibitor and the enzyme active site have groups that change protonation state with pH. With the enzyme having two acidic groups and the inhibitor one basic group, enzyme–inhibitor complexes with three (EH3I), two (EH2I), one (EHI), or no protons (EI), are possible. In the present work an analysis method is presented that from pH-inhibition data allows one to distinguish between the different complexes and determine which protonation state is preferred. It is also possible to determine the pH-independent binding constants of the inhibitor. Analysis of pH data for imino- and azasugar inhibition of β-glucosidases revealed that basic glycosidase inhibitors bind as the monoprotonated (EHI) complex. Three neutral inhibitors were also studied and two of these were also bound exclusively as the EHI complex while a third bound both as a EHI and a EH2I complex.  相似文献   

7.
Positively charged pyridiniums are unique inhibitors to probe the structural and functional properties of the ubiquinone reduction site of bovine heart mitochondrial complex I. In this study, we synthesized a series of neutral as well as pyridinium analogues of MP-24 (N-methyl-4-[2-methyl-2-(p-tert-butylbenzyl)propyl]pyridinium), a selective inhibitor of one of the two proposed binding sites of these pyridinium-type inhibitors of complex I (H. Miyoshi et al., J. Biol. Chem. 273 (1998) 17368-17374), to elucidate the origin of its selectivity. Inhibitory potencies of all neutral and pyridinium analogues with tetraphenylboron (TPB(-)), which forms an ion-pair with pyridiniums, were comparable, although the degrees of selective inhibition by pyridiniums without TPB(-) were entirely different. In contrast to MP-24, the dose-response curves of nonselective pyridiniums and all neutral analogues were not affected by incubation conditions. These results strongly suggested that the process of the inhibitor passage to the binding sites is responsible for the selective inhibition.  相似文献   

8.
We have examined the interaction of the nicotinic acetylcholine receptor with decidium diiodide, a bisquaternary analogue of ethidium containing 10 methylene groups between the endocyclic and trimethylamino quaternary nitrogens. Decidium inhibits mono-[125I]iodo-alpha-toxin binding, inhibits agonist-elicited 22Na+ influx in intact cells, augments agonist competition with mono-[125I]iodo-alpha-toxin binding, and enhances [3H]phencyclidine (PCP) binding to a noncompetitive inhibitor site. These effects occur over similar concentration ranges (half-maximum effects between 0.1 and 0.4 microM). Thus, decidium binds to the agonist site and converts the receptor to a desensitized state exhibiting increased affinity for agonist and heterotropic inhibitors. These properties are similar to metaphilic antagonists characterized in classical pharmacology. At higher concentrations decidium associates directly with the noncompetitive inhibitor site identified by [3H]phencyclidine binding. Dissociation constants of decidium at this site in the resting and desensitized states are determined to be 29 and 1.2 microM, respectively. Analysis of fluorescence excitation and emission maxima reveal that binding to both the agonist and noncompetitive inhibitor sites is associated with approximately 2-fold enhancement of fluorescence. The excitation maximum for decidium bound at the agonist site appears at 490 nm while that for decidium bound at the noncompetitive inhibitor site appears at 530 compared to 480 nm in buffer. These results suggest that decidium experiences a more hydrophobic environment upon binding to the nicotinic acetylcholine receptor sites, particularly to the noncompetitive inhibitor site. Fluorescence energy transfer between N'-fluorescein isothiocyanate-lysine-23 alpha-toxin (FITC-toxin), and decidium is not detected when each is bound to one of the two agonist sites on the receptor. This allows a minimal distance to be estimated between fluorophores. In contrast, energy transfer is observed between decidium nonspecifically associated with the membrane or with nonspecific sites and the FITC-toxin at the agonist sites.  相似文献   

9.
The zinc(II) complex of pyridine-2-azo-p-dimethylaniline is bound to several acid proteinases, at pH 5.0, accompanied by a change is the visible absorption spectrum. Streptomyces pepsin inhibitor, which was discovered by Satoi and Murao (Satoi, S. and Murao, S. (1970) Agric. Biol. Chem. 34, 1265-1267 and Satoi, S. and Murao, S. (1971) Agric. Biol. Chem. 35, 1482-1487), is also bound to acid proteinases. Spectrophotometric studies with ten acid proteinases from different sources have revealed that in several acid proteinases, zinc(II)-pyridine-2-azo-p-dimethylaniline is released from the enzyme by the inhibitor, while some acid proteinase forms a quaternary complex, zinc(II)-pyridine-2-azo-p-dimethylaniline-inhibitor-enzyme. It is speculated that zinc(II)-pyridine-2-azo-p-dimethylaniline is bound to two catalytic carboxylate groups in the active site of the acid proteinases and the inhibitor is bound mainly to the substrate-binding site of the enzymes. The binding of the inhibitor may overlap the catalytic site completely or partially. The degree of overlapping is characteristic of the kind of acid proteinases.  相似文献   

