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.
Generalized binding phenomena in an allosteric macromolecule   总被引:2,自引:0,他引:2  
A general macromolecular partition function is developed in terms of chemical ligand activity, temperature and pressure for systems described by an array of species which are characterized by their state of allosteric conformation and ligand stoichiometry. The effects of chemical ligand binding, enthalpy change, and volume change are treated in a parallel manner. From a broad viewpoint all of these effects can be regarded as specific cases of generalized binding phenomena. This approach provides a general method for analyzing calorimetric and ligand binding experiments. Several applications are given: (1) Thermal scanning data for tRNAphe (P.L. Privalov and V.V. Filimonov, J. Mol. Biol. 122 (1978) 447) are shown to fit a general model with six conformational states. By application of linkage theory it is shown that sodium chloride is expelled as the molecule denatures. (2) The results of calorimetric titrations on the arabinose binding protein (H. Fukada, J.M. Sturtevant and F.A. Quiocho, J. Mol. Biol. 258 (1983) 13193) are shown to fit a simple two-state allosteric model. (3) A thermal binding curve is simulated for an unusual respiratory protein, trout I hemoglobin (B.G. Barisas and S.J. Gill, Biophys. Chem. 9 (1979) 235), in order to illustrate both the similarities and differences between enthalpy and chemical ligand binding processes.  相似文献   

2.
C H Robert  S J Gill  J Wyman 《Biochemistry》1988,27(18):6829-6835
We present a general framework for analysis of two closely related problems in biochemical studies: (1) The first is analysis of binding data obtained under conditions in which a second, linked ligand is present in limited total quantity. In such conditions the free activity of the second ligand varies throughout the primary ligand binding curve, and the resultant behavior can be quite complex. Analysis of such curves enables one to quantitatively extract detailed information regarding the linkage of the two ligands at intermediate stages of ligation. The treatment is applied in an accompanying paper to oxygen binding in human hemoglobin in the presence of organic phosphates [Robert, C.H., Fall, L., & Gill, S. J. (1988) Biochemistry (following paper in this issue)]. (2) The second treatment we outline regards the analogous problem of analyzing differential scanning calorimetry (DSC) data obtained for a macromolecule binding a ligand present in limited quantity. A simple model is presented that accounts for dual transitions like those already seen in DSC data for human serum albumin in the presence of nonsaturating amounts of fatty acids [Ross, P., & Shrake, A. (1987) Abstracts of the 42nd Calorimetry Conference, University of Colorado, Boulder, CO].  相似文献   

3.
Reversible binding of a ligand to an enzyme active site can elicit a variety of changes in the protein, such as conformational changes (close to the site of binding or communicated over long distances), changes in the ionization state of surrounding amino acid side chains, changes in the interaction of the target protein with other subunits (or other proteins), or even changes in the thermodynamic stability of the protein. Relatively little attention has been given to studying these effects in proteins to which the ligand has been irreversibly bound, yet this can be a convenient way of studying the effects of ligand binding in the absence of association/dissociation equilibria. We report the dramatic changes which occur to the shikimate pathway enzyme dehydroquinase when ligand is attached to its active site after borohydride reduction of the mechanistically important Schiff's base intermediates. The effects of this modification have been characterized by limited proteolysis, circular dichroism, guanidine hydrochloride denaturation, and differential scanning calorimetry. The conclusions from these studies are that although anchoring the ligand at the active site does not cause a gross change in conformation, it does increase markedly the conformational stability of the protein. This is conclusively established by three separate experiments: 1) the modified protein is completely resistant to proteases, whereas the unmodified protein is very susceptible to proteolysis; 2) the concentration of guanidine hydrochloride required to unfold the ligand-linked dehydroquinase is 3-4-fold greater than that of the unmodified protein; 3) the melting temperature (Tm) of the modified protein is 40 degrees C higher than that of the unmodified protein. These results are a very clear example of the thermodynamic link between ligand binding, conformational stability, and proteolytic susceptibility in vitro and will be a useful system for dissecting the contributions of individual protein-ligand interactions to these parameters.  相似文献   

4.
An investigation of the crystal structure of cytosolic pigheart aspartate aminotransferase (AAT, E.C.2.6.1.1) was carried out to determine the structural requirements for ligand recognition by the active site. Structural differences were observed between the two active sites of the AAT dimer. The natural ligand, l-aspartate, was docked into both active sites using various methods. However, due to structural differences, the ligand was able to form all the necessary interactions for initial binding in only one of the active sites. The program GRID (P. J. Goodford. J. Med. Chem. 1985, 28, 849-857) was used to predict favorable binding sites for the functional groups of the aspartate ligand. These binding sites corresponded to the position of the docked aspartate ligand, indicating that substrate recognition takes place before any major conformational changes occur within the enzyme.  相似文献   

