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.
J Secnik  Q Wang  C M Chang  J E Jentoft 《Biochemistry》1990,29(34):7991-7997
The structural and functional properties of the nucleocapsid (NC) protein of the avian myeloblastosis virus were examined by steady-state fluorescence and fluorescence anisotropy measurements of the complex between the NC and the extrinsic fluorophore 4,4'-bis(phenylamino)(1,1'-binaphthalene)-5,5'-disulfonic acid (bis-ANS). The intrinsic fluorescence of bis-ANS is enhanced many fold upon forming a complex with the NC. Between 2 and 10 molecules of bis-ANS bind strongly to the NC, with an overall Kd of less than 10(-6) M. The emission of bis-ANS in the complex can also be induced by excitation at 298 nm, indicating that energy is transferred from Trp 80, the sole tryptophan in the NC protein, to bis-ANS. The energy transferred between the Trp 80 and bis-ANS was analyzed to yield a calculated distance of separation between these fluorophores of 28 +/- 3 A; thus, Trp 80 is well removed from the nearest bound bis-ANS. The fluorescence emission of bis-ANS in the NC.bis-ANS complex is efficiently quenched by added salts and by poly(A), suggesting that salt (presumably anions), nucleic acid, and bis-ANS bind to the same, positively charged region on the NC protein. A site size of six nucleotides was determined for nucleic acid binding to the NC protein, with an estimated Kd of less than 10(-6) M. Salt (anion) binding is strong, but nonspecific, with a Kapp of 4 mM, raising the possibility that anion binding to the NC protein might regulate the interaction of the NC with viral RNA inside the host cell.  相似文献   

2.
Chloroplast coupling factor 1 (CF1) contains a high-affinity binding site for 8-anilino-1-napthalene sulphonate (ANS,Kd = 5-6 microM). The binding of ANS to the enzyme is associated with a fluorescence enhancement and a blue-shift in the emission spectrum. ANS only slightly inhibits ATP hydrolysis by CF1. Adenine nucleotides and inorganic phosphate induce a fast ANS fluorescence quenching of about 50% which is due to a decrease in the affinity of the enzyme for ANS (Kd increases from 6 microM to 22 microM) and in the fluorescence quantum yield of the bound probe (by 33%) but not in the number of ANS sites (n = 1). Conversely, Mg and Ca ions induce a fluorescence enhancement of bound ANS. Inactivation of the enzyme enhances ANS fluorescence, eliminates the response to adenine nucleotides and inorganic phosphate but increases the response to divalent metals. The affinity of latent CF1 for ADP (Kd = 12 microM) is considerably higher than for ATP (Kd = 95 microM) in buffer containing EDTA. The Kd for inorganic phosphate is 140 microM. Mg increases the apparent affinity for ATP (Kd = 28 microM) but not for ADP or Pi. Binding of ATP to the tight-sites does not inhibit the ADP or Pi-induced fluorescence quenching but decreases the affinity for ADP (Kd = 34 microM) and for inorganic phosphate (Kd = 320 microM). These results suggest that the ADP and phosphate binding sites are different but not independent from the tight sites. Activation of a Mg-specific ATPase in CF1 by octyl glucoside decreases the affinity for ADP and inorganic phosphate by about threefold but increases the affinity for ATP. ATPase activation of CF1 also increases the Ki for ADP inhibition of ATP hydrolysis. ATPase activation also influences the ANS responses to Ca and Mg. Ca-ATPase activation increases the fluorescence enhancement and the apparent affinity for Ca whereas Mg-ATPase activation specifically increases the Mg-induced fluorescence enhancement. The fluorescence of CF1-bound ANS is enhanced by Dio-9 and quenched by phloridzin, quercetin, Nbf-Cl and FITC. Nbf-Cl and FITC completely inhibit the ADP-induced fluorescence quenching whereas Dio-9 inhibits the Mg-induced fluorescence enhancement. ANS does not relieve the quercetin or phloridzin inhibition of ATP hydrolysis indicating that these inhibitors do not compete with ANS for a common binding site. ANS may be used, therefore, as a sensitive probe to detect conformational changes in CF1 in response to activation or inactivation and to binding of substrates and of inhibitors.  相似文献   

