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 37 毫秒
1.
After short preincubations with N-[(3)H]methylscopolamine ([(3)H]NMS) or R(-)-[(3)H]quinuclidinyl benzilate ([(3)H]QNB), radioligand dissociation from muscarinic M(1) receptors in Chinese hamster ovary cell membranes was fast, monoexponential, and independent of the concentration of unlabeled NMS or QNB added to reveal dissociation. After long preincubations, the dissociation was slow, not monoexponential, and inversely related to the concentration of the unlabeled ligand. Apparently, the unlabeled ligand becomes able to associate with the receptor simultaneously with the already bound radioligand if the preincubation lasts for a long period, and to hinder radioligand dissociation. When the membranes were preincubated with [(3)H]NMS and then exposed to benzilylcholine mustard (covalently binding specific ligand), [(3)H]NMS dissociation was blocked in wild-type receptors, but not in mutated (D99N) M(1) receptors. Covalently binding [(3)H]propylbenzilylcholine mustard detected substantially more binding sites than [(3)H]NMS. The observations support a model in which the receptor binding domain has two tandemly arranged subsites for classical ligands, a peripheral one and a central one. Ligands bind to the peripheral subsite first (binding with lower affinity) and translocate to the central subsite (binding with higher affinity). The peripheral subsite of M(1) receptors may include Asp-99. Experimental data on [(3)H]NMS and [(3)H]QNB association and dissociation perfectly agree with the predictions of the tandem two-site model.  相似文献   

2.
The first report of slow-tight inhibition of xylanase by a bifunctional inhibitor alkalo-thermophilic Bacillus inhibitor (ATBI), from an extremophilic Bacillus sp. is described. ATBI inhibits aspartic protease (Dash, C., and Rao, M. (2001) J. Biol. Chem., 276, 2487-2493) and xylanase (Xyl I) from a Thermomonospora sp. The steady-state kinetics revealed time-dependent competitive inhibition of Xyl I by ATBI, consistent with two-step inhibition mechanism. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI*, and the dissociation of EI* were 13 +/- 1 x 10(-6) s(-1) and 5 +/- 0.5 x 10(-8) s(-1), respectively. The K(i) value for the formation of EI complex was 2.5 +/- 0.5 microm, whereas the overall inhibition constant K(i)* was 7 +/- 1 nm. The conformational changes induced in Xyl I by ATBI were monitored by fluorescence spectroscopy and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. ATBI binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde (OPTA)-labeled Xyl I. Our results revealed that the inactivation of Xyl I is due to the disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis and a model depicting the probable interaction between ATBI or OPTA with Xyl I has been proposed.  相似文献   

3.
This is the first report that describes the inhibition mechanism of xylanase from Thermomonospora sp. by pepstatin A, a specific inhibitor toward aspartic proteases. The kinetic analysis revealed competitive inhibition of xylanase by pepstatin A with an IC50 value 3.6 +/- 0.5 microm. The progress curves were time-depended, consistent with a two-step slow tight binding inhibition. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI* and the dissociation of EI* were 15 +/- 1 x 10(-5) and 3.0 +/- 1 x 10(-8) s(-1), respectively. The Ki value for the formation of EI complex was 1.5 +/- 0.5 microm, whereas the overall inhibition constant Ki* was 28.0 +/- 1 nm. The conformational changes induced in Xyl I by pepstatin A were monitored by fluorescence spectroscopy, and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. Pepstatin A binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde-labeled xylanase. Our results revealed that the inactivation of xylanase is due to the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis, and a model depicting the probable interaction between pepstatin A with xylanase has been proposed.  相似文献   

