首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The binding of divalent haptens to IgE-class antibodies leads predominantly to their oligomerization into open and closed dimers. Kinetics of the open dimer formation was investigated by fluorescence titrations of Fab fragments of monoclonal DNP-specific IgE antibodies with divalent haptens having different spacer length (gamma = 14-130 A). Binding was monitored by quenching of intrinsic tryptophan emission of the Fab. Addition of divalent haptens with short spacers (gamma = 14-21 A) to the Fabs at rates larger than a distinct threshold value caused a significant decrease of Fab-binding site occupation in the initial phase of the titration. This finding was interpreted to reflect a nonequilibrium state of hapten-Fab-binding. Such nonequilibrium titrations were analyzed by inserting a kinetic model into a theory of antibody aggregation as presented by Dembo and Golstein (Histamine release due to bivalent penicilloyl haptens. 1978. J. Immunol. 121, 345). Fitting of this model to the fluorescence titrations yielded dissociation rate constants of 7.8 x 10(-3) s-1 and 6 x 10(-3) s-1 for the Fab dimers formed by the flexible divalent haptens N alpha, N epsilon-di(dinitrophenyl)-L-lysine (gamma = 16 A) and bis(N beta-2,4-dinitrophenyl-alanyl)-meso-diamino-succinate (gamma = 21 A). Making the simplifying assumption that a single step binding equilibrium prevails, the corresponding dimer formation rate constants were calculated to be 1.9 x 10(5) M-1 s-1 and 1.1 x 10(4) M-1 s-1, respectively. In contrast, all haptens with spacers longer than 40 A (i.e., bis(N alpha-2,4-dinitrophenyl-tri-D-alanyl)-1,7-diamino-heptane, and di(N epsilon-2,4-dinitrophenyl)-6-aminohexanoate-aspartyl-(prolyl)n-L-l ysyl (n = 24, 27, 33) exhibit a relative fast dimerization rate of the Fab fragments (greater than 7 x 10(6) M-1 s-1). These observations were interpreted as being caused by orientational constraints set by the limited solid angle of the reaction between the macromolecular reactants. Thus, ligands having better access to the binding site would react faster.  相似文献   

2.
The interaction of L-tyrosine, L-tyrosyladenylate and tRNA-Tyr with tyrosyl-tRNA synthetase from Bacillus stearothermophilus was studied by equilibrium dialysis, gel filtration and fluorescence spectroscopy. The enzyme, which consists of two identical subunits (mol. wt 2 x 44000), binds only a single molecule of L-tyrosine per dimer with a K-d of 2 x 10-5 M at pH 7.8 and 23 degrees C. The tyrosyl-tRNA synthetase--tyrosyladenylate complex which was isolated by gel filtration also has one adenylate bound per dimeric enzyme molecule. In contrast, two tRNA-Tyr molecules bind per enzyme dimer, but the two binding sites are not equivalent having K-d values of 2 x 10-7 M and 1.3 x 10-6 M respectively at pH 6.5 and 25 degrees C. Since crystallographic analysis of the free enzyme [2] shows that the monomer is the asymmetric unit, the data indicate that substrate binding induces asymmetry in the enzyme.  相似文献   

3.
For monovalent ligands interacting with cell surface receptors we have directly observed the functional dependence of the forward rate constant on the number of receptors per cell (N). The experimental system we studied consisted of monovalent ligand, 2,4-dinitrophenyl (DNP)-aminocaproyl-L-tyrosine (DCT), binding to bivalent, monoclonal anti-DNP immunoglobulin E (IgE) anchored to its high affinity receptor on rat basophilic leukemia (RBL) cells. To measure the fractional occupation of antibody combining sites by DNP we employed a recently developed fluorescence technique (Erickson, J., Kane, B. Goldstein, D. Holowka, and B. Baird, 1986, Mol. Immunol., 72:769-781). Our results are well fitted by the equation (Berg and Purcell, 1977, Biophys. J., 20:193-219) konc = 4 pi DaN kappa on/[4 pi Da + N kappa on] where konc is the forward rate constant for binding to the cell, D is the diffusion constant of the ligand, a is the radius of the cell, and kappa on is the intrinsic forward rate constant describing a single IgE combining site-DNP interaction. If D is fixed at 10(-5) cm2/s, the best fit of accumulated data predicts an average cell radius of approximately 4 microns and kappa on of approximately 1.8 x 10(-13) cm3/s [1.1 x 10(8)(M . s)-1]; both in excellent agreement with RBL cell size and the single-site forward rate constant for the binding of DCT to IgE in solution, respectively. We believe this is the first report of experimental evidence that directly illustrates the effect of surface density in determining the rates of binding for small molecules to membrane receptors.  相似文献   