3.
G Musci  L J Berliner 《Biochemistry》1985,24(15):3852-3856
The binding of the fluorescent probe 4,4'-bis[1-(phenylamino)-8-naphthalenesulfonate] (bis-ANS) to bovine alpha-lactalbumin (alpha-LA) was investigated. A strong dependence of the Kd value with the bound calcium stoichiometry was found, with Kd values ranging from 6.2 +/- 0.4 to 64.6 +/- 5.9 microM for apo-alpha-LA and 1:1 Ca(II)-alpha-LA, respectively. A 350-fold enhancement of the bis-ANS emission was observed in the protein-bis-ANS complex, along with an approximately 33-nm blue shift. Both appeared to be related to the hydrophobicity of the binding site and were independent of the Ca(II) ion content. From the difference in bis-ANS affinity between apo-alpha-LA and Ca(II)-alpha-LA, we demonstrated that Zn(II) and Al(III) were able to "lock" the protein into a new "apo-like" conformation, which was similar to, but not identical with, the apo conformation. The protein could be interconverted between all three conformations in a Mn(II) titration. The first Mn(II) shifted the apoprotein to the Ca(II) conformation; at higher Mn(II) levels, binding to the second site shifted the protein toward the apo-like conformation. The same behavior was observed with calcium in large excess. The evidence supported a model for the mutually nonexclusive binding of metals both to site I ("calcium site") and to site II ("zinc site") simultaneously. The results suggest that alpha-lactalbumin possesses a hydrophobic surface that becomes somewhat less accessible upon 1:1 calcium binding in the absence of metals that also bind to the zinc site.  相似文献   

4.
The effect of aurovertin on the binding parameters of ADP and ATP to native F1 from beef heart mitochondria in the presence of EDTA has been explored. Three exchangeable sites per F1 were titrated by ADP and ATP in the absence or presence of aurovertin. Curvilinear Scatchard plots for the binding of both ADP and ATP were obtained in the absence of aurovertin, indicating one high affinity site (Kd for ADP = 0.6-0.8 microM; Kd for ATP = 0.3-0.5 microM) and two lower affinity sites (Kd for ADP = 8-10 microM; Kd for ATP = 7-10 microM). With a saturating concentration of aurovertin capable of filling the three beta subunits of F1, the curvilinearity of the Scatchard plots was decreased for ATP binding and abolished for ADP binding, indicating homogeneity of ADP binding sites in the F1-aurovertin complex (Kd for ADP = 2 microM). When only the high affinity aurovertin site was occupied, maximal enhancement of the fluorescence of the F1-aurovertin complex was attained with 1 mol of ADP bound per mol of F1 and maximal quenching for 1 mol of ATP bound per mol of F1. When the F1-aurovertin complex was incubated with [3H]ADP followed by [14C]ATP, full fluorescence quenching was attained when ATP had displaced the previously bound ADP. In the case of the isolated beta subunit, both ADP and ATP enhanced the fluorescence of the beta subunit-aurovertin complex. The Kd values for ADP and ATP in the presence of EDTA were 0.6 mM and 3.7 mM, respectively; MgCl2 decreased the Kd values to 0.1 mM for both ADP and ATP. It is postulated that native F1 possesses three equivalent interacting nucleotide binding sites and exists in two conformations which are in equilibrium and recognize either ATP (T conformation) or ADP (D conformation). The negative interactions between the nucleotide binding sites of F1 are strongest in the D conformation. Upon addition of aurovertin, the site-site cooperativity between the beta subunits of F1 is decreased or even abolished.  相似文献   

5.
J C Lee  L C Yeh  P M Horowitz 《Biochimie》1991,73(9):1245-1247
Binding studies of yeast 40S ribosome with bis (1,8-anilinonaphthalenesulfonate) (bis-ANS) revealed the binding of 3-4 molecules of bis-ANS per ribosome with a dissociation constant (Kd) of 1.45 microM. Binding of AUG to the 40S subunits resulted in a concentration-dependent decrease in the bis-ANS fluorescence without displacing all of the bound bis-ANS from the ribosomes. The residual bis-ANS fluorescence at saturation with AUG corresponds to about 3 molecules of bis-ANS per ribosome. Thus AUG displaces one of the bound bis-ANS molecules. The data suggest that AUG binds at a hydrophobic site on the yeast 40S subunit.  相似文献   