4.
Lin G  Liao WC  Ku ZH 《The protein journal》2005,24(4):201-207
The pre-steady states of Pseudomonas species lipase inhibitions by p-nitrophenyl-N-substituted carbamates (1-6) are composed of two steps: (1) formation of the non-covalent enzyme-inhibitor complex (E:I) from the inhibitor and the enzyme and (2) formation of the tetrahedral enzyme-inhibitor adduct (E-I) from the E:I complex. From a stopped-flow apparatus, the dissociation constant for the E:I complex, KS, and the rate constant for formation of the tetrahedral E-I adduct from the E:I complex, k2 are obtained from the non-linear least-squares of curve fittings of first-order rate constant (k(obs)) versus inhibition concentration ([I]) plot against k(obs)=k2+k2[I]/(KS+[I]). Values of pKS, and log k2 are linearly correlated with the sigma* values with the rho* values of -2.0 and 0.36, respectively. Therefore, the E:I complexes are more positive charges than the inhibitors due to the rho* value of -2.0. The tetrahedral E-I adducts on the other hand are more negative charges than the E:I complexes due to the rho* value of 0.36. Formation of the E:I complex from the inhibitor and the enzyme are further divided into two steps: (1) the pre-equilibrium protonation of the inhibitor and (2) formation of the E:I complex from the protonated inhibitor and the enzyme.  相似文献   

5.
Gonadotropin-releasing hormone (GnRH) binding sites have been characterized in the fully mature common carp ovary, using an analog of salmon GnRH ([D-Arg6,Trp7,Leu8,Pro9-NEt]-GnRH; sGnRH-A) as a labeled ligand. Binding of sGnRH-A to carp follicular membrane preparation was found to be time-, temperature-, and pH-dependent. Optimal binding was achieved after 40 min of incubation at 4 degrees C at pH 7.6; binding was found to be unstable at room temperature. Binding of radioligand was a function of tissue concentration, with a linear correlation over the range of 8.0-40.0 micrograms membrane protein per tube. Incubation of membrane preparations with increasing levels of [125I]sGnRH-A revealed saturable binding at radioligand concentrations greater than 400 nM. The binding of [125I]sGnRH-A to the carp ovary was also found to be reversible; addition of unlabeled sGnRH-A (10(-6) M) after reaching equilibrium resulted in complete dissociation of [125I]sGnRH-A within 30 min, and the log dissociation plot indicated the existence of a single class of binding sites. Addition of unlabeled sGnRH-A displaced the bound [125I]sGnRH-A in a dose-related manner. Hill plot as well as Scatchard analysis suggested the presence of one class of high affinity GnRH binding sites. Bound [125I]sGnRH-A was also found to be displaceable by other GnRH peptides, including sGnRH ([Trp7,Leu8]-GnRH), cGnRH-II ([His5,Trp7,Tyr8]-GnRH) and a GnRH antagonist ([D-pGlu1,D-Phe2,D-PTrp3,6]-GnRH; GnRH-ANT) in a parallel fashion, indicating that these peptides bind to the same class of binding sites.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The kinetics of slow onset inhibition of Proteinase K by a proteinaceous alkaline protease inhibitor (API) from a Streptomyces sp. is presented. The kinetic analysis revealed competitive inhibition of Proteinase K by API with an IC50 value 5.5 +/- 0.5 x 10-5 m. The progress curves were time-dependent, consistent with a two-step slow tight binding inhibition. The first step involved a rapid equilibrium for formation of reversible enzyme-inhibitor complex (EI) with a Ki value 5.2 +/- 0.6 x 10-6 m. The EI complex isomerized to a stable complex (EI*) in the second step because of inhibitor-induced conformational changes, with a rate constant k5 (9.2 +/- 1 x 10-3 s-1). The rate of dissociation of EI* (k6) was slower (4.5 +/- 0.5 x 10-5 s-1) indicating the tight binding nature of the inhibitor. The overall inhibition constant Ki* for two-step inhibition of Proteinase K by API was 2.5 +/- 0.3 x 10-7 m. Time-dependent dissociation of EI* revealed that the complex failed to dissociate after a time point and formed a conformationally altered, irreversible complex EI**. These conformational states of enzyme-inhibitor complexes were characterized by fluorescence spectroscopy. Tryptophanyl fluorescence of Proteinase K was quenched as a function of API concentration without any shift in the emission maximum indicating a subtle conformational change in the enzyme, which is correlated to the isomerization of EI to EI*. Time-dependent shift in the emission maxima of EI* revealed the induction of gross conformational changes, which can be correlated to the irreversible conformationally locked EI** complex. API binds to the active site of the enzyme as demonstrated by the abolished fluorescence of 5-iodoacetamidofluorescein-labeled Proteinase K. The chemoaffinity labeling experiments lead us to hypothesize that the inactivation of Proteinase K is because of the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the catalytic triad and other residues involved in catalysis.  相似文献   

