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 828 毫秒
1.
Nieland TJ  Xu S  Penman M  Krieger M 《Biochemistry》2011,50(11):1818-1830
Scavenger receptor class B, type I (SR-BI), is a high-density lipoprotein (HDL) receptor, which also binds low-density lipoprotein (LDL), and mediates the cellular selective uptake of cholesteryl esters from lipoproteins. SR-BI also is a coreceptor for hepatitis C virus and a signaling receptor that regulates cell metabolism. Many investigators have reported that lipoproteins bind to SR-BI via a single class of independent (not interacting), high-affinity binding sites (one site model). We have reinvestigated the ligand concentration dependence of (125)I-HDL binding to SR-BI and SR-BI-mediated specific uptake of [(3)H]CE from [(3)H]CE-HDL using an expanded range of ligand concentrations (<1 μg of protein/mL, lower than previously reported). Scatchard and nonlinear least-squares model fitting analyses of the binding and uptake data were both inconsistent with a single class of independent binding sites binding univalent lipoprotein ligands. The data are best fit by models in which SR-BI has either two independent classes of binding sites or one class of sites exhibiting negative cooperativity due to either classic allostery or ensemble effects ("lattice model"). Similar results were observed for LDL. Application of the "infinite dilution" dissociation rate method established that the binding of (125)I-HDL to SR-BI at 4 °C exhibits negative cooperativity. The unexpected complexity of the interactions of lipoproteins with SR-BI should be taken into account when interpreting the results of experiments that explore the mechanism(s) by which SR-BI mediates ligand binding, lipid transport, and cell signaling.  相似文献   