6.
The regulation of prostacyclin (PGI2) synthesis by cultured human umbilical vein endothelium (HUVEC) was investigated. HUVEC monolayer generation of PGI2 was monitored by RIA of 6-keto PGF1 alpha and dose-dependent increases observed with human alpha- and gamma-thrombins, histamine, or arachidonate. Alpha thrombin (10 nM) produced levels of 6-keto PGF1 alpha approximating responses with 1 microM gamma-thrombin, 5 microM arachidonate, or 10 microM histamine. Diisopropyl phosphorofluoridate-inactivated alpha-thrombin did not stimulate PGI2 release, demonstrating that catalytic activity was required for thrombin-stimulated PGI2 release. Sodium fluoride (NaF), at concentrations known to activate guanine nucleotide regulatory proteins (G proteins), directly stimulated HUVEC PGI2 synthesis in a dose-dependent and time-dependent manner (20 mM NaF, 4.4 +/- 0.5-fold increase at 10 min, 11.9 +/- 1.5-fold increase at 30 min). Neither alpha-thrombin nor NaF-stimulated PGI2 release was dependent upon the availability of extracellular Ca++). The hypothesis that G proteins are involved in agonist-stimulated PGI2 synthesis was further supported by studies using digitonin-permeabilized HUVEC monolayers challenged with another G protein activator, guanosine 5'-0-3-thiotrisphosphate (GTP gamma S), which effected significant dose-dependent increases in PGI2 synthesis compared with control levels of 6-keto PGF1 alpha. In contrast, the G-protein inhibitor GDP beta S, (guanosine 5'-0-2-thiodiphosphate), attenuated alpha-thrombin-mediated prostaglandin generation. Treatment of HUVEC monolayers with pertussis toxin (1 microgram/ml) did not inhibit the PGI2 synthesis stimulated by either alpha-thrombin, NaF, or histamine but catalyzed the ADP ribosylation of a 40 kDa membrane protein which cross-reacted with antisera against a synthetic peptide corresponding to an amino acid sequence common to the alpha-subunit of other G-proteins. Preincubation of HUVEC microsomal membranes with alpha-thrombin diminished pertussis toxin-catalyzed ADP ribosylation in a time-dependent manner. These data suggest that thrombin stimulation of PGI2 synthesis by HUVEC monolayers requires the catalytically functional enzyme and further suggests that the thrombin-occupied receptor is coupled to phospholipase activities by a pertussis toxin-insensitive guanine nucleotide regulatory protein in human endothelial cell membranes.  相似文献   

7.
We have used a fluorescence assay to measure the binding of Acanthamoeba profilin to monomeric Acanthamoeba and rabbit skeletal muscle actin labeled on cysteine-374 with pyrene iodoacetamide. The wavelengths of the pyrene excitation and emission maxima are constant at 346 and 386 nm, but the fluorescence is enhanced up to 50% by profilin. The higher fluorescence is largely due to higher absorbance in the presence of profilin. The fluorescence enhancement has a hyperbolic dependence on the concentration of profilin, suggesting a single class of binding sites. Linear Scatchard plots yield an estimate of the dissociation constant, Kd, of the complex of profilin with pyrenyl-actin. In low-ionic-strength buffers with 2 to 6 mM imidazole (pH 7.0) and 0.1 mM CaCl2 the Kd is 9 microM for both muscle and Acanthamoeba actin. In 50 mM KCl the Kd for the complex with Acanthamoeba actin is 16 microM, while the Kd for the complex with muscle actin is greater than 50 microM.  相似文献   