4.
Kinetic parameters of the interaction of the toxic lectins abrin and ricin with human erythrocytes and HeLa cells have been measured. The binding of 125I-labeled abrin and ricin to human erythrocytes and to HeLa cells at 37 degrees was maximal around pH 7, whereas at 0 degrees the binding was similar over a broad pH range. The binding occurred at similar rates at 0 degrees and 37 degrees with rate constants in the range 0.9 to 3.0 X 10(5) M-1 s-1. The dissociation was strongly temperature-dependent with rate constants in the range 3.4 to 45 X 10(-4) s-1 at 0 degrees and 3.9 to 18 X 10(-3) s-1 at 37 degrees. The presence of unlabeled lectins as well as lactose increased the rate of dissociation. The association constants measured at equilibrium or calculated from the rate constants were between 0.64 X 10(8) M-1 and 8.2 X 10(8) M-1 for abrus lectins, and between 8.0 X 10(6) M-1 and 4.2 X 10(8) M-1 for ricinus lectins. The association constants for the toxins were lower at 37 degrees than at 0 degrees. Isolated ricin B chain appeared to bind with similar affinity as intact ricin. The number of binding sites was estimated to be 2 to 3 X 10(6) per erythrocyte and 1 to 3 X 10(7) per HeLa cell. The binding sites of HeLa cells all displayed a uniform affinity towards abrin and ricin, both at 0 degrees and at 37 degrees. The same was the case with the binding sites of erythrocytes at 0 degrees. However, the data indicated that at 20 degrees erythrocytes possessed binding sites with two different affinities. Only a fraction of the cell-bound toxin appeared to be irreversibly bound and could not be removed by washing with 0.1 M lactose. The fraction of the total amount of bound toxin which became irreversibly bound to HeLa cells was for both toxins about 2 X 10(-3)/min at 37 degrees, whereas no toxin was irreversibly bound at 0 degrees. In the case of erythrocytes no toxin became irreversibly bound, either at 0 degrees or 37 degrees, indicating that the toxins are unable to penetrate into these cells.  相似文献   

5.
A mAb was isolated (mAb BD6) that recognized a surface glycoprotein on rat basophilic leukemia cells (RBL-2H3). The antibody bound to 2 x 10(6) molecules/cell and specifically blocked IgE binding (50% inhibition with 3.48 +/- 0.51 micrograms/ml; mean +/- SEM), although neither IgE nor anti-high affinity IgE receptor (anti-Fc epsilon RI) mAb blocked mAb BD6 binding to the cells. mAb BD6 did not affect the rate of dissociation of cell-bound IgE, nor did it induce or inhibit the internalization of IgE. mAb BD6 did not release histamine. However, it did cause rapid spreading of the cells. By 1 h the cells had retracted to a spherical shape with their surface covered with membranous spikes, and they could easily be detached from the tissue culture plate. These changes differed from those observed after Fc epsilon RI activation. mAb BD6 immunoprecipitated a complex of two proteins, 38 to 50 kDa and 135 kDa from 125I-surface labeled rat basophilic leukemia cells that are not subunits of Fc epsilon RI. Chemical cross-linking studies showed that these molecules are associated on the cell surface. By immunoblotting, mAb BD6 reacted with a 40-kDa protein. Therefore, mAb BD6 binds to a surface protein that is close to the Fc epsilon RI and sterically inhibits 125I-IgE binding.  相似文献   