7.
Previous studies [Dautry-Varsat, A., Cohen, G. N., & Stadtman, E.R. (1979) J. Biol. Chem. 254, 3124-3128; Lei, M., Aebi, U., Heidner, E. G., & Eisenberg, D. (1979) J. Biol. Chem. 254, 3129-3134] have shown that Escherichia coli glutamine synthetase (GS) can be cleaved by proteases to form a limited digestion species called nicked glutamine synthetase (GS). The present study gives the amino acid sequence of the protease-sensitive region of glutamine synthetase. The present study also shows that GS is enzymatically active, but this activity is low compared to the activity of GS. The apparent Michaelis constant value for glutamate was 90 mM for GS as compared to 3 mM for GS, while the Michaelis constant values for ATP were similar for GS and GS*. The dissociation constant values for ATP, as determined by intrinsic fluorescence measurements, were similar for GS and GS*. Glutamate decreased the dissociation constant value of ATP for GS because of synergism between the two binding sites; glutamate did not decrease the dissociation constant value of ATP for GS*. The glutamate analogue methionine sulfoximine bound very tightly to GS and inactivated the enzyme in the presence of ATP. Methionine sulfoximine did not appear to bind to GS* and did not inactivate GS* in the presence of ATP. The ATP analogue 5'-[p-(fluorosulfonyl)benzoyl]adenosine bound to GS and inactivated the enzyme by forming a covalent bond with it. Glutamate accelerated this inactivation because of the synergism between the ATP and glutamate binding sites of GS.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Lavendustin-A was reported to be a potent tyrosine kinase inhibitor of the epidermal growth factor (EGF) receptor (Onoda, T., Iinuma, H., Sasaki, Y., Hamada, M., Isshibi, K., Naganawa, H., Takeuchi, T., Tatsuta, K., and Umezawa, K. (1989) J. Nat. Prod. 52, 1252-1257). Its inhibition kinetics was studied in detail using the baculovirus-expressed recombinant intracellular domain of the EGF receptor (EGFR-IC). Lavendustin-A (RG 14355) is a slow and tight binding inhibitor of the receptor tyrosine kinase. The pre-steady state kinetic analysis demonstrates that the inhibition corresponds to a two-step mechanism in which an initial enzyme-inhibitor complex (EI) is rapidly formed followed by a slow isomerization step to form a tight complex (EI*). The dissociation constant for the initial rapid forming complex is 370 nM, whereas the overall dissociation constant is estimated to be less than or equal to 1 nM. The difference between the two values is due to the tight binding nature of the inhibitor to the enzyme in EI*. The kinetic analysis using a preincubation protocol to pre-equilibrate the enzyme with the inhibitor in the presence of one substrate showed that Lavendustin-A is a hyperbolic mixed-type inhibitor with respect to both ATP and the peptide substrate, with a major effect on the binding affinities for both substrates. An analogue of Lavendustin-A (RG 14467) showed similar inhibition kinetics to that of Lavendustin-A. The results of the pre-steady state analysis are also consistent with the proposed two-step mechanism. The dissociation constant for the initial fast forming complex in this case is 3.4 microM, whereas the overall dissociation constant is estimated to be less than or equal to 30 nM. It is a partial (hyperbolic) competitive inhibitor with respect to ATP. Its inhibition is reduced to different extents by different peptide substrates, when the peptide is added to the enzyme simultaneously with the inhibitor. When studied with the least protective peptide, K1 (a peptide containing the major autophosphorylation site of the EGF receptor), RG 14467 acts as a hyperbolic noncompetitive inhibitor with respect to the peptide.  相似文献   