8.
The binding of substrates and the herbicide N-(phosphonomethyl)glycine (glyphosate) to enolpyruvoylshikimate-3-phosphate (EPSP) synthase was evaluated by stopped-flow and equilibrium fluorescence measurements. Changes in protein fluorescence were observed upon the binding of EPSP and upon the formation of the enzyme-shikimate 3-phosphate-glyphosate ternary complex; no change was seen with either shikimate 3-phosphate (S3P) or glyphosate alone. By fluorescence titrations, the dissociation constants were determined for the formation of the enzyme binary complexes with S3P (Kd,S = 7 +/- 1.2 microM) and EPSP (Kd,EPSP = 1 +/- 0.01 microM). The dissociation constant for S3P was determined by competition with EPSP or by measurements in the presence of a low glyphosate concentration. At saturating concentrations of S3P, glyphosate bound to the enzyme--S3P binary complex with a dissociation constant of 0.16 +/- 0.02 microM. Glyphosate did not bind significantly to free enzyme, so the binding is ordered with S3P binding first: (formula; see text) where S refers to S3P, G refers to glyphosate, and E.S.G. represents the complex with altered fluorescence. The kinetics of binding were measured by stopped-flow fluorescence methods. The rate of glyphosate binding to the enzyme--S3P complex was k2 = (7.8 +/- 0.2) X 10(5) M-1 s-1, from which we calculated the dissociation rate k-2 = 0.12 +/- 0.02 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Calmodulin (CaM) binding by turkey gizzard myosin light chain kinase (MLCK) causes subtle changes in the fluorescence emission and polarization excitation spectra of the enzyme. Fluorescence experiments using 9-anthroyl-choline (9AC), which competes with ATP in binding, demonstrate mutually stabilizing interactions between the CaM and ATP binding sites corresponding to delta G = -0.6 to -0.7 kcal/mol. Fluorescence titrations in the presence of 9AC or 5,5'-bis[8-(phenylamino)-1-naphthalenesulfonate] confirm the stoichiometry of 1 mol of CaM/MLCK. Phosphorylation of MLCK has no effect on either the protein fluorescence or the binding of ATP and 9AC. The dissociation constant for the MLCL-CaM complex is increased approximately 500-fold on phosphorylation. Values of Kd for the phosphorylated enzyme range from 0.5 to 1.1 microM in 0.2 N KCl, pH 7.3, 25 degrees C. We showed competition between MLCK and other CaM binding proteins and peptides by using both fluorescence and catalytic activity measurements. Competition for CaM occurs with ACTH, beta-endorphin, substance P, glucagon, poly(L-arginine), myelin basic protein, troponin I, and histone H2A. Phosphorylation of the last three proteins by the adenosine cyclic 3',5'-phosphate dependent protein kinase diminishes their ability to compete. Phosphorylation of MLCK by the protein kinase gives 0.95 +/- 0.04 and 2.2 +/- 0.4 mol of incorporated 32P in the presence and absence of CaM, respectively. These stoichiometries agree with those recently reported [Conti, M. A. & Adelstein, R. S. (1981) J. Biol. Chem. 256, 3178].  相似文献   

10.
Bothrojaracin, a 27-kDa C-type lectin from Bothrops jararaca venom, is a selective and potent thrombin inhibitor (K(d) = 0.6 nM) which interacts with the two thrombin anion-binding exosites (I and II) but not with its catalytic site. In the present study, we analyzed the allosteric effects produced in the catalytic site by bothrojaracin binding to thrombin exosites. Opposite effects were observed with alpha-thrombin, which possesses both exosites I and II, and with gamma-thrombin, which lacks exosite I. On the one hand, bothrojaracin altered both kinetic parameters K(m) and k(cat) of alpha-thrombin for small synthetic substrates, resulting in an increased efficiency of alpha-thrombin catalytic activity. This effect was similar to that produced by hirugen, a peptide based on the C-terminal hirudin sequence (residues 54-65) which interacts exclusively with exosite I. On the other hand, bothrojaracin decreased the amidolytic activity of gamma-thrombin toward chromogenic substrates, although this effect was observed with higher concentrations of bothrojaracin than those used with alpha-thrombin. In agreement with these observaions, bothrojaracin produced opposite effects on the fluorescence intensity of alpha- and gamma-thrombin derivatives labeled at the active site with fluorescein-Phe-Pro-Arg-chloromethylketone. These observations support the conclusion that bothrojaracin binding to thrombin produces two different structural changes in its active site, depending on whether it interacts exclusively with exosite II, as seen with gamma-thrombin, or with exosite I (or both I and II) as observed with alpha-thrombin. The ability of bothrojaracin to evoke distinct modifications in the thrombin catalytic site environment when interacting with exosites I and II make this molecule an interesting tool for the study of allosteric changes in the thrombin molecule.  相似文献   