5.
In continuation of earlier work [Hu, C. Q., & Sturtevant, J.M. (1987) Biochemistry 26, 178-182], differential scanning calorimetry has been employed in a study of the effects on the thermal denaturation of yeast phosphoglycerate kinase of two inhibitors of the enzyme, sulfate ion and the dye Cibracron blue F3GA. Sulfate ion, as is usual with ligands that dissociate during unfolding of the host protein, raises t1/2, the temperature of half-completion of the denaturation, has only a modest effect, stemming from the enthalpy of dissociation of the ligand, on the enthalpy of denaturation, and has little or no effect on the heat capacity change resulting from denaturation. In sharp contrast, Cibacron blue F3GA lowers t1/2 and drastically decreases both the enthalpy and heat capacity changes due to denaturation. The DSC results with sulfate ion are consistent with previous kinetic data [Scopes, R. K. (1978) Eur. J. Biochem. 91, 119-129; Khamis, M. H., & Larsson-Raznikiewicz, M. (1981) Biochim. Biophys. Acta 657, 190-194], which indicate two binding sites for sulfate ion at one of which the ligand acts as a competitive inhibitor. The results with Cibacron blue F3GA indicate that the dye induces a major destabilizing structural change in the enzyme in addition to rendering it enzymically inactive.  相似文献   

6.
We measured time-dependent concentration changes of human interferon-alpha 2a (IFN) and human tumor necrosis factor-alpha (TNF) bound at the plasma membrane and internalized by human lung alveolar carcinoma A549 cells in the presence of excess free ligand. Concentration changes for these two ligands were substantially different. We modified our compartmental kinetic model encompassing receptor synthesis and receptor loss (Myers, A. C., Kovach, J. S., and Vuk-Pavlovi?, S. (1987) J. Biol. Chem. 262, 6494-6499) to include receptor recycling. We solved analytically the equations of three variants of the model of receptor recycling. All parameters (rate constants) were identifiable when the data sets consisted of time-resolved concentrations of IFN and TNF at the cell surface and internalized by cells. By least squares fitting we derived the best fit values for the first order rate constants for internalization of the ligand-receptor complex, receptor recycling, turnover of free receptors, elimination of the ligand from cells, and the rate of insertion of free receptors into the membrane. The best fit to data for interactions of cells with IFN was obtained without inclusion of the term for recycling of receptors to the membrane. The simplest model including receptor recycling was necessary and sufficient for the fit to the respective data for TNF. These results demonstrate that the contribution of receptor recycling to the metabolism of the ligand and the receptor can be quantitated by compartmental modeling. Receptor recycling does not contribute to the kinetics of Type I IFN receptor in A549 cells. In contrast, recycling contributes significantly to endocytosis mediated by the TNF receptor.  相似文献   

7.
8.
Transforming growth factor beta1 (TGFbeta) simultaneously induces the expression of fibronectin, fibronectin receptor, laminin, and laminin receptor (alpha6beta1 integrin) in the human colon cancer cell line Moser (Int J Cancer, 57:742, 1994). Induction of fibronectin and induction of fibronectin receptor by TGFB are tightly coupled, and disrupting fibronectin induction disrupts the induction of fibronectin receptor and cellular adhesion to fibronectin (J Cellular Physiol, 170:138, 1997). We recently demonstrated the efficacy of using antisense chain-specific laminin RNA expression vectors to disrupt the induction by TGFP of the multichain laminin molecule (J Cellular Physiol, 178:296, 1999). We now show in this report that Moser cells used alpha6 and beta1 integrins to adhere to laminin, and, as is the fibronectin and fibronectin receptor system, disrupting the induction by TGFbeta of the ligand laminin by the expression of antisense laminin A chain RNA disrupted the induction of 125I-laminin binding and cellular adhesion to laminin. Disrupting laminin induction also blocked the induction of alpha6 and beta1 integrin laminin receptor by TGFbeta. We conclude that disrupting the induction of the ligand laminin by TGFbeta disrupts TGFbeta-regulated laminin receptor function by suppressing the induction of alpha6 and beta1 integrins. Therefore, targeted disruption of the ligand laminin may be an effective means in disrupting the function of both the ligand and its receptor in cells that utilize the laminin and laminin receptor system in malignant cell behavior.  相似文献   