2.
3.
Integrins are heterodimeric adhesion receptors that regulate immune cell adhesion. Integrin-dependent adhesion is controlled by multiple conformational states that include states with different affinity to the ligand, states with various degrees of molecule unbending, and others. Affinity change and molecule unbending play major roles in the regulation of cell adhesion. The relationship between different conformational states of the integrin is unclear. Here we have used conformationally sensitive antibodies and a small LDV-containing ligand to study the role of the inside-out signaling through formyl peptide receptor and CXCR4 in the regulation of α4β1 integrin conformation. We found that in the absence of ligand, activation by formyl peptide or SDF-1 did not result in a significant exposure of HUTS-21 epitope. Occupancy of the ligand binding pocket without cell activation was sufficient to induce epitope exposure. EC50 for HUTS-21 binding in the presence of LDV was identical to a previously reported ligand equilibrium dissociation constant at rest and after activation. Furthermore, the rate of HUTS-21 binding was also related to the VLA-4 activation state even at saturating ligand concentration. We propose that the unbending of the integrin molecule after guanine nucleotide-binding protein-coupled receptor-induced signaling accounts for the enhanced rate of HUTS-21 binding. Taken together, current results support the existence of multiple conformational states independently regulated by both inside-out signaling and ligand binding. Our data suggest that VLA-4 integrin hybrid domain movement does not depend on the affinity state of the ligand binding pocket.In the bloodstream circulating leukocytes respond to inflammatory signals by rapid changes of cell adhesive properties. These include cell tethering, rolling, arrest, and firm adhesion, all of which are well described steps of leukocyte recruitment to the sites of inflammation (1). Leukocyte arrest and firm adhesion are mediated exclusively by integrin receptors (2). At the same time integrins can also mediate tethering and rolling (3). These largely diverse cell adhesive properties are achieved by sophisticated conformational regulation; multiple states of the same molecule with different affinity for its ligand and different degrees of molecular unbending are attributed to various types of “cellular behavior.” It is proposed that the low affinity bent state translates into a non-adhesive resting cell, the low affinity unbent or extended state of integrin results in cell rolling, and the high affinity state promotes cell arrest (4, 5). However, the exact sequence of conformational events and the relationship between integrin conformational and functional activity remain key questions (6).Integrin conformation is regulated through G-protein-coupled receptors by a signaling pathway which is initiated by ligand binding to a GPCR,3 propagated inside the cell, and results in the binding of signaling proteins (such as talin and others) to cytoplasmic domains of integrin subunits. This binding leads to a separation of the integrin cytoplasmic domains and inside-out activation (6). Chemokines (chemotactic cytokines) as well as “classical” chemoattractants (such as formyl peptide) preferentially signal through heterotrimeric G-proteins coupled to the Gαi subunit (1). Activation by these ligands results in up-regulation of integrin affinity and/or conformational unbending (extension) of the integrin molecule. These conformational changes lead to cell arrest and firm adhesion. G-protein receptors coupled to Gαs-coupled subunit (adenylyl cyclase/cAMP signaling pathway) can actively down-regulate the affinity state of the ligand binding pocket without changing integrin conformational unbending. This provides an anti-adhesive signal and results in cell de-adhesion (7). Thus, interaction of multiple G-protein-coupled receptors on a single cell creates a plethora of conformational states. Understanding of the relationship between inside-out signaling through GPCRs and integrin conformational regulation will provide valuable insight into the dynamic regulation of cell adhesion.One technique to study conformational changes of integrins uses conformationally sensitive mAbs that bind to epitopes which are hidden in one conformation and exposed under certain conditions. Lately, it has been accepted that integrins exhibit two major conformations, resting and activated. A number of mAbs for “activated” integrins have been described, and the epitopes have been mapped. Together with mapping of these epitopes into three-dimensional structures of integrin (8), epitope exposure can provide helpful information about integrin conformational changes upon signaling. Moreover, because integrin inside-out activation through different signaling pathways can result in different activation states, the use of previously mapped mAbs can help dissect conformational changes upon activation.Although it is clear that inside-out activation results in a conformational rearrangement of the integrin molecule, the relationship between affinity state of the ligand binding pocket and overall molecule conformation is still debated. Currently, two contrasting models of integrin inside-out integrin activation are described. The “switchblade” model implies that an open head structure with swung-out β-hybrid domain represents the high (or at least intermediate) affinity state. A feature of this model is that integrin extension provides space for hybrid domain swing. The “deadbolt” model proposes that the movement of β-hybrid domain is not related to the inside-out signal. Ligand binding by itself can provide the energy for the hybrid domain swing out (for details, see Ref. 9 and references therein). Because these two models assign different roles to the hybrid domain motion, we evaluated the exposure of VLA-4 hybrid domain epitopes upon activation through two Gαi-coupled GPCRs (FPR and CXCR4) and ligand binding using the conformationally sensitive HUTS-21 mAb with an epitope mapped to the hybrid domain of β1-integrin (10).We found that contrary to previous reports, where these mAbs were reported to bind or used for the detection of activated integrin (1013), formyl peptide or SDF-1 treatment alone did not result in any significant exposure of HUTS-21 epitope despite the fact that the VLA-4 affinity up-regulation was detected in parallel on the same batch of cells. Quantitative analysis of mAb binding in real time on live cells suggests that for both the low (resting) and high affinity (induced by inside-out pathway) states, occupancy of the ligand binding pocket rather than inside-out signaling by itself causes the conformational change. Thus, these data support the idea that the hybrid domain movement, which results in the exposure of the mAb epitope, and the high affinity state of the binding pocket are regulated separately and independently of each other, a feature of the deadbolt model of inside-out activation.  相似文献   