11.
In a new strategy for labeling the active sites of serine proteinases with fluorescence probes (Bock, P. E. (1988) Biochemistry 27, 6633-6639), a thioester peptide chloromethyl ketone inhibitor is incorporated into the enzyme active center and used to produce a unique thiol group which provides a site for selective chemical modification with any one of many thiol-reactive fluorescence probes. This approach was developed to increase the opportunities for identifying fluorescent proteinase derivatives that act as reporters of binding interactions by allowing a large number of derivatives, representing a broad range of probe spectral properties, to be readily prepared. In the studies described here, the specificity of the labeling approach was evaluated quantitatively for the labeling of human alpha and beta/gamma-thrombin with the thioester peptide chloromethyl ketones, N alpha-[(acetylthio)acetyl]-D-Phe-Pro-Arg-CH2Cl and N alpha-[(acetylthio)acetyl]-D-Phe-Phe-Arg-CH2Cl, and the thiol-reactive fluorescence probe, 5-(iodoacetamido)fluorescein. Irreversible inactivation of thrombin by the inhibitors was accompanied by incorporation of 0.98 +/- 0.06 mol/mol of the thioester group into the active site, independent of a 470-fold difference between the thioester peptide chloromethyl ketones in the bimolecular rate constants of alpha-thrombin affinity labeling. Subsequent mild treatment of the covalent thrombin-inhibitor complexes with NH2OH in the presence of 5-(iodoacetamido)fluorescein resulted in generation of the thiol group together with its selective modification and incorporation of 0.96 +/- 0.07 mol of probe/mol of active sites. The incorporated label was localized to a 9000 molecular weight region of alpha and beta/gamma-thrombin containing the catalytic-site histidine residue. Evaluation of competing, side reactions showed that they did not significantly compromise the active site specificity of labeling. These results demonstrated equivalent, active-site-selective fluorescence probe labeling of alpha and beta/gamma-thrombin by use of either of the thioester peptide chloromethyl ketones, with a site specificity of greater than or equal to 94%.  相似文献   

12.
The two-way and three-way interactions among active-site-blocked bovine thrombin, bovine protein C, and the elastase fragment of rabbit thrombomodulin (elTM) were examined by analytical ultracentrifugation at 23.3 degrees C in 100 mM NaCl, 50 mM Tris (pH 7.65), and 1 mM benzamidine, in the presence of 0 to 5 mM calcium chloride. Thrombin and elTM form a tight (Kd less than 10(-8) M) 1:1 complex in the absence of Ca2+ that weakens with the addition of Ca2+ (Kd approximately 4 microM in 5 mM Ca2+). Without Ca2+, thrombin and protein C form a 1:1 complex (Kd approximately 1 microM) and what appears to be a 1:2 thrombin-protein C complex. The Kd for the 1:1 complex weakens over 100-fold in 5 mM CaCl2. Protein C and elTM form a Ca(2+)-independent 1:1 complex (Kd approximately 80 microM). Nearly identical binding to thrombin and elTM is observed when active-site-blocked activated bovine protein C is substituted for protein C. Thrombin inhibited by diisopropyl fluorophosphate and thrombin inhibited by a tripeptide chloromethyl ketone exhibited identical behavior in binding experiments, suggesting that the accessibility of protein C to the substrate recognition cleft of these two forms of thrombin is nearly equal. Human protein C binds with lower affinity than bovine protein C. Ternary mixtures also were examined. Protein C, elTM, and thrombin form a 1:1:1 complex which dissociates with increasing [Ca2+]. In the absence of Ca2+, protein C binds to the elTM-thrombin complex with an apparent Kd approximately 1 microM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
In the present report we describe the platelet-binding characteristics of applaggin and echistatin, potent inhibitors of fibrinogen-dependent platelet aggregation derived from Agkistrodon piscivorus piscivorus and Echis carinatus snake venoms, respectively. Both molecules bound to unstimulated platelets in a specific and saturable manner. At saturation there were 37,100 +/- 3,150 (mean, +/- S.D.) molecules of applaggin and 27,200 +/- 2,816 molecules of echistatin bound/platelet, with dissociation constants (Kd) of 1.4 +/- 0.6 x 10(-7) M and 4.9 +/- 1.2 x 10(-7) M, respectively. Stimulation of platelets with ADP (10 microM) + epinephrine (2 microM) resulted in an increase in the number of molecules bound at saturation to 42,300 +/- 2,105 for applaggin and 32,185 +/- 3,180 for echistatin, with a Kd of 5.6 +/- 0.3 x 10(-8) M and 1.8 +/- 0.6 x 10(-7) M, respectively. The synthetic peptide (Arg)8-Gly-Asp-Val was a competitive antagonist of applaggin and echistatin binding to unstimulated platelets (Ki = 25 and 36 microM, respectively). Applaggin and echistatin inhibited the binding of fibrinogen to stimulated platelets in a dose-dependent manner, with an IC50 of 9 and 25 nM, respectively. In concert with inhibition of platelet aggregation, applaggin and echistatin inhibited platelet secretion and synthesis of thromboxane A2 induced by ADP, collagen, and human gamma-thrombin. The monclonal antibody, LJ-CP3, which inhibits the binding of Arg-Gly-Asp containing ligands to platelet GPIIb.IIIa, also inhibited applaggin binding to unstimulated platelets in a competitive manner (Ki = 4.5 microM). Thus, applaggin and echistatin bind to the platelet GPIIb.IIIa complex, and the Arg-Gly-Asp sequence plays a central role in mediating this interaction.  相似文献   