6.
1. The interactions of lanthanide metals and dinitrophenyl spin-label haptens with the Fv fragment of the mouse myeloma protein MOPC 315 were investigated by the techniques of fluorescence, e.s.r. (electron spin resonance) and high-resolution n.m.r. (nuclear magnetic resonance). 2. The protein fluorescence of Fv fragment at 340nm is quenched by the haptens (fluorescence enhancement, epsilon=0.15) and enhanced by Gd(III) (epsilon=1.14) and other lanthanides. The binding of the haptens studied here is insensitive to pH in the range 5.5-7.0 (dissociation constant KH=0.3-1.0 muM) and shows 1:1 stoicheiometry. The binding of Gd(III) also shows 1:1 stoicheiometry, but is pH-dependent; the binding constant (KM) varies from 10 muM at pH7.0 to 700 muM at pH4.8. La(III) binding is less sensitive to pH. The pH-dependences of the metal-binding constants imply that a group in the protein with pKa greater than or equal to 6.2 is involved in the binding, and probably also other groups with lower pKa values. 3. The apparent binding of the haptens is weakened about 20-fold by Gd(III), and vice versa. An equilibrium scheme involving a ternary complex with an interaction between the two binding sites is derived in Appendix I to explain the experimental results at two pH values. 4. Time-dependent fluorescence changes are observed in the presence of Gd(III) at pH5.5. A two-state kinetic scheme involving a 'slow' conformational change in the Fv fragment is derived in Appendix II to explain this time-dependence. This scheme is consistent with the antagonistic equilibrium behaviour. 5. The e.s.r. changes in the spin-label haptens on binding to Fv fragment and on the subsequent addition of lanthanides are consistent with the binding scheme for haptens and lanthanides proposed from the fluorescence studies. A difference between the limiting quenching of the e.s.r. signal from the bound haptens in the presence of saturating concentrations of Gd(III) and La(III) is attributed to dipolar interactions between bound Gd(III) and the nitroxide moiety of the bound hapten. The residual quenching with Gd(III) allows an estimate of 1.2nm to be made for the distance between the two paramagnetic centres. 6. The 270 MHz proton difference spectrum of the Fv fragment resulting from the addition of La(III) suggests that any metal-induced conformational changes are small and involve relatively few amino acid residues on the Fv fragment...  相似文献   

7.
1. Rat alpha-foetoprotein, an oestrogen-binding foetal globulin, was isolated in large quantities from amniotic fluid and serum by preparative electrophoresis on polyacrylamide slab gels or by chromatography on an immunoadsorbent column. Subsequently the two electrophoretic forms of this protein were separated by electrophoresis on the same medium. 2. Both forms were found to show identical binding with oestradiol. From the extrinsic fluorescence of the bound dye 8-anilinonaphthalene-1-sulphonic acid it was shown that the polarity of the binding site is practically identical for both forms. One residue of tryptophan was determined for both forms. The two electrophoretic variants display the same amount of secondary structure as demonstrated by circular dichroism. 3. The affinity of total alpha-foetoprotein for oestradiol as a function of pH was studied by using a Sephadex G-25 gel-equilibration method. Maximal binding occurred at pH8.5. Only a fractional number of binding sites per molecule could be measured at pH7.4, whereas at higher pH the number of sites was very close to unity. There was no significant effect of pH on the value of the association constant (K(a)=4.3x10(7)+/-1.2x10(7)m(-1)). 4. Displacement experiments of bound labelled oestradiol with various steroids have permitted investigation of the specificity of alpha-foetoprotein. This foetal globulin binds rather strongly compounds that display the rigid structure of the oestratriene skeleton (oestradiol, oestrone). Diminished binding for diethylstilboestrol and a diethylstilboestrol affinity label was observed. No binding was measured with a more flexible structure such as hexoestrol [4,4'-(1,2-diethylethane-1,2-diyl)bisphenol]. 5. Chemical modification of cysteine residues of alpha-foetoprotein with two alkylating reagents [iodoacetic acid and 8-[N-(iodoacetylaminoethyl)amino]naphthalene-1-sulphonic acid] has very little effect on the oestrogen binding. It is suggested that the oestrogen-binding site does not contain a cysteine residue. From the kinetics of alkylation and from the fluorescence properties of the chemically bound thiol reagent 8-[N-(iodoacetylaminoethyl)amino]naphthalene-1-sulphonic acid], it was demonstrated that the very-slow-reacting thiol group is probably located in a non-polar region of the molecule.  相似文献   