4.
5.
G M Keserü  D K Menyhárd 《Biochemistry》1999,38(20):6614-6622
Monte Carlo protein simulations with continuum solvation were used to explore the conformational mobility of NO within the active site of metmyoglobin. To the best of our knowledge this is the first application of a continuum solvation model for exploring protein binding sites. The usefulness of the Monte Carlo conformational analysis was demonstrated in comparative molecular dynamics simulations. Analysis of conformer populations revealed that Monte Carlo conformational analysis is more effective in mapping the relevant region of the potential surface. Distribution of low-energy conformers obtained for the metmyoglobin-NO complex was found to depend on the orientation of proximal His93. Different charge distributions corresponding to the two experimentally verified possible torsions of this proximal residue result in strong binding of NO or its release to a nearby hydrophobic trap. Conformer populations obtained by Monte Carlo conformational analysis were grouped into three main families: one, with the NO directly bound to the iron, appears when the CA-CB-CG-CD2 torsion of His93 is at its ligand binding value (-113 degrees); and two conformers exist where NO is trapped in a nearby hydrophobic pocket, the same cavity that was determined to be the geminate trap of CO in ferrous Mb as a result of the torsional flip of His93 to its ligand releasing state (-125 degrees). Based on this analysis, we suggest that the electrostatic rearrangement coupled to the conformational fluctuation of the proximal His leads to the geminate trapping of the ligand. Conformational rearrangement of the proximal side would provide the possibility of rebinding of the ligand to Fe.  相似文献   

6.
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   

7.
We have mathematically analyzed ligand-induced monomerization and dimerization in a protein monomer-dimer equilibrium system, in which the monomer has one and the dimer two binding sites. These dimer sites have the same binding constants for the first ligand but may cooperatively interact when one of them is occupied by a ligand molecule. In this system, the apparent dimerization constant and the apparent molecular weight are functions of free ligand concentration, and depend on the intrinsic binding constants of the ligand molecule to the monomer and the dimer. The behavior of these functions is classified into 17 cases according to the values of the three intrinsic binding constants, and some calculated examples are shown graphically for selected parameters. The theory was also applied to D-amino acid oxidase [EC 1.4.3.3], a flavoprotein, and the pH dependence of the apparent dimerization constant and the apparent molecular weight in the presence of ligand, p-aminobenzoate, were studied theoretically using parameters obtained in our previous experiments (5).  相似文献   

8.
Homo- and heterodimerization have emerged as prominent features of G-protein-coupled receptors with possible impact on the regulation of their activity. Using a sensitive bioluminescence resonance energy transfer system, we investigated the formation of CXCR4 and CCR2 chemokine receptor dimers. We found that both receptors exist as constitutive homo- and heterodimers and that ligands induce conformational changes within the pre-formed dimers without promoting receptor dimer formation or disassembly. Ligands with different intrinsic efficacies yielded distinct bioluminescence resonance energy transfer modulations, indicating the stabilization of distinct receptor conformations. We also found that peptides derived from the transmembrane domains of CXCR4 inhibited activation of this receptor by blocking the ligand-induced conformational transitions of the dimer. Taken together, our data support a model in which chemokine receptor homo- and heterodimers form spontaneously and respond to ligand binding as units that undergo conformational changes involving both protomers even when only one of the two ligand binding sites is occupied.  相似文献   

9.
The calcium binding properties of annexin I as observed by thermodynamic DSC studies have been compared to the structural information obtained from X-ray investigation. The calorimetric experiment permitted to evaluate both the reaction scheme - including binding of ligand and conformational changes - and the energetics of each reaction step. According to published X-ray data Annexin I has six calcium binding sites, three medium-affinity type II and three low-affinity type III sites.The present study shows that at 37 degrees C annexin I binds in a Hill type fashion simultaneously two calcium ions in a first step with medium affinity at a concentration of 0.6 mM and another three Ca(2+) ions again cooperatively at 30 mM with low affinity. Therefore it can be concluded that only two medium-affinity type II binding sites are available. The third site, that should be accessible in principle appears to be masked presumably due to the presence of the N terminus. In view of the large calcium concentration needed for saturation of the binding sites, annexin I may be expected to be Ca(2+) free in vivo unless other processes such as membrane interaction occur simultaneously. This assumption is consistent with the finding, that the affinity of annexins to calcium is usually markedly increased by the presence of lipids.  相似文献   