9.
The present article reports a low molecular weight aspartic protease inhibitor, API, from a newly isolated thermo-tolerant Bacillus licheniformis. The inhibitor was purified to homogeneity as shown by rp-HPLC and SDS-PAGE. API is found to be stable over a broad pH range of 2-11 and at temperature 90 degrees C for 2 1/2h. It has a Mr (relative molecular mass) of 1363 Da as shown by MALDI-TOF spectra and 1358 Da as analyzed by SDS-PAGE .The amino acid analysis of the peptide shows the presence of 12 amino acid residues having Mr of 1425 Da. The secondary structure of API as analyzed by the CD spectra showed 7% alpha-helix, 49% beta-sheet and 44% aperiodic structure. The Kinetic studies of Pepsin-API interactions reveal that API is a slow-tight binding competitive inhibitor with the IC(50) and Ki values 4.0 nM and (3.83 nM-5.31 nM) respectively. The overall inhibition constant Ki* value is 0.107+/-0.015 nM. The progress curves are time-dependent and consistent with slow-tight binding inhibition: E+I -->/<-- (k(4), k(5)) EI -->/<-- (k(6), k(7)) EI*. Rate constant k(6)=2.73+/-0.32 s(-1) reveals a fast isomerization of enzyme-inhibitor complex and very slow dissociation as proved by k(7)=0.068+/-0.009 s(-1). The Rate constants from the intrinsic tryptophanyl fluorescence data is in agreement with those obtained from the kinetic analysis; therefore, the induced conformational changes were correlated to the isomerization of EI to EI*.  相似文献   

10.
alpha(1)-Antichymotrypsin is a member of the serine proteinase inhibitor, or serpin, family that typically forms very long-lived, enzymatically inactive 1:1 complexes (denoted E*I*) with its target proteinases. Serpins share a conserved tertiary structure, in which an exposed region of amino acid residues (called the reactive center loop or RCL) acts as bait for a target proteinase. Within E*I*, the two proteins are linked covalently as a result of nucleophilic attack by Ser(195) of the serine proteinase on the P1 residue within the RCL of the serpin. This species is formally similar to the acyl enzyme species normally seen as an intermediate in serpin proteinase catalysis. However, its subsequent hydrolysis is extremely slow as a result of structural changes within the enzyme leading to distortion of the active site. There is at present an ongoing debate concerning the structure of the E*I* complex; in particular, as to whether the enzyme, bound to P1, maintains its original position at the top of the serpin molecule or instead translocates across the entire length of the serpin, with concomitant insertion of RCL residues P1-P14 within beta-sheet A and a large separation of the enzyme and RCL residue P1'. We report time-resolved fluorescence energy transfer and rapid mixing/quench studies that support the former model. Our results indicate that the distance between residue P1' in alpha(1)-antichymotrypsin and the amino terminus of chymotrypsin actually decreases on conversion of the encounter complex E.I to E*I*. These results led us to formulate a comprehensive mechanism that accounted both for our results and for those of others supporting the two different E*I* structures. In this mechanism, partial insertion of the RCL, with no large perturbation of the P1' enzyme distance, is followed by covalent acyl enzyme formation. Full insertion can subsequently take place, in a reversible fashion, with the position of equilibrium between the partially and fully inserted complexes depending on the particular serpin-proteinase pair under consideration.  相似文献   

11.
In the intravesicle scooting mode of interfacial catalysis, the interfacial complex E*S is formed by the interaction of the membrane bound phospholipase A2 (E*) with the substrate monomer (S) in the interface. In the presence of nonhydrolyzable substrate analogs (I) the kinetics of interfacial catalysis is modified. If phospholipase A2 is added to a mixture of the vesicles of L-DMPMe ester and of DTPMe ether or D-DMPMe ester, the extent of hydrolysis, A, decreases and the interfacial scooting rate constant, ki, remains unchanged. On the other hand, when the enzyme is added to the vesicles prepared from premixed L-DMPMe ester with D-DMPMe ester or L-DTPMe ether, ki decreases but A remains constant. Qualitatively, these results are in excellent accord with the Scheme I for interfacial catalysis. However, a quantitative departure has been noted, which suggests that the interfacial dissociation constant for E*S is larger than that for E*I. These results are interpreted to suggest that the catalytic rate constant for decomposition of E*S to E* + P is larger than the rate constant for decomposition of E*S to E* + S. Broader implications of the scooting mode of interfacial catalysis are discussed.  相似文献   