9.
It has been proposed that regulatory multienzyme complex formation between yeast ornithine transcarbamoylase (OTCase) and arginase is triggered by a conformational change promoted by the binding of ornithine to a regulatory site in OTCase (Wiame, J.-M. (1971) Curr. Top. Cell. Regul. 4, 1-38). To isolate the binding of ornithine to the proposed regulatory site, the active site was blocked with the high affinity (Ki = 13 +/- 1.4 nM) bisubstrate analogue, delta-N-phosphonacetyl-L-ornithine (PALO). The binding of PALO to the active site produces large changes in the absorption (delta A290-296 = 0.010/mg of enzyme) and in the fluorescence (25% quenching) of the protein. These changes both saturate at one PALO/polypeptide chain. The binding of PALO also changes the rate constant for diffusional acrylamide quenching by 43% and increases the midpoint for the thermal denaturation of the enzyme by 13 degrees C. Finally, PALO binding results in a +2.8% change in the sedimentation coefficient demonstrating that these spectral and energetic changes are associated with a gross structural change in the enzyme. In an effort to detect ligand binding to the proposed effector site on OTCase, ornithine was added to the enzyme saturated with PALO, and consequent conformational changes were tested for using methodologies identical to those which demonstrated active site ligand binding-promoted conformational changes. In no instance were any additional differences observed. Hence, strong support for isosteric effector binding-promoted conformational changes cannot be presented. We conclude that active site ligand binding events themselves are responsible for conformational changes which promote enzyme-enzyme association of OTCase with arginase.  相似文献   

10.
The block to polyspermy in Xenopus laevis involves an interaction between a cortical granule lectin, released at fertilization, and a ligand located in the egg extracellular matrix. The egg extracellular matrix in X. laevis consists of a vitelline envelope and three distinct jelly layers, designated J1, J2 and J3. To localize cortical granule lectin ligand in the egg extracellular matrix, we used enzyme-linked lectin assays that showed that cortical granule lectin ligands were absent in J2, J3 and the vitelline envelope. Cortical granule lectin bound to a ligand(s) in J1 in a galactose-dependent fashion. In addition, we separated egg jelly macromolecules electrophoretically and, in conjunction with western blotting, have shown that J1 contains two major, high molecular weight ligands for cortical granule ligand. Finally, using confocal microscopy, we demonstrated that the ligand(s) for cortical granule lectin occupies a 20–30 μm thick band in a region of J1 just proximal to the vitelline envelope.  相似文献   

11.
Incubation of alveolar macrophages or hepatocytes in media in which Na+ is replaced by K+ ("isotonic-K buffer") inhibited the movement of internalized ligand from late endosomes to lysosomes (Ward et al.: Journal of Cell Biology 110:1013-1022, 1990). In this study we investigate the mechanism responsible for the isotonic-K+ block in movement of ligand from late endosomes to lysosomes. We observed that iso-K+ inhibition of endosome-lysosome fusion is not unique to alveolar macrophages or hepatocytes but can be seen in a variety of cell types including J774 and Hela cells. The inhibition in intracellular ligand movement was time dependent with the maximum change occurring after 60 minutes. Once established the inhibition resulted in a prolonged and apparently permanent decrease in vesicle movement. Cells were able to recover from the effects of iso-K+ buffers over a time course of 5-10 minutes when placed back in Na(+)-containing media. The effect of iso-K+ buffers was independent of intracellular pH changes and appeared to involve cell swelling. When cells were incubated in iso-K+ buffers under conditions in which cell volume changes were reduced, intracellular ligand movement approached normal levels. Such conditions included replacing Cl- with the less permeant anion gluconate, and by addition of sucrose to isotonic-K+ buffers. Analysis of the mechanism by which changes in cell volume could alter intracellular movement ruled out changes in cyclic nucleotides, Ca2+, or microtubules. These results suggest that changes in cell shape or volume can alter intracellular transport systems by novel routes.  相似文献   