8.
The kinetics of the binding of the fluorescence indicator N-phenyl naphthylamine to bilayer vesicles of C12-methyl-phosphatidic acid have been investigated by means of the temperature-jump relaxation technique utilizing fluorescence light detection. Single-exponential relaxation curves were observed, with time constants in the range 0.2-3 ms. The concentration dependence of the relaxation time yielded an apparent association rate constant (expressed in terms of monomeric phospholipid) of k(on) = 5 x 10(6) M(-1) s(-1) in aqueous solution at 25 degrees . The activation energy and viscosity dependence associated with the binding rate show that this process is actually diffusion-controlled. The theory of diffusion-controlled reactions then allows a determination of the average size of the bilayer vesicles and of the true rate constant for the association of the indicator molecules with the vesicles. Assuming spherical geometry for the vesicles, the values are: r(ves) = 190 A, which corresponds to 20000 lipid molecules per vesicle and k'(on) = 1 x 10(11) M(-1) s(-1) (25 degrees). The correctness of this size-determination was confirmed semi-quantitatively by electron microscopy. Since in fact a distribution of vesicle sizes must be present, a discussion is included of the relaxation function which the system is expected to take in the general case. Biological implications of diffusion control for the transport of non-polar substances and for lipid mixing are indicated.  相似文献   

9.
The effect of temperature on the binding of thyroxine and triiodothyronine to thyroxine-binding globulin has been studied by equilibrium dialysis. Inclusion of ovalbumin in the dialysis mixture stabilized thyroxine-binding globulin against losses in binding activity which had been found to occur during equilibrium dialysis. Ovalbumin by itself bound the thyroid hormones very weakly and its binding could be neglected when analyzing the experimental results. At pH 7.4 and 37 degrees in 0.06 M potassium phosphate/0.7 mM EDTA buffer, thyroxine was bound to thyroxine-binding globulin at a single binding site with apparent association constants: at 5 degrees, K = 4.73 +/- 0.38 X 10(10) M-1; at 25 degrees, K = 1.55 +/- 0.17 X 10(10) M-1; and at 37 degrees, K = 9.08 +/- 0.62 X 10(9) M-1. Scatchard plots of the binding data for triiodothyronine indicated that the binding of this compound to thyroxine-binding globulin was more complex than that found for thyroxine. The data for triiodothyronine binding could be fitted by asuming the existence of two different classes of binding sites. At 5 degrees and pH 7.4 nonlinear regression analysis of the data yielded the values n1 = 1.04 +/- 0.10, K1 = 3.35 +/- 0.63 X 10(9) M-1 and n2 = 1.40 +/- 0.08, K2 = 0.69 +/- 0.20 X 10(8) M-1. At 25 degrees, the values for the binding constants were n1 = 1.04 +/- 0.38, K1 = 6.5 +/- 2.8 X 10(8) M-1 and n2 = 0.77 +/- 0.22, K2 = 0.43 +/- 0.62 X 10(8) M-1. At 37 degrees where less curvature was observed, the estimated binding constants were n1 = 1.02 +/- 0.06, K1 = 4.32 +/- 0.59 X 10(8) M-1 and n2K2 = 0.056 +/- 0.012 X 10(8) M-1. When n1 was fixed at 1, the resulting values obtained for the other three binding constants were at 25 degrees, K1 = 6.12 +/- 0.35 X 10(8) M-1, n2 = 0.72 +/- 0.18, K2 = 0.73 +/- 0.36 X 10(8) M-1; and at 37 degrees K1 = 3.80 +/- 0.22 X 10(8) M-1, n2 = 0.44 +/- 0.22, and K2 = 0.43 +/- 0.38 X 10(8) M-1. The thermodynamic values for thyroxine binding to thyroxine-binding globulin at 37 degrees and pH 7.4 were deltaG0 = -14.1 kcal/mole, deltaH0 = -8.96 kcal/mole, and deltaS0 = +16.7 cal degree-1 mole-1. For triiodothyronine at 37 degrees, the thermodynamic values for binding at the primary binding site were deltaG0 = -12.3 kcal/mole, deltaH0 = -11.9 kcal/mole, and deltaS0 = +1.4 cal degree-1 mole-1. Measurement of the pH dependence of binding indicated that both thyroxine and triiodothyronine were bound maximally in the region of physiological pH, pH 6.8 to 7.7.  相似文献   