12.
Studies of CO ligand binding revealed that two protein states with different ligand affinities exist in the protoglobin from Methanosarcina acetivorans (in MaPgb*, residue Cys(E20)101 was mutated to Ser). The switch between the two states occurs upon the ligation of MaPgb*. In this work, site-directed mutagenesis was used to explore the role of selected amino acids in ligand sensing and stabilization and in affecting the equilibrium between the “more reactive” and “less reactive” conformational states of MaPgb*. A combination of experimental data obtained from electronic and resonance Raman absorption spectra, CO ligand-binding kinetics, and X-ray crystallography was employed. Three amino acids were assigned a critical role: Trp(60)B9, Tyr(61)B10, and Phe(93)E11. Trp(60)B9 and Tyr(61)B10 are involved in ligand stabilization in the distal heme pocket; the strength of their interaction was reflected by the spectra of the CO-ligated MaPgb* and by the CO dissociation rate constants. In contrast, Phe(93)E11 is a key player in sensing the heme-bound ligand and promotes the rotation of the Trp(60)B9 side chain, thus favoring ligand stabilization. Although the structural bases of the fast CO binding rate constant of MaPgb* are still unclear, Trp(60)B9, Tyr(61)B10, and Phe(93)E11 play a role in regulating heme/ligand affinity.  相似文献   

13.
Serpins form enzymatically inactive covalent complexes (designated E*I*) with their target proteinases, corresponding most likely to the acyl enzyme that resembles the normal intermediate in substrate turnover. Formation of E*I* involves large changes in the conformation of the reactive center loop (residues P17 to P9') and of the serpin molecule in general. The "hinge" region of the reactive center loop, including residues P10-P14, shows facile movement in and out of beta-sheet A, and this movement appears to be crucial in determining whether E*I* is formed (the inhibitor pathway) or whether I is rapidly hydrolyzed to I* (the substrate pathway). Here, we report stopped-flow and rapid quench studies investigating the pH dependence of the conversion of the alpha1-antichymotrypsin.alpha-chymotrypsin encounter complex, E.I, to E*I*. These studies utilize fluorescent derivatives of cysteine variants of alpha1-antichymotrypsin at the P11 and P13 residues. Our results demonstrate three identifiable intermediates, EIa, EIb, and EIc, between E.I and E*I* and permit informed speculation regarding the nature of these intermediates. Partitioning between inhibitor and substrate pathways occurs late in the process of E*I* formation, most likely from a species occurring between EIc and E*I*.  相似文献   

14.
Modification of transducin (T) with iodoacetic acid (IAA) inhibited its light-dependent guanine nucleotide-binding activity. Approximately 1 mol of [(3)H]IAA was incorporated per mole of T. Cys(347), located on the alpha-subunit of T (T(alpha)), was identified as the major labeled residue in the [(3)H]IAA-modified holoenzyme. In contrast, Cys(135) and Cys(347) were modified with [(3)H]IAA in the isolated T(alpha). IAA-modified T was able to bind tightly to photoexcited rhodopsin (R*), but GTP did not promote the dissociation of the complex between alkylated T and R*. In addition, R* protected against the inhibition of T by IAA. A comparable inactivation of T and analogous interactions between T and R* were observed when 2-nitro 5-thiocyanobenzoic acid (NTCBA) was used as the modifying reagent (J. O. Ortiz and J. Bubis, 2001, Effects of differential sulfhydryl group-specific labeling on the rhodopsin and guanine nucleotide binding activities of transducin, Arch. Biochem. Biophys. 387, 233-242). However, while carboxymethylated T was capable of liberating GDP in the presence of R*, NTCBA-modified T was unable to release the guanine nucleotide diphosphate upon incubation with the photoactivated receptor. Thus, IAA-labeling stabilized a T:R* complex intermediate carrying the empty nucleotide pocket conformation of T. On the other hand, NTCBA-modified T seemed to be "locked" in the GDP-bound state of T, even in the presence of R*.  相似文献   