10.
Mechanism of action of inter-alpha-trypsin inhibitor   总被引:1,自引:0,他引:1  
C W Pratt  S V Pizzo 《Biochemistry》1987,26(10):2855-2863
Inter-alpha-trypsin inhibitor (I alpha I) is a unique proteinase inhibitor that can be proteolyzed by the same enzymes that are inhibited, to generate smaller inhibitors. This study examines the reactions of I alpha I with trypsin, chymotrypsin, plasmin, and leukocyte elastase. Complexes of I alpha I and proteinase were demonstrated by gel filtration chromatography. Complete digestion of I alpha I by each proteinase was not accompanied by a comparable loss of inhibition of that enzyme or a different enzyme. Following proteolysis, inhibitory activity was identified in I alpha I fragments of molecular weight 50,000-100,000 and less than 40,000. Addition of a second proteinase inhibitor prevented proteolysis. Both I alpha I and its complex with proteinase were susceptible to degradation. Kinetic parameters for both the inhibition and proteolysis reactions of I alpha I with four proteinases were measured under physiological conditions. On the basis of these results, a model for the mechanism of action of I alpha I is proposed: Proteinase can react with either of two independent sites on I alpha I to form an inhibitory complex or a complex that leads to proteolysis. Both reactions occur simultaneously, but the inhibitory capacity of I alpha I is not significantly affected by proteolysis since the product of proteolysis is also an inhibitor. For a given proteinase, the inhibition equilibrium constant and the Michaelis constant for proteolysis describe the relative stability of the inhibition and proteolysis complexes; the second-order rate constants for inhibition and proteolysis indicate the likelihood of either reaction. The incidence of inhibition or proteolysis reactions involving I alpha I in vivo cannot be assessed without knowledge of the exact concentrations of inhibitor and proteinases; however, analysis of inhibition rate constants suggests that I alpha I might be involved in plasmin inhibition.  相似文献   

11.
The interaction of several inhibitors with muscle phosphofructokinase has been studied by both equilibrium binding measurements and kinetic analysis. At low concentrations of citrate a maximum of 1 mol is bound per mol of enzyme protomer. Tight binding requires MgATP and very weak binding is observed in the absence of either magnesium ion or ATP. ITP at low concentrations cannot replace ATP. In the presence of MgATP and at pH 7.0, the dissociation constant for the enzyme-citrate complex is 20 muM. At 50 muM citrate and excess magnesium ion, the concentration of ATP required to give half-maximal binding of citrate is approximately 3 muM . Both P-enolpyruvate and 3-P-glycerate compete for the binding of citrate and the estimated Ki values are 480 and 52 muM, respectively. Creatine-P, another inhibitor of muscle phosphofructokinase, does not compete with the binding of citrate. Measurement of the equilibrium binding of ATP shows that citrate, 3-P-glycerate, P-enolpyruvate, and creatine-P all increase the affinity of enzyme for MgATP with the concentration required to give an effect increasing in the order given. In kinetic studies, citrate, 3-P-glycerate and P-enolpyruvate each act synergistically with ATP to inhibit the phosphofructokinase reaction. This is indicated by the observation that the three metabolites do not inhibit the enzyme with ITP as the phosphoryl donor and that they inhibit at ATP concentrations that are not themselves inhibitory. Furthermore, the sensitivity to the inhibitors increases with increasing ATP concentrations. Striking differences in the extent of inhibition can be seen by varying the order of addition of assay components. Preincubation of the enzyme with ATP and citrate, 3-P-glycerate, or P-enolpyruvate results in greater inhibition than when the inhibitor is added after the reaction is started with fructose-6-P. Furthermore, the inhibition is reversed partially 10 to 15 min after the addition of fructose-6-P. This phenomenon is particularly striking with creatine-P as the inhibitor. Very high concentrations of this inhibitor are required to show any effect if the inhibitor is added after fructose-6-P. These effects are interpreted as reflecting slow conformational changes between an active form with high affinity for fructose-6-P and an inactive, or less active, conformation that binds the inhibitors. Citrate, 3-P-glycerate, P-enolpyruvate, and creatine-P increase the rate of the phosphofructokinase at subsaturating concentrations of MgITP. The results indicate a common binding site on the enzyme for citrate, 3-P-glycerate, and P-enolpyruvate that is distinct from the ATP inhibitory site. An additional site (or sites) for creatine-P is indicated. All four inhibitors act synergistically with ATP by increasing the affinity of the enzyme for MgATP at an inhibitory site. The inhibitors appear also to increase the affinity of the catalytic nucleoside triphosphate site for substrate.  相似文献   