12.
The mechanistic basis underlying the striking cooperativity observed for the assembly of TGF-β family ligand/receptor complexes is not well understood. We report here an investigation in which we used a novel ligand sequestration assay, in combination with immunofluorescent light microscopy and flow cytometry analyses, to examine and quantify cooperative assembly of TGF-β ligand/receptor complexes on the cell surface, as well as ligand/receptor complex internalization. We analyzed the roles played by the ecto/transmembrane (ecto/TM) domains and endodomains of RI and RII TGF-β receptors in these processes by transfecting 293 or HeLa cells with different combinations of receptor mutants. We found that the ecto/TM domains of RII and RI cooperated together to promote the formation of cell surface receptor/ligand complexes. Furthermore, in agreement with the recently determined structure of the TGF-β3/RII ectodomain/RI ectodomain complex [J. Groppe, C.S. Hinck, P. Samavarchi-Tehrani, C. Zubieta, J.P. Schuermann, A.B. Taylor, P.M. Schwarz, J.L. Wrana, A.P. Hinck, Cooperative assembly of TGF-beta superfamily signaling complexes is mediated by two disparate mechanisms and distinct modes of receptor binding, Mol. Cell 29 (2008) 157–168], we observed that the N-terminus of the RII ectodomain was required for full assembly. With respect to endodomains, we found that the RI endodomain enhanced cooperative complex assembly at the cell surface, whereas both the RI and RII endodomains enhanced internalization. Finally, we observed that ligand/receptor internalization, but not complex assembly at the cell surface, was partly raft-dependent. In light of these results, currently proposed mechanisms of cooperative ligand/receptor assembly are discussed.  相似文献   

13.
The kinetics of the reaction with oxygen and carbon monoxide of the homodimeric hemoglobin from the bivalve mollusc Scapharca inaequivalvis has been extensively investigated by flash and dye-laser photolysis, temperature jump relaxation, and stopped flow methods. The results indicate that cooperativity in ligand binding, already observed for oxygen at equilibrium, finds its kinetic counterpart in a large decrease of the oxygen dissociation velocity in the second step of the binding reaction. In the case of carbon monoxide, cooperativity is clearly evident in the increase of the combination velocity constant as the reaction proceeds. Therefore, the ligand-binding kinetics of this dimeric hemoglobin shows the characteristic features of the corresponding reactions of tetrameric hemoglobins. Analysis of the data in terms of the allosteric model proposed by Monod et al. (Monod, J., Wyman, J., and Changeux, J. P. (1965) J. Mol. Biol. 12, 88-118) has shown that the values of the allosteric parameters cannot be fixed uniquely for a dimeric hemoglobin. The rapid changes in absorbance observed at the isosbestic points of unliganded and liganded hemoglobin following laser photolysis provided a value of 7 X 10(4) S-1 at 20 degrees C for the rate of the ligand-free quarternary conformational change, postulated on the basis of cooperative ligand binding. Comparison of the rapid absorbance changes observed during ligand rebinding in this hemoglobin with those observed in tuna hemoglobin indicate that, at full photolysis, binding to the T state is followed by further binding and conversion to the liganded R state; at partial photolysis, population of the liganded T state occurs immediately and is followed by a decay to the liganded R state upon further ligand binding. These new results, in conjunction with previous equilibrium data on the same system, show unequivocally that the presence of two different types of chain is not an absolute prerequisite for cooperativity in hemoglobins, contrary to currently accepted ideas.  相似文献   

14.
The Src homology 2 (SH2) domain of interleukin-2 tyrosine kinase (Itk) is a critical component of the regulatory apparatus controlling the activity of this immunologically important enzyme. To gain insight into the structural features associated with the activated form of Itk, we have solved the NMR structure of the SH2 domain bound to a phosphotyrosine-containing peptide (pY) and analyzed changes in trans-hydrogen bond scalar couplings ((3h)J(NC')) that result from pY binding. Isomerization of a single prolyl imide bond in this domain is responsible for simultaneous existence of two distinct SH2 conformers. Prolyl isomerization directs ligand recognition: the trans conformer preferentially binds pY. The structure of the SH2/pY complex provides insight into the ligand specificity; the BG loop in the ligand-free trans SH2 conformer is pre-arranged for optimal contacts with the pY+3 residue of the ligand. Analysis of (3h)J(NC') couplings arising from hydrogen bonds has revealed propagation of structural changes from the pY binding pocket to the CD loop containing conformationally heterogeneous proline as well as to the alphaB helix, on the opposite site of the domain. These findings offer a structural framework for understanding the roles of prolyl isomerization and pY binding in Itk regulation.  相似文献   