10.
Mosberg HI 《Biopolymers》1999,51(6):426-439
The elaboration of a pharmacophore model for the delta opioid receptor selective ligand JOM-13 (Tyr-c[D-Cys-Phe-D-Pen]OH) and the parallel, independent development of a structural model of the delta receptor are summarized. Although the backbone conformation of JOM-13's tripeptide cycle is well defined, considerable conformational lability is evident in the Tyr(1) residue and in the Phe(3) side chain, key pharmacophore elements of the ligand. Replacement of these flexible features of the ligand by more conformationally restricted analogues and subsequent correlation of receptor binding and conformational properties allowed the number of possible binding conformations of JOM-13 to be reduced to two. Of these, one was chosen as more likely, based on its better superposition with other conformationally constrained delta receptor ligands. Our model of the delta opioid receptor, constructed using a general approach that we have developed for all rhodopsin-like G protein-coupled receptors, contains a large cavity within the transmembrane domain that displays excellent complementarity in both shape and polarity to JOM-13 and other delta ligands. This binding pocket, however, cannot accommodate the conformer of JOM-13 preferred from analysis of ligands, alone. Rather, only the "alternate" allowed conformer, identified from analysis of the ligands but "disfavored" because it does not permit simultaneous superposition of all pharmacophore elements of JOM-13 with other delta ligands, fits the binding site. These results argue against a simple view of a single, common fit to a receptor binding site and suggest, instead, that at least some binding site interactions of different ligands may differ.  相似文献   

11.
12.
We have devised a molecular model building computer program (DEZYMER) which builds new ligand binding sites into a protein of known three-dimensional structure. It alters only the sequence and the side-chain structure of the protein, leaving the protein backbone fold intact by definition. The program searches for a constellation of backbone positions arranged such that if appropriate side-chains were placed there, they would bind the ligand according to a pre-defined geometry of interaction specified by the experimentalist. These binding sites are introduced by the program by taking into account simple rules such as steric hindrance, atomic close-packing and hydrogen bond patterns, which are known to maintain the integrity of a protein structure to a first approximation. A test case is presented in this paper where the copper binding site found in blue-copper proteins such as plastocyanin, azurin and cupredoxin is introduced into Escherichia coli thioredoxin. The model building of one of the solutions found by the program is presented in some detail. The experimental construction and properties of this new protein are described in an accompanying paper. It is hoped that this program provides a general method for the design of ligand binding sites and enzyme active sites, which can then be tested experimentally.  相似文献   

13.

Background  

Riboswitches are a type of noncoding RNA that regulate gene expression by switching from one structural conformation to another on ligand binding. The various classes of riboswitches discovered so far are differentiated by the ligand, which on binding induces a conformational switch. Every class of riboswitch is characterized by an aptamer domain, which provides the site for ligand binding, and an expression platform that undergoes conformational change on ligand binding. The sequence and structure of the aptamer domain is highly conserved in riboswitches belonging to the same class. We propose a method for fast and accurate identification of riboswitches using profile Hidden Markov Models (pHMM). Our method exploits the high degree of sequence conservation that characterizes the aptamer domain.  相似文献   

14.
Weikl TR  von Deuster C 《Proteins》2009,75(1):104-110
The binding of a ligand molecule to a protein is often accompanied by conformational changes of the protein. A central question is whether the ligand induces the conformational change (induced-fit), or rather selects and stabilizes a complementary conformation from a pre-existing equilibrium of ground and excited states of the protein (selected-fit). We consider here the binding kinetics in a simple four-state model of ligand-protein binding. In this model, the protein has two conformations, which can both bind the ligand. The first conformation is the ground state of the protein when the ligand is off, and the second conformation is the ground state when the ligand is bound. The induced-fit mechanism corresponds to ligand binding in the unbound ground state, and the selected-fit mechanism to ligand binding in the excited state. We find a simple, characteristic difference between the on- and off-rates in the two mechanisms if the conformational relaxation into the ground states is fast. In the case of selected-fit binding, the on-rate depends on the conformational equilibrium constant, whereas the off-rate is independent. In the case of induced-fit binding, in contrast, the off-rate depends on the conformational equilibrium, while the on-rate is independent. Whether a protein binds a ligand via selected-fit or induced-fit thus may be revealed by mutations far from the protein's binding pocket, or other "perturbations" that only affect the conformational equilibrium. In the case of selected-fit, such mutations will only change the on-rate, and in the case of induced-fit, only the off-rate.  相似文献   