12.
Inhibition of porcine pancreatic alpha-amylase (1,4-alpha-D-glucan glucanohydrase) [EC 3.2.1.1] with maltotriitol (G3OH) and 4-phenylimidazole was investigated by using maltohexaitol (G6OH) and p-nitrophenyl-alpha-D-maltoside (G2PNP) as substrates. When G6OH was the substrate, both G3OH and 4-phenylimidazole behaved as competitive inhibitors. On the other hand, when G2PNP was the substrate, G3OH behaved as a competitive inhibitor, whereas 4-phenylimidazole behaved as a non-competitive inhibitor. Further inhibition study in the presence of both G3OH and 4-phenylimidazole, with G6OH as the substrate, showed that the two inhibitors compete with each other for the active site of the enzyme. Based on a consideration of the productive (reactive) binding modes of G2PNP and G6OH, and a nonproductive (nonreactive) binding mode of G2PNP, it is suggested that the binding sites of the two inhibitors may be partially overlapping around the catalytic site of the enzyme and that the rest of the binding site of each inhibitor lies along the substrate binding cleft of the enzyme.  相似文献   

13.
J K Wright  M Takahashi 《Biochemistry》1977,16(8):1541-1548
The aspartokinase activity of the aspartokinase-homoserine dehydrogenase complex of Escherichia coli was affinity labeled with substrates ATP, aspartate, and feedback inhibitor threonine. Exchange-inert ternary adducts of Co(III)-aspartokinase and either ATP, aspartate or threonine were formed by oxidation of corresponding Co(II) ternary complexes with H2O2. The ternary enzyme-Co(III)-threonine adduct (I) had 3.8 threonine binding sites per tetramer, one-half that of the native enzyme. The binding of threonine to I was still cooperative as determined by equilibrium dialysis (nH = 2.2) or by studying inhibition of residual dehydrogenase activity (nH = 2.7). Threonine still protected the SH groups of I against 5,5'-dithiobis(2-nitrobenzoate) (DTNB) reaction but the number of SH groups reacting with thiol reagents (DTNB) was reduced by 1-2 per subunit in the absence of threonine. This suggests either that Co(III) is bound to the enzyme via sulfhydryl groups or that 1-2SH groups are buried or rendered inaccessible in I. The binding of threonine to sites not blocked by the affinity labeling produced changes in the circular dichroism of the complex comparable to changes produced by threonine binding to native enzyme and also protected against proteolytic digestion. The major conformational changes produced by threonine are thus ascribable to binding at this one class of regulatory sites. The interactions of kinase substrates with various aspartokinase-Co(III) complexes containing ATP, aspartate, or threonine and a threonine-insensitive homoserine dehydrogenase produced by mild proteolysis were studied. The inhibition of homoserine dehydrogenase by kinase substrates is not due to binding of these inhibitors at the kinase active site but was shown to be due to binding to sites within the dehydrogenase domain of the enzyme. L-alpha-Aminobutyrate, a presumed threonine analogue, also inhibits the dehydrogenase by binding at the same or similar sites in the dehydrogenase domain and not at threonine regulatory site.  相似文献   

14.
B Mao 《Biophysical journal》1991,60(4):966-973
The mass-weighted molecular dynamics simulation method was developed previously for sampling the multidimensional conformational space of linear and cyclic polypeptides and studying their conformational flexibility. Herein results from molecular dynamics simulations of the protein-ligand complex of the aspartyl protease rhizopuspepsin and a polypeptide inhibitor are reported. The dihedral conformational space sampling for the linear peptide inhibitor in situ was found to be increased in the mass-weighted simulation as in other molecular systems previously studied. More significantly, the physical space of the enzyme binding pocket was also sampled efficiently in the simulations and multiple binding sites were identified for the inhibitor. These results suggest that it may be possible now to study, by computer simulations, the putative initial enzyme-inhibitor complex suggested experimentally from the time-dependent kinetics of enzyme inhibition by slow-binding inhibitors (Morrison, J. F., and C. T. Walsh. 1988. Adv. Enzymol. 61:201), and/or conformational substates in protein-ligand complexes suggested in the study of reassociation dynamics of myoglobin and carbon monoxide following photolysis (Austin, R. H., K. W. Beeson, L. Eisenstein, H. Frauenfelder, and I. C. Gunsalus. 1975. Biochemistry. 14:5355). Moreover, the intermediate binding steps and the molecular flexibility of the inhibitor shown in the MWMD simulation may have crucial roles in the ligand binding process.  相似文献   