15.
Site-directed mutagenesis was used to probe the structural and functional roles of two highly conserved residues, Tyr-52 and Tyr-73, in interfacial catalysis by bovine pancreatic phospholipase A2 (PLA2, overproduced in Escherichia coli). According to crystal structures, the side chains of these two active site residues form H-bonds with the carboxylate of the catalytic residue Asp-99. Replacement of either or both Tyr residues by Phe resulted in only very small changes in catalytic rates, which suggests that the hydrogen bonds are not essential for catalysis by PLA2. Substitution of either Tyr residue by nonaromatic amino acids resulted in substantial decreases in the apparent kcat toward 1,2-dioctanoyl-sn-glycero-3-phosphocholine (DC8PC) micelles and the v(o) (turnover number at maximal substrate concentration, i.e., mole fraction = 1) toward 1,2-dimyristoyl-sn-glycero-3-phosphomethanol (DC14PM) vesicles in scooting mode kinetics [Berg, O. G., Yu, B.-Z., Rogers, J., & Jain, M. K. (1991) Biochemistry 30, 7283-7297]. The Y52V mutant was further analyzed in detail by scooting mode kinetics: the E to E* equilibrium was examined by fluorescence; the dissociation constants of E*S, E*P, and E*I (KS*, KP*, and KI*, respectively) in the presence of Ca2+ were measured by protection of histidine-48 modification and by difference UV spectroscopy; the Michaelis constant KM* was calculated from initial rates of hydrolysis in the absence and presence of competitive inhibitors; and the turnover number under saturating conditions (kcat, which is a theoretical value since the enzyme may not be saturated at the interface) was calculated from the vo and KM* values. The results indicated little perturbation in the interfacial binding step (E to E*) but ca. 10-fold increases in KS*, KP*, KI*, and KM* and a less than 10-fold decrease in kcat. Such changes in the function of Y52V are not due to global conformational changes since the proton NMR properties of Y52V closely resemble those of wild-type PLA2; instead, it is likely to be caused by perturbed enzyme-substrate interactions at the active site. Tyr-73 appears to play an important structural role. The conformational stability of all Tyr-73 mutants decreased by 4-5 kcal/mol relative to that of the wild-type PLA2. The proton NMR properties of Y73A suggested significant conformational changes and substantially increased conformational flexibility. These detailed structural and functional analyses represent a major advancement in the structure-function study of an enzyme involved in interfacial catalysis.  相似文献   

16.
J Myung  W P Jencks 《FEBS letters》1991,278(1):35-37
The E-E* model for calcium pumping by the CaATPase of sarcoplasmic reticulum includes two distinct conformational states of the enzyme, E and E*. Exterior Ca2+ binds only to E and interior Ca2+ binds only to E*. Therefore, it is expected that there will be competition between the binding of calcium to the unphosphorylated enzyme from the two sides of the membrane. The equilibrium concentration of cECa2, the enzyme with Ca2+ bound at the exterior site, was measured at different Ca2+ concentrations with empty sarcoplasmic reticulum vesicles (SRV) and with SRV loaded with 40 mM Ca2+ by reaction with 0.5 mM [gamma-32P]ATP plus 20 mM EGTA for 13 ms (100 mM KCl, 5 mM MgSO4, 40 mM Mops/KOH, pH 7.0, 25 degrees C). The sigmoidal dependence on free exterior calcium concentration of the concentration of cECa2, measured as [32P]phosphoenzyme, is identical with empty and loaded SRV, within experimental error. The value of K0.5 is 2.8 microM, and the Hill coefficient is 2. This result shows that there is no competition between binding of Ca2+ to the outside and the inside of the membrane. This is consistent with a model in which the vectorial specificity for calcium binding is controlled by the chemical state of the enzyme, rather than a simple conformational change. It is concluded that there are not two interconverting forms of the free enzyme, E and E*, instead the vectorial specificity for binding and dissociation of Ca2+ is determined by the state of phosphorylation of the CaATPase.  相似文献   