10.
Bovine serum albumin (BSA) is routinely utilized in vitro to prevent the adverse detergent effects of long-chain acyl-CoA esters (i.e., palmitoyl-CoA) in enzyme assays. Determination of substrate saturation kinetics in the presence of albumin would only be valid if the relationship between bound and free substrate concentrations was known. To elucidate the relationship between bound and free palmitoyl-CoA concentrations in the presence of BSA, several different techniques including equilibrium dialysis, equilibrium partitioning, fluorescence polarization and direct fluorescence enhancement were investigated. Direct fluorescence enhancement using a custom synthesized fluorescent probe, 16-(9-anthroyloxy)palmitoyl-CoA (AP-CoA), was the best approach to this question. Measurement of the relationship between mol of palmitoyl-CoA bound per mol of BSA (nu) versus -log[free palmitoyl-CoA] revealed that the binding of palmitoyl-CoA to BSA, like palmitate was nonlinear, suggesting the presence of more than one class of acyl-CoA binding sites. Computer analyses of the binding data gave a best fit to the 2,4 two-class Scatchard model, suggesting the presence of two high-affinity primary binding sites (k1 = (1.55 +/- 0.46) x 10(-6) M-1) and four lower affinity secondary binding sites (k2 = (1.90 +/- 0.09) x 10(-8) M-1). Further analyses using the six parameter stoichiometric (stepwise) ligand binding model supports the existence of six binding sites with the higher affinities associated with the binding of the first mole of palmitoyl-CoA and weaker binding occurring after the first two sites are occupied. The association constants from this model of multiple binding diminish sequentially (i.e., K1 greater than K2 greater than K3 greater than...greater than or equal to K6), suggesting that each mol of long-chain acyl-CoA binds to BSA with decreasing affinities.  相似文献   

11.
The interaction of several metallo-porphyrins with the galactose-specific lectin from Trichosanthes cucumeirna (TCSL) has been investigated. Difference absorption spectroscopy revealed that significant changes occur in the Soret band region of the porphyrins upon binding to TCSL and these changes have been monitored to obtain association constants (Ka) and stoichiometry of binding (n). The dimeric lectin binds two porphyrin molecules and the presence of the specific saccharide lactose did not affect porphyrin binding significantly, indicating that the sugar and the porphyrin bind at different sites. The Ka values obtained for the binding of different porphyrins with TCSL at 25 degrees C were in the range of 2 x 10(3)-5 x 10(5) m(-1). Association constants for meso-tetra(4-sulphonatophenyl)porphyrinato copper(II) (CuTPPS), a porphyrin bearing four negative charges and meso-tetra(4-methylpyridinium)porphyrinato copper(II) (CuTMPyP), a porphyrin with four positive charges, were determined at several temperatures; from the temperature dependence of the association constants, the thermodynamic parameters change in enthalpy (DeltaH degrees ) and change in entropy (DeltaS degrees ) associated with the binding process were estimated. The thermodynamic data indicate that porphyrin binding to TCSL is driven largely by a favourable entropic contribution; the enthalpic contribution is very small, suggesting that the binding process is governed primarily by hydrophobic forces. Stopped-flow spectroscopic measurements show that binding of CuTMPyP to TCSL takes place by a single-step process and at 20 degrees C, the association and dissociation rate constants were 1.89 x 10(4) m(-1).s(-1) and 0.29 s(-1), respectively.  相似文献   