14.
V Fleury  E Anglés-Cano 《Biochemistry》1991,30(30):7630-7638
In the present study we have quantitatively characterized the interaction of purified human Glu- and Lys-plasminogen with intact and degraded fibrin by ligand-binding experiments using a radioisotopic dilution method and antibodies against human plasminogen. A fibrinogen monolayer was covalently linked to a solid support with polyglutaraldehyde and was treated with thrombin or with thrombin and then plasmin to respectively obtain intact and degraded fibrin surfaces. Under these conditions, a well-defined surface of fibrin is obtained (410 +/- 4 fmol/cm2) and, except for a 39-kDa fragment, most of the fibrin degradation products remain bound to the support. New binding sites for plasminogen were detected on the degraded surface of fibrin. These sites were identified as carboxy-terminal lysine residues both by inhibition of the binding by the lysine analogue 6-aminohexanoic acid and by carboxy-terminal end-group digestion with carboxypeptidase B. The binding curves exhibited a characteristic Langmuir adsorption isotherm saturation profile. The data were therefore analyzed accordingly, assuming a single-site binding model to simplify the analysis. Equilibrium dissociation constants (Kd) and the maximum number of binding sites (Bmax) were derived from linearized expression of the Langmuir isotherm equation. The Kd for the binding of Glu-plasminogen to intact fibrin was 0.99 +/- 0.17 microM and for degraded fibrin was 0.66 +/- 0.22 microM. The Kd for the binding of Lys-plasminogen to intact fibrin was 0.41 +/- 0.22 microM and for degraded fibrin was 0.51 +/- 0.12 microM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Nucleotide binding to sarcoplasmic reticulum vesicles was investigated in the absence of calcium using both filtration and fluorescence measurements. Filtration assays of binding of radioactive nucleotides at concentrations up to 0.1 mM gave a stoichiometry of one ATP-binding site/sarcoplasmic reticulum ATPase molecule. When measured in the presence of calcium under otherwise similar conditions, ATPase velocity rose 4-8-fold (depending on pH and magnesium concentration) when the ATP concentration was increased from 1 microM to 0.1 mM. Binding of ATP and ADP enhanced the intrinsic fluorescence of sarcoplasmic reticulum ATPase, but AMP and adenosine did not affect it. Both filtration and fluorescence measurements showed that binding of metal-free ATP is independent of pH (Kd = 20-25 microM) but that the presence of magnesium induces pH dependence of the binding of the Mg.ATP complex (Kd = 10 microM at pH 6.0 and 1.5 microM at pH 8.0). Binding of metal-free ADP was pH-dependent but was not affected by magnesium. High magnesium concentrations inhibited nucleotide binding. These results suggest that ATP interacts with two different domains of Ca-ATPase that form the catalytic site. The first domain may bind the adenine moiety of the substrate, and the pH dependence of ADP binding suggests the participation of His683 in this region. The second domain of the catalytic site may bind the gamma-phosphate and the magnesium ion of the Mg.ATP complex and constitute the locus of the electrostatic interactions between the substrate and the enzyme.  相似文献   