17.
18.
Abstract

Analysis of (-)[125]iodo-N6-(4-hydroxyphenylisopropyl)-adenosine ([125I]HPIA) binding to purified sarcolemmal preparations of guinea pig and bovine hearts revealed two classes of binding sites when unlabeled iodo-HPIA (100 μmol/1) was used as non-specific binding marker. In the presence of 1 mmol/1 theophylline, however, only the high affinity component was detected. Adenosine receptor agonists caused biphasic displacement of [125I]HPIA binding, with a high affinity potency rank order typical of interaction with A1-adenosine receptors. Biphasic competition curves were also observed with 8-phenyltheophylline and isobutylmethylxanthine, whereas the theophylline curve was monophasic up to 1 mmol/1. In brain membranes, specific binding of [125I]HPIA as well as of [3H]PIA was further reduced when unlabeled iodo-HPIA replaces theophylline as the non-specific binding marker. These results suggest the presence of two [125I]HPIA binding sites on cardiac sarcolemma and brain membranes, but receptor function can only be ascribed to the high affinity sites. The low affinity site probably represents an artefact, which is often observed when non-specific binding is defined with the unlabeled counterpart or a structurally related ligand of the radioligand used.  相似文献   

19.
The rate of dissociation of labeled estradiol from [3H] estradiol-8-9 S receptor complexes ([3H]E2-8-9 S ER) molybdate-stabilized was determined in the presence of either an excess of unlabeled hormone ("chase") or of charcoal/dextran suspension ("stripping"). Biphasic dissociation of the hormone was observed in both cases, but the fraction of the fast-dissociating component was dramatically reduced (5% instead of 60%) when stripping was used. As the dissociation patterns were independent of the degree of saturation of the receptor, the results do not favor the possibility of cooperative effects between binding sites in the 8-9 S ER. After pretreatment of cytosol by charcoal at 28 degrees C for 15 min, the dissociation studied by chase displayed only the slowly dissociating component (t1/2 approximately 65 min). This effect was dependent on temperature and influenced by the ligand bound to 8-9 S ER, being pronounced with estradiol (E2) and absent with [3H]4-hydroxytamoxifen. The slow-dissociating component obtained after charcoal treatment was reconverted to fast-dissociating state by adding dithiothreitol or by incubation with cytosol at 20 degrees C. The charcoal treatment did not change the sedimentation coefficient (approximately 9 S) and the Stokes radius (approximately 7 nm) of the [3H]E2-8-9 S ER, and the slow-dissociating form obtained did not bind to DNA-cellulose either in the presence or absence of molybdate ions. Thus there are likely small but functionally significant changes of structure in the 8-9 S ER which remain in a non-DNA-binding form, whereas the rate of estradiol dissociation is modified.  相似文献   

20.
To examine the existence of negative cooperativity among formyl peptide chemotaxis receptors, steady-state binding of f Met-Leu-[3H]Phe to viable rat neutrophils and their purified plasma membranes was measured and the data were subjected to statistical analysis and to computer curve fitting using the NONLIN computer program. Curvilinear, concave upward Scatchard plots were obtained. NONLIN and statistical analysis of the binding data indicated that a two-saturable-sites model was preferable to a one-saturable-site model and statistically valid by the F-test (P less than .010). In addition, Hill coefficients of 0.80 +/- 0.02 were obtained. Kinetic dissociation experiments using purified plasma membranes showed evidence of site-site interactions of the destabilizing type (negative cooperativity). Thus, unlabeled f Met-Leu-Phe accelerated the dissociation of f Met-Leu-[3H]Phe under conditions where no rebinding of radioligand occurred. The rate of dissociation of f Met-Leu-[3H]Phe from the plasma membranes was dependent on the fold excess of unlabeled f Met-Leu-Phe used in the dilution medium; at the highest concentration tested (10,000-fold excess), the dissociation rate was more than double the dissociation rate seen with dilution alone. In addition, occupancy-dependent affinity was ascertained directly by studying the effect of increasing fractional receptor saturation with labeled ligand on the dissociation rate of the receptor-bound labeled ligand. These data showed that the f Met-Leu-[3H]Phe dissociation rate was dependent on the degree of binding site occupancy over the entire biologically relevant range of formyl peptide concentrations. Furthermore, monitoring of the time course of dissociation of the receptor/f Met-Leu-[3H]Phe receptor/f Met-Leu-[3H]Phe complex as a function of receptor occupancy revealed that receptor affinity for f Met-Leu-Phe remained occupancy-dependent during the entire time of dissociation examined (up to 10 min). Finally, the average affinity profile of the equilibrium binding data demonstrated a 60% decrease in receptor affinity in changing from the high affinity to the low affinity conformation.  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