12.
6,7-Dimethyllumazine derivatives, substituted at the 8-position with aldityls or monohydroxyalkyl groups, have been examined for their binding ability to lumazine apo-protein from two strains of Photobacterium phosphoreum using fluorescence dynamics techniques. On the protein the lumazine has a nearly monoexponential decay of fluorescence with lifetime 13.8 ns (20 degrees C). In free solution the lifetime is 9.6 ns. The concentration of free and bound lumazine in an equilibrium mixture can be recovered readily by analysis of the fluorescence decay. Only the aldityl derivatives D-xylityl and 3'-deoxy-D-ribityl, having stereoconfigurations at the 2' and 4' positions identical to the natural ligand, 8-(1'-D-ribityl), show comparable dissociation constants (0.3 microM, 20 degrees C, pH 7.0). D-Erythrityl and L-arabityl have dissociation constants of 1-2 microM. All other ligands show no interaction at all or have dissociation constants in the range 6-80 microM, which can still be determined semi-quantitatively using the fluorescence decay technique. In the case of these very weakly bound ligands, unambiguous detection of bound ligand can be shown by a long correlation time (23 ns, 2 degrees C) for the fluorescence anisotropy decay. Examination of the bound D-xylityl compound's fluorescence anisotropy decay at high time resolution (< 100 ps) shows rigid association, i.e. no mobility independent of the macromolecule. All bound ligands appear to be similarly positioned in the binding site. The influence of the stereoconfiguration at the 8-position found for lumazine protein parallels that previously observed for the enzyme riboflavin synthase, where the lumazines are substrates or inhibitors. This is consistent with the finding of significant sequence similarity between these proteins. The binding rigidity may have implications for the mechanism of the enzyme.  相似文献   

13.
Immunization of BALB/c mice with a cadmium-chelate-protein conjugate resulted in the isolation of two hybridoma cell lines (A4 and E5) that synthesized antibodies with different variable regions, but similar metal-chelate affinity. The ability of these two monoclonal antibodies to interact with 12 different metal-chelate complexes was studied using the KinExA 3000 immunoassay instrument. The two antibodies showed the highest affinity for cadmium and mercury complexes of ethylenediamine N,N,N',N'-tetraacetic acid (EDTA). The E5 antibody bound to EDTA complexes of cadmium and mercury with equilibrium dissociation constants (K(d)) of 1.62 x 10(-)(9) M and 3.64 x 10(-)(9) M, respectively. The corresponding values for the A4 antibody were 14.7 x 10(-)(9) M and 3.56 x 10(-)(9) M. Addition of a cyclohexyl ring to the EDTA backbone increased the affinity of E5 for the metal-chelate haptens, while decreasing the binding of A4 to the same haptens. Based on available crystal structures, molecular models were constructed for five different divalent metal-chelate complexes. The models were compared to determine structural features of the haptens that may influence antibody recognition. Difference distance matrixes were used to identify areas of the metal-chelate haptens that differed in three-dimensional space. Antibody affinity correlated well with the extent of total structural difference for these metal-EDTA complexes.  相似文献   