16.
The interaction of vanadate ions with the Ca-ATPase from sarcoplasmic reticulum vesicles was studied in a native and a fluorescein-labeled ATPase preparation (Pick, U., and Karlish, S. J. D. (1980) Biochim. Biophys. Acta 626, 255-261). Vanadate induced a fluorescence enhancement in a fluorescein-labeled enzyme, indicating that it shifts the equilibrium between the two conformational states of the enzyme by forming a stable E2-Mg-vanadate complex (E2 is the low affinity Ca2+ binding conformational state of the sarcoplasmic reticulum Ca-ATPase). Indications for tight binding of vanadate to the enzyme (K1/2 = 10 microM) in the absence of Ca2+ and for a slow dissociation of vanadate from the enzyme in the presence of Ca2+ are presented. The enzyme-vanadate complex was identified by the appearance of a time lag in the onset of Ca2+ uptake and by a slowing of the fluorescence quenching response to Ca2+. Ca2+ prevented the binding of vanadate to the enzyme. Pyrophosphate (Kd = 2 mM) and ATP (Kd = 25 microM) competitively inhibited the binding of vanadate, indicating that vanadate binds to the low affinity ATP binding site. Binding of vanadate inhibited the high affinity Ca2+ binding to the enzyme at 4 degrees C. Vanadate also inhibited the phosphorylation reaction by inorganic phosphate (Ki = 10 microM) but had no effect on the phosphorylation by ATP. It is suggested that vanadate binds to a special region in the low affinity ATP binding site which is exposed only in the E2 conformation of the enzyme in the absence of Ca2+ and which controls the rate of the conformation transition in the dephosphorylated enzyme. The implications of these results to the role of the low affinity ATP binding sites are discussed.  相似文献   

17.
Ethanol effects on warfarin binding to human serum albumin (HSA) have been studied by equilibrium dialysis and fluorescence methods at pH 7.4 in phosphate-buffered saline at 37 degrees C. In the presence of various amounts of ethanol fluorescence intensity of bound warfarin decreased significantly but this intensity reduction was not solely from displacement of bound warfarin from HSA. By comparing fluorescence and equilibrium dialysis data we concluded that fluorescence intensity reduction of warfarin was mainly the result of changes in the surrounding environment of the warfarin binding site by ethanol interaction with HSA and that displacement of bound warfarin was not significant compared to the fluorescence intensity changes. The dissociation constant of warfarin binding to HSA decreased with an increasing amount of ethanol. From the changes in fluorescence intensity upon warfarin binding to HSA with the presence of ethanol ranging from 0 to 5.0% the following dissociation constants (Kd) were determined: 0% ethanol 5.39 +/- 0.2 microM, 0.1% ethanol 5.86 +/- 0.1 microM, 0.3% ethanol 5.83 +/- 0.2 microM, 0.5% ethanol 6.76 +/- 0.1 microM, 1% ethanol 7.01 +/- 0.1 microM, 3% ethanol 9.9 +/- 0.7 microM, 5% ethanol 13.01 +/- 0.1 microM. From the equilibrium dialysis with the same ranges of ethanol presence the following Kd values were obtained: 0% ethanol 6. 62 +/- 1.6 microM, 0.1% ethanol 6.81 +/- 1.1 microM, 0.3% ethanol 8. 26 +/- 2.5 microM, 0.5% ethanol 8.86 +/- 1.9 microM, 1% ethanol 11. 01 +/- 4.2 microM, 3% ethanol 20.75 +/- 2.4 microM, 5% ethanol 21.67 +/- 2.2 microM. The results suggest that warfarin bound to HSA was displaced by ethanol. These data indicate that ethanol influence on warfarin binding to HSA may alter the pharmacokinetics of warfarin.  相似文献   