15.
Wako H  Endo S 《Biophysical chemistry》2011,159(2-3):257-266
The conformational change of a protein upon ligand binding was examined by normal mode analysis (NMA) based on an elastic-network model (ENM) for a full-atom system using dihedral angles as independent variables. Specifically, we investigated the extent to which conformational change vectors of atoms from an apo form to a holo form of a protein can be represented by a linear combination of the displacement vectors of atoms in the apo form calculated for the lowest-frequency m normal modes (m=1, 2,…, 20). In this analysis, the latter vectors were best fitted to the former ones by the least-squares method. Twenty-two paired proteins in the holo and apo forms, including three dimer pairs, were examined. The results showed that, in most cases, the conformational change vectors were reproduced well by a linear combination of the displacement vectors of a small number of low-frequency normal modes. The conformational change around an active site was reproduced as well as the entire conformational change, except for some proteins that only undergo significant conformational changes around active sites. The weighting factors for 20 normal modes optimized by the least-squares fitting characterize the conformational changes upon ligand binding for these proteins. The conformational changes sampled around the apo form of a protein by the linear combination of the displacement vectors obtained by ENM-based NMA may help solve the flexible-docking problem of a protein with another molecule because the results presented herein suggest that they have a relatively high probability of being involved in an actual conformational change.  相似文献   

16.
TLR3 (Toll-like receptor 3) recognizes dsRNA, a potent indicator of viral infection. The extracellular domain of TLR3 dimerizes when it binds dsRNA, and the crystal structure of the dimeric complex reveals three sites of interaction on each extracellular domain, two that bind dsRNA and one that is responsible for dimer formation. The goal of this study was to determine which amino acid residues are essential for forming a stable receptor·ligand complex and whether dimerization of TLR3 is required for dsRNA binding. Using a novel ELISA to analyze dsRNA binding by mutant TLR3 constructs, we identified the essential interacting residues and determined that the simultaneous interaction of all three sites is required for ligand binding. In addition, we show that TLR3 is unable to bind dsRNA when dimerization is prevented by mutating residues in the dimerization site or by immobilizing TLR3 at low density. We conclude that dimerization of TLR3 is essential for ligand binding and that the three TLR3 contact sites individually interact weakly with their binding partners but together form a high affinity receptor·ligand complex.  相似文献   

17.
18.
Hellmann N 《IUBMB life》2011,63(5):329-336
Cooperative effects in ligand binding and dissociation kinetics are much less investigated than steady state kinetics or equilibrium binding. Nevertheless, cooperativity in ligand binding leads necessarily to characteristic properties with respect to kinetic properties of the system. In case of positive cooperativity as found in oxygen binding proteins, a typical property is an autocatalytic ligand dissociation behavior leading to a time dependent, apparent ligand dissociation rate. To follow systematically the influence of the various potentially involved parameters on this characteristic property, simulations based on the simple MWC model were performed which should be relevant for all types of models based on the concept of an allosteric unit. In cases where the initial conformational distribution is very much dominated by the R-state, the intrinsic kinetic properties of the T-state are of minor influence for the observed ligand dissociation rate. Even for fast conformational transition rates, the R-state properties together with the size of the allosteric unit and the allosteric equilibrium constant define the shape of the curve. In such a case, a simplified model of the MWC-scheme (the irreversible n-chain model) is a good approximation of the full scheme. However, if in the starting conformational distribution some liganded T-molecules are present (a few percent is enough), the average off-rates can be significantly altered. Thus, the assignment of the initial rates to R-state properties has to be done with great care. However, if the R-state strongly dominates initially it is even possible to get an estimation of the lower limit for the number of interacting subunits from kinetic data: similar to the Hill-coefficient for equilibrium conditions, a measure for "kinetic cooperativity" can be derived by comparing initial and final ligand dissociation rates.  相似文献   