14.
Binding and proximity relationships of hydrophobic ligands on human serum albumin have been studied using absorption, fluorescence, circular dichroism, and electron paramagnetic resonance spectroscopy. The ligands studied were bilirubin, two conjugated linear polyene fatty acids, cis-parinaric acid and cis-eleostearic acid, and three nitroxide derivatives of stearic acid with doxyl groups at positions 5, 10, and 12, respectively. Binding of polyene fatty acids was monitored by absorption peak shifts, induced circular dichroism, enhancement of fluorescence, and energy transfer between albumin's single tryptophanyl residue and the polyene chromophore. Induced circular dichroism studies indicate excitonic ligand-ligand interaction between bound fatty acids. Fluorescence enhancement of cis-parinaric acid was analyzed using a stepwise multiple equilibrium model, and six binding constants in the range 10(8) to 10(6) M-1 were obtained, in agreement with previous measurements for other fatty acids. The temperature dependence of the equilibrium constants indicates that the binding enthalpy is nearly zero. Fluorescence energy transfer was similarly used to quantitate bilirubin binding to albumin. Energy transfer, nitroxide quenching of fluorescence, and electron paramagnetic resonance spectroscopy were used to elucidate binding geometries which support and extend proposed structural models for albumin. It is suggested that the first two fatty acids bind side-by-side in an antiparallel fashion in domain III of human serum albumin.  相似文献   

15.
Tropomyosin, cross-linked at cysteine 190, was found to bind more weakly to actin filaments than uncross-linked tropomyosin. Cross-linking of tropomyosin can cause actin filaments nearly completely covered with tropomyosin to be uncovered almost completely. The critical monomer concentration of actin is not significantly changed by binding of cross-linked or uncross-linked tropomyosin to actin filaments. The binding curves were analyzed quantitatively, thereby taking into account the polar end-to-end contact of tropomyosin molecules bound by actin and the overlap of the seven subunit binding sites along the actin filament. Under the conditions of the experiment (80 mM KCl, 1 mM MgCl2, pH 7.5, 38-42 degrees C), the equilibrium constant for isolated binding of tropomyosin to actin filaments is in the range 1 x 10(3)-3 x 10(3) M-1. The equilibrium constants for binding of tropomyosin to binding sites along the actin filament with one or two neighbouring tropomyosin molecules are in the range of 10(6) or 10(8) to 10(9) M-1, respectively. The equilibrium constants for binding of tropomyosin to binding sites along the actin filament with one or two neighbouring tropomyosin molecules are in the range of 10(6) or 10(8) to 10(9) M-1, respectively. The equilibrium constants for cross-linked and uncross-linked tropomyosin differ by a factor of only about two. Owing to the highly cooperative binding, these differences are sufficient so that actin filaments nearly completely covered with uncross-linked tropomyosin are uncovered almost completely by cross-linking tropomyosin at cysteine 190.  相似文献   

16.
We studied galactose (Gal)-specific binding of the soluble purified 260-kDa Entamoeba histolytica adherence protein to glycosylation deficient Chinese hamster ovary (CHO) cell mutants. Our goal was to further define the lectin's functional activity and carbohydrate receptor specificity. The adherence protein was purified by acid elution from an immunoaffinity column; however, exposure of the surface membrane lectin on viable trophozoites to identical acid pH conditions had no effect on carbohydrate binding activity. Saturable Gal-specific binding of soluble lectin to parental CHO cells was demonstrated at 4 degrees C by radioimmunoassay; the dissociation coefficient (Kd) was 2.39 x 10(-8) M-1 with 5.97 x 10(4) lectin receptors present per CHO cell. Gal-specific binding of lectin to Lec2 CHO cell mutants, which have increased N- and O-linked terminal Gal residues on cell surface carbohydrates, was increased due to an enhanced number of receptors (2.41 x 10(5)/cell) rather than a significantly reduced dissociation constant (4.93 x 10(-8) M-1). At 4 degrees C, there was no measurable Gal-specific binding of the adherence protein to the Lec1 and 1dlD.Lec1 CHO mutants, which contain surface carbohydrates deficient in terminal Gal residues. Binding of lectin (20 micrograms/ml) to CHO cells was equivalent at 4 degrees C and 37 degrees C and unaltered by adding the microfilament inhibitor, Cytochalasin D (10 micrograms/ml). Gal-specific binding of the lectin at 4 degrees C was calcium independent and reduced by 81% following adsorption of only 0.2% of the lectin to CHO cells.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
W Y Lin  C D Eads  J J Villafranca 《Biochemistry》1991,30(14):3421-3426
TNS, 2-p-toluidinylnaphthalene-6-sulfonate, has been used as a fluorescent probe to determine the binding constants of metal ions to the two binding sites of Escherichia coli glutamine synthetase (GS). TNS fluorescence is enhanced dramatically when bound to proteins due to its high quantum yield resulting from its interactions with hydrophobic regions in proteins. The fluorescence energy transfer from a hydrophobic tryptophan residue of GS to TNS has been detected as an excitation band centered at 280 nm. Therefore, TNS is believed to be bound to a hydrophobic site on the GS surface other than the active site and is located near a hydrophobic Trp residue of GS. GS binds lanthanide ions [Ln(III)] more tightly than either Mn(II) or Mg(II), and the binding constants of several lanthanide ions were determined to be in the range (2.1-4.6) x 10(10) and (1.4-3.0) x 10(8) M-1 to the two metal binding sites of GS, respectively. The intermetal distances between the two metal binding sites of GS were also determined by measuring the efficiencies of energy transfer from Tb(III) to other Ln(III) ions. The intermetal distances of Tb(III)-Ho(III) and Tb(III)-Nd(III) were 7.9 and 6.8 A, respectively.  相似文献   