15.
Metal ions affect ligand binding to G-protein-coupled receptors by as yet unknown mechanisms. In particular, Na(+) increases the affinity for antagonists but decreases it for agonists. We had modeled the mu-opioid receptor (muR) based on the low-resolution structure of rhodopsin by G. F. X. Schertler, C. Villa, and R. Henderson (1993, Nature 362, 770-772) and proposed that metal ions may be directly involved in the binding of ligands and receptor activation (B. S. Zhorov and V. S. Ananthanarayanan, 1998, J. Biomol. Struct. Dyn. 15, 631-637). Developing this concept further, we present here homology models of muR using as templates the structure of rhodopsin elaborated by I. D. Pogozheva, A. L. Lomize, and H. I. Mosberg (1997, Biophys. J. 70, 1963-1985) and J. M. Baldwin, G. F. X. Schertler, and V. M. Unger (1997, J. Mol. Biol., 272, 144-164). Using the Monte Carlo minimization (MCM) method, we docked the Na(+)-bound forms of muR ligands: naloxone, bremazocine, and carfentanyl. The resultant low-energy complexes showed that the two positive charges in the protonated metal-bound ligands interact with the two negative charges at Asp(3.32) and Asp(2.50) (for notations, see J. A. Ballesteros and H. Weinstein, 1995, Methods Neurosci. 25, 366-426). MCM computation on morphine docked inside the model of muR by I. D. Pogozheva, A. L. Lomize, and H. I. Mosberg (1998, Biophys. J. 75, 612-634) yielded two binding modes with the ligand's ammonium group salt-bridged either to Asp(3.32) (generally regarded as the ligand recognition site) or to Asp(2.50). The latter is the presumed site for Na(+) ion, which is known to modulate ligand binding. Assuming that in the low-dielectric transmembrane region of muR, organic and inorganic cations would compete for Asp(3.32) and Asp(2.50), we propose that ligand binding, as visualized in the above models, would first displace Na(+) from Asp(3.32). A subsequent progress of the ligand toward Asp(2.50) would result in either the retention of Na(+) at Asp(2.50) in the case of antagonists or the displacement of Na(+) from Asp(2.50) in the case of agonists. The displaced Na(+) would move toward the salt-bridged Asp(3.49)-Arg(3.50) and disengage the salt bridge. This, in turn, would result in conformational changes at the cytoplasmic face of the receptor that facilitate the interaction with the G-protein.  相似文献   

16.
17.
18.
19.
W Bujalowski  T M Lohman 《Biochemistry》1987,26(11):3099-3106
We describe a general method for the analysis of ligand-macromolecule binding equilibria for cases in which the interaction is monitored by a change in a signal originating from the ligand. This method allows the absolute determination of the average degree of ligand binding per macromolecule without any assumptions concerning the number of modes or states for ligand binding or the relationship between the fractional signal change and the fraction of bound ligand. Although this method is generally applicable to any type of signal, we discuss the details of the method as it applies to the analysis of binding data monitored by a change in fluorescence of a ligand upon binding to a nucleic acid. We apply the analysis to the equilibrium binding of Escherichia coli single-strand binding (SSB) protein to single-stranded nucleic acids, which is monitored by the quenching of the intrinsic tryptophan fluorescence of the SSB protein. With this method, one can quantitatively determine the relationship between the fractional signal change of the ligand and the fraction of bound ligand, LB/LT, and rigorously test whether the signal change is directly proportional to LB/LT. For E. coli SSB protein binding to single-stranded nucleic acids in its (SSB)65 binding mode [Lohman, T. M., & Overman, L. B. (1985) J. Biol. Chem. 260, 3594; Chrysogelos, S., & Griffith, J. (1982) Proc. Natl. Acad. Sci. U.S.A. 79, 5803], we show that the fractional quenching of the SSB fluorescence is equal to the fraction of bound SSB.  相似文献   

20.
Starting from the Monod-Wyman-Changeux (MWC) model (Monod, J., J. Wyman, and J. P. Changeux. 1965. J. Mol. Biol. 12:88-118), we obtain an analytical expression for the slope of the Hill plot at any ligand concentration. Furthermore, we derive an equation satisfied by the ligand concentration at the position of maximum slope. From these results, we derive a set of formulas which allow determination of the parameters of the MWC model (kR, C, and L) from the value of the Hill coefficient, nH, the ligand concentration at the position of maximum slope [( A]0), and the value of nu/(n-nu) at this point. We then outline procedures for utilizing these equations to provide a "best fit" of the MWC model to the experimental data, and to obtain a refined set of the parameters. Finally, we demonstrate the applicability of the technique by analysis of oxygen binding data for Octopus hemocyanin.  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