18.
Studies of the kinetics of association and dissociation of the formycin nucleotides FTP and FDP with CF1 were carried out using the enhancement of formycin fluorescence. The protein used, derived from lettuce chloroplasts by chloroform induced release, contains only 4 types of subunit and has a molecular weight of 280 000. In the presence of 1.25 mM MgCl2, 1 mol of ATP or FTP is bound to the latent enzyme, with Kd = 10(-7) or 2 . 10(-7), respectively. The fluorescence emission (lambdamax 340 nm) of FTP is enhanced 3-fold upon binding, and polarization of fluorescence is markedly increased. The fluorescence changes have been used to follow FTP binding, which behaves as a bimolecular process with k1 = 2.4 . 10(4) M-1 . s-1. FTP is displaced by ATP in a process apparently involving unimolecular dissociation of FTP with K-1 = 3 . 10(-3) S-1. The ratio of rates is comparable to the equilibrium constant and no additional steps have been observed. The protein has 3 sites for ADP binding. Rates of ADP binding are similar in magnitude to those for FTP. ADP and ATP sites are at least partly competitive with one another. The kinetics of nucleotide binding are strikingly altered upon activation of the protein as an ATPase. The rate of FTP binding increases to at least 10(6) M-1 . s-1. This suggests that activation involves lowering of the kinetic barriers to substrate and product binding-dissociation and has implications for the mechanism of energy transduction in photophosphorylation.  相似文献   

19.
The environment of the biotin binding site on avidin was investigated by determining the fluorescence enhancement of a series of fluorescent probes that are anilinonaphthalene sulfonic acid derivatives. Of the compounds tested, 2-anilinonaphthalene-6-sulfonic acid (2,6-ANS) exhibited the greatest enhancement under the conditions used (which would reflect both molar fluorescence enhancement and binding affinity) and exhibited more than 95% reversal upon addition of biotin. Thus, 2,6-ANS was chosen for more detailed characterization of the interaction with avidin. Only a single class of binding sites for 2,6-ANS was identified; the mean value for the Kd was 203 +/- 16 microM (X +/- 1 S.D.), and the molar ratio of 2,6-ANS binding sites to biotin binding sites was approx. 1. These results provide evidence that the biotin binding site and the 2,6-ANS binding site are at least partially overlapping, but the possibility that the probe binding site is altered by a conformational change induced by biotin binding cannot be excluded. At excitation = 328 nm and emission = 408 nm, the molar fluorescence of the bound probe was 6.8 +/- 1.0 microM-1 and that of the free probe was 0.061 +/- 0.008 microM-1 giving an enhancement ratio (molar fluorescence of bound probe/molar fluorescence of free probe) of 111 +/- 22. Upon binding, the wavelength of maximum fluorescence decreases. These findings also provide evidence that the fluorescence enhancement associated with the interaction of 2,6-ANS and avidin reflects the environment of the biotin binding site. The Kosower's Z factor, an empirical index of apolarity, was 82.1 for the 2,6-ANS binding site on avidin. This value reflects a degree of apolarity that is similar to apolar environments observed for substrate binding sites on several enzymes; although not the dominant factor, this environment may contribute to the strong binding of biotin.  相似文献   

20.
A Muhlrad 《Biochemistry》1989,28(9):4002-4010
The 23-kDa N-terminal tryptic fragment was isolated from the heavy chain of rabbit skeletal myosin subfragment 1 (S-1). The heavy-chain fragments were dissociated by guanidine hydrochloride following limited trypsinolysis, and the 23-kDa fragment was isolated by gel filtration and ion-exchange chromatography. Finally, the fragment was renatured by removing the denaturants. The CD spectrum of the renatured fragment shows the presence of ordered structure. The tryptophan fluorescence emission spectrum of the fragment is considerably shifted to the red upon adding guanidine hydrochloride which indicates that the tryptophans are located in relatively hydrophobic environments. The two 23-kDa tryptophans, unlike the rest of the S-1 tryptophans, are fully accessible to acrylamide as indicated by fluorescence quenching. The isolated 23-kDa fragment cosediments with F-actin in the ultracentrifuge and significantly increases the light scattering of actin in solution which indicates actin binding. The binding is rather tight (Kd = 0.1 microM) and ionic strength dependent (decreasing with increasing ionic strength). ATP, pyrophosphate, and ADP dissociate the 23-kDa-actin complex with decreasing effectiveness. The isolated 23-kDa fragment does not have ATPase activity; however, it inhibits the actin-activated ATPase activity of S-1 by competing presumably with S-1 for binding sites on actin. F-Actin binds to the 23-kDa fragment immobilized on the nitrocellulose membrane. The fragment was further cleaved, and one of the resulting peptides, containing the 130-204 stretch of residues, was found to bind actin on the nitrocellulose membrane, indicating that this region of the 23-kDa fragment participates in forming an actin binding site.  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