15.
The effect of methylation of histidine-57 of alpha-chymotrypsin with Streptomyces subtilisin inhibitor was examined. Methylchymotrypsin was isolated by affinity chromatography on inhibitor-Sepharose, and the interaction of this inactive enzyme with inhibitor was quantitatively analyzed by two different methods: the spectrophotometric titration of difference spectrum resulted in the complex formation and the application of competitive enzyme assay by using substrates of large Km values. The former method gave values of 8.6 . 10(-6) M as dissociation constant (Kd) of methylchymotrypsin . inhibitor complex and 0.91 as the number of binding sites (n) per inhibitor monomer, both of which were almost equivalent to those for native enzyme . inhibitor complex. By the latter novel method, values of 7.9 . 10(-6) M and 1.08 were obtained for Kd and n, respectively, for interaction of inhibitor with alpha-chymotrypsin, and 8 . 10(-6) M as Kd for methylchymotrypsin . inhibitor complex. These results indicate that methylation of histidine-57 of active site in alpha-chymotrypsin molecule does not affect essentially the binding ability to inhibitor and the modified enzyme binds stoichiometrically to inhibitor, as the native enzyme does, with a molar ratio of 1:1 per inhibitor monomer.  相似文献   

16.
Transition state analogs pepstatin methylester (PME) and L685458 have been shown to inhibit gamma-secretase non-competitively (Tian, G., Sobotka-Briner, C., Zysk, J., Liu, X., Birr, C., Sylvester, M. A., Edwards, P. D., Scott, C. W., and Greenberg, B. D. (2002) J. Biol. Chem. 277, 31499-31505). This unusual kinetics suggests physical separation of the sites for substrate binding and catalysis with binding of the transition state analogs to the catalytic site and not to the substrate binding site. Methods of inhibitor cross-competition kinetics and competition ligand binding were utilized to address whether non-transition state small molecule inhibitors, which also display non-competitive inhibition of gamma-secretase, inhibit the enzyme by binding to the catalytic site as well. Inhibitor cross-competition kinetics indicated competitive binding between the transition state analogs PME and L685458 and between small molecules arylsulfonamides and benzodiazepines, but non-competitive binding between the transition state analogs and the small molecule inhibitors. These results were indicative of two inhibitor binding sites, one for transition state analogs and the other for non-transition state small molecule inhibitors. The presence of two inhibitor binding sites for two different classes of inhibitors was corroborated by results from competition ligand binding using [3H]L685458 as the radioligand. Although L685458 and PME displaced the radioligand at the same concentrations as for enzyme inhibition, arylsulfonamides and benzodiazepines did not displace the radioligand at their Ki values, a result consistent with the presence of two inhibitor binding sites. These findings provide useful insights into the catalytic and regulatory mechanisms of gamma-secretase that may facilitate the design of novel gamma-secretase inhibitors.  相似文献   

17.
Gamma-secretase is one of the critical enzymes required for the generation of amyloid-beta peptides from the beta-amyloid precursor protein. Because amyloid-beta peptides are generally accepted to play a key role in Alzheimer disease, gamma-secretase inhibition holds the promise for a disease-modifying therapy for this neurodegenerative condition. Although recent progress has enhanced the understanding of the biology and composition of the gamma-secretase enzyme complex, less information is available on the actual interaction of various inhibitor classes with the enzyme. Here we show that the two principal classes of inhibitor described in the scientific and patent literature, aspartyl protease transition state analogue and small molecule non-transition state inhibitors, display fundamental differences in the way they interact with the enzyme. Taking advantage of a gamma-secretase enzyme overexpressing cellular system and different radiolabeled gamma-secretase inhibitors, we observed that the maximal binding of non-transition state gamma-secretase inhibitors accounts only for half the number of catalytic sites of the recombinant enzyme complex. This characteristic stoichiometry can be best accommodated with a model whereby the non-transition state inhibitors bind to a unique site at the interface of a dimeric enzyme. Subsequent competition studies confirm that this site appears to be targeted by the main classes of small molecule gamma-secretase inhibitor. In contrast, the non-steroidal anti-inflammatory drug gamma-secretase modulator sulindac sulfide displayed noncompetitive antagonism for all types of inhibitor. This finding suggests that non-steroidal anti-inflammatory drug-type gamma-secretase modulators target an alternative site on the enzyme, thereby changing the conformation of the binding sites for gamma-secretase inhibitors.  相似文献   