18.
Immunogenic peptides have been shown to bind detergent-solubilized class II (Ia) molecules from mice. In this investigation, we report that highly purified HLA-DR (DR) molecules in detergent solution are capable of binding a synthetic peptide (HAp) derived from the influenza hemagglutinin sequence. Although the presentation of this peptide has been demonstrated only to DR1-restricted Th cells, the association rate constants for the formation of HAp-DR1, -DR5, and -DR8 complexes were essentially identical (ka = 1.1 x 10(2) to 1.6 x 10(2) M-1 s-1). By contrast, the value of the rate constants for the dissociation of preformed HAp-DR1, -DR5, and -DR8 complexes varied nearly threefold (kd = 1.6 x 10(6) to 4.4 x 10(-6) s-1). The value of the equilibrium dissociation constants (KD) derived from these rate constants were 13 nM, 24 nM, and 28 nM, for HAp-DR1, -DR5, and -DR8 complexes, respectively. Scatchard analysis demonstrated that the KD obtained from the rate constants for the HAp-DR1 reaction was in excellent agreement with that obtained under equilibrium conditions. SDS-PAGE confirmed that the HAp-DR complexes were remarkably stable, as HAp remained associated with the DR alpha beta heterodimer after treatment of the complexes with SDS and beta-mercaptoethanol. Steady-state binding studies demonstrated that 18% of all DR1 molecules had bound HAp at equilibrium, whereas only 3.8% of all DR8 molecules had bound HAp under identical conditions. The slight differences in the KD for HAp-DR complexes suggest that differences in the affinity of a peptide for DR alleles alone may not always explain the process of MHC restriction.  相似文献   

19.
Bacteriophage T7 gene 2.5 protein has been purified to homogeneity from cells overexpressing its gene. Native gene 2.5 protein consists of a dimer of two identical subunits of molecular weight 25,562. Gene 2.5 protein binds specifically to single-stranded DNA with a stoichiometry of approximately 7 nucleotides bound per monomer of gene 2.5 protein; binding appears to be noncooperative. Electron microscopic analysis shows that gene 2.5 protein is able to disrupt the secondary structure of single-stranded DNA. The single-stranded DNA is extended into a chain of gene 2.5 protein dimers bound along the DNA. In fluorescence quenching and nitrocellulose filter binding assays, the binding constants of gene 2.5 protein to single-stranded DNA are 1.2 x 10(6) M-1 and 3.8 x 10(6) M-1, respectively. Escherichia coli single-stranded DNA-binding protein and phage T4 gene 32 protein bind to single-stranded DNA more tightly by a factor of 25. Fluorescence spectroscopy suggests that tyrosine residue(s), but not tryptophan residues, on gene 2.5 protein interacts with single-stranded DNA.  相似文献   

20.
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   

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