19.
D Watters  A Maelicke 《Biochemistry》1983,22(8):1811-1819
We have studied 20 monoclonal antibodies directed against both the solubilized and the membrane-bound receptor from Torpedo marmorata. We find the following: (i) Six of the antibodies compete with cholinergic ligands for receptor binding and, hence, are directed against the ligand binding regions. (ii) Of these six antibodies, two cross-react with receptor from Electrophorus electricus, rat myotubes, and chicken sympathetic ganglia. These two antibodies therefore define a preserved structure within the ligand binding regions. The other four antibodies bind to structures not common between the receptor preparations tested. (iii) From competition binding studies using internally 3H-labeled antibodies, nine nonoverlapping antigenic regions were defined at the surface of the receptor. Three of these regions overlap with the ligand binding regions. Since two of these three regions do not overlap with each other, two structurally distinct ligand binding regions must exist at the receptor. (iv) From competition binding studies with representative cholinergic ligands, the antibodies directed against the ligand binding regions can be subdivided into three groups: one group competes with all ligands tested; the second group competes with all ligands except the bismethonium compounds; the third group competes with all ligands except the bismethonium compounds and tubocurarine. The results are summarized in a model of the organization of ligand binding sites at the receptor: There are two ligand binding regions differing in their antigenic properties. Furthermore, either there exists separate sites for distinct groups of ligands within each of these binding regions or some ligands produce conformational changes of the receptor that reversibly abolish some antigenic sites. In any case, the cholinergic ligands must interact with the receptor by more and/or other structural determinants than are provided by the structure of acetylcholine.  相似文献   

20.
Proteins are dynamic molecules and often undergo conformational change upon ligand binding. It is widely accepted that flexible loop regions have a critical functional role in enzymes. Lack of consideration of binding site flexibility has led to failures in predicting protein functions and in successfully docking ligands with protein receptors. Here we address the question: which sequence and structural features distinguish the structurally flexible and rigid binding sites? We analyze high-resolution crystal structures of ligand bound (holo) and free (apo) forms of 41 proteins where no conformational change takes place upon ligand binding, 35 examples with moderate conformational change, and 22 cases where a large conformational change has been observed. We find that the number of residue-residue contacts observed per-residue (contact density) does not distinguish flexible and rigid binding sites, suggesting a role for specific interactions and amino acids in modulating the conformational changes. Examination of hydrogen bonding and hydrophobic interactions reveals that cases that do not undergo conformational change have high polar interactions constituting the binding pockets. Intriguingly, the large, aromatic amino acid tryptophan has a high propensity to occur at the binding sites of examples where a large conformational change has been noted. Further, in large conformational change examples, hydrophobic-hydrophobic, aromatic-aromatic, and hydrophobic-polar residue pair interactions are dominant. Further analysis of the Ramachandran dihedral angles (phi, psi) reveals that the residues adopting disallowed conformations are found in both rigid and flexible cases. More importantly, the binding site residues adopting disallowed conformations clustered narrowly into two specific regions of the L-Ala Ramachandran map. Examination of the dihedral angles changes upon ligand binding shows that the magnitude of phi, psi changes are in general minimal, although some large changes particularly between right-handed alpha-helical and extended conformations are seen. Our work further provides an account of conformational changes in the dihedral angles space. The findings reported here are expected to assist in providing a framework for predicting protein-ligand complexes and for template-based prediction of protein function.  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