18.
The crystal structure of the class-B beta-lactamase, BlaB, from the pathogenic bacterium, Chryseobacterium meningosepticum, in complex with the inhibitor, d-captopril, has been solved at 1.5-A resolution. The enzyme has the typical alphabeta/betaalpha metallo-beta-lactamase fold and the characteristic two metal binding sites of members of the subclass B1, in which two Zn2+ ions were identified. d-Captopril, a diastereoisomer of the commercial drug, captopril, acts as an inhibitor by displacing the catalytic hydroxyl ion required for antibiotic hydrolysis and intercalating its sulfhydryl group between the two Zn2+ ions. Interestingly, d-captopril is located on one side of the active site cleft. The x-ray structure of the complex of the closely related enzyme, IMP-1, with a mercaptocarboxylate inhibitor, which also contains a sulfhydryl group bound to the two Zn2+ ions, shows the ligand to be located on the opposite side of the active site cleft. A molecule generated by fusion of these two inhibitors would cover the entire cleft, suggesting an interesting approach to the design of highly specific inhibitors.  相似文献   

19.
The ox heart mitochondrial inhibitor protein may be iodinated with up to 0.8 mol 125I per mol inhibitor with no loss of inhibitory activity, with no change in binding affinity to submitochondrial particles, and without alteration in the response of membrane-bound inhibitor to energisation. Tryptic peptide maps reveal a single labelled peptide, consistent with modification of the single tyrosine residue of the protein. A single type of high-affinity binding site (Kd=96 . 10 (-9)M) for the inhibitor protein has been measured in submitochondrial particles. The concentration of this site is proportional to the amount of membrane-bound F1, and there appears to be one such site per F1 molecule. The ATp hydrolytic activity of submitochondrial particles is inversely proportional to the occupancy of the high-affinity binding site for the inhibitor protein. No evidence is found for a non-inhibitory binding site on the membrane or on other mitochondrial proteins. In intact mitochondria from bovine heart, the inhibitor protein is present in an approx. 1:1 ratio with F1. Submitochondrial particles prepared by sonication of these mitochondria with MgATP contain about 0.75 mol inhibitor protein per mol F1, and show about 25% of the ATPase activity of inhibitor-free submitochondrial particles. Additional inhibitor protein can be bound to these particles to a level of 0.2 mol/mol F1, with consequent loss of ATPase activity. If MgATP is omitted from the medium, or inhibitors of ATP hydrolysis are present, the rate of combination between F1 and its inhibitor protein is very much reduced. The equilibrium level of binding is, however, unaltered. These results suggest the presence of a single, high-affinity, inhibitory binding site for inhibitor protein on membrane-bound F1. The energisation of coupled submitochondrial particles by succinate oxidation or by ATP hydrolysis results in both the dissociation of inhibitor protein into solution, and the activation of ATP hydrolysis. At least 80% of the membrane-bound F1-inhibitor complex responds to this energisation by participating in a new equilibrium between bound and free inhibitor protein. This finding suggests that a delocalised energy pool is important in promoting inhibitor protein release from F1. Dissipation of the electrochemical gradient by uncouplers, or the binding of oligomycin or efrapetin effectively blocks energised release of the inhibitor protein. Conversely, the addition of aurovertin or adenosine 5'--[beta, lambda--imido]triphosphate enhances energy-driven release. The mode of action of various inhibitors on binding and energised release of the protein inhibitor is discussed.  相似文献   

20.
Determination of the number of ubiquinone- and inhibitor-binding sites in the mitochondrial complex I (NADH:ubiquinone oxidoreductase) is a controversial question with a direct implication for elaborating a suitable model to explain the bioenergetic mechanism of this complicated enzyme. We have used combinations of both selective inhibitors and common ubiquinone-like substrates to demonstrate the multiplicity of the reaction centers in the complex I in contrast with competition studies that have suggested the existence of a unique binding site for ubiquinone. Our results provide new evidence for the existence of at least two freely exchangeable ubiquinone-binding sites with different specificity for substrates, as well as for a different kinetic interaction of inhibitors with the 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