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.
Polar lipids and membrane proteins are major components of biological membranes, both cell membranes and membranes of enveloped viruses. How these two classes of membrane components interact with each other to influence the function of biological membranes is a fundamental question that has attracted intense interest since the origins of the field of membrane studies. One of the most powerful ideas that driven the field is the likelihood that lipids bind to membrane proteins at specific sites, modulating protein structure and function. However only relatively recently has high resolution structure determination of membrane proteins progressed to the point of providing atomic level structure of lipid binding sites on membrane proteins. Analysis of X-ray diffraction, electron crystallography and NMR data over 100 specific lipid binding sites on membrane proteins. These data demonstrate tight lipid binding of both phospholipids and cholesterol to membrane proteins. Membrane lipids bind to membrane proteins by their headgroups, or by their acyl chains, or binding is mediated by the entire lipid molecule. When headgroups bind, binding is stabilized by polar interactions between lipid headgroups and the protein. When acyl chains bind, van der Waals effects dominate as the acyl chains adopt conformations that complement particular sites on the rough protein surface. No generally applicable motifs for binding have yet emerged. Previously published biochemical and biophysical data link this binding with function. This Article is Part of a Special Issue Entitled: Membrane Structure and Function: Relevance in the Cell's Physiology, Pathology and Therapy.  相似文献   

2.
Ligand binding to proteins: the binding landscape model.   总被引:1,自引:3,他引:1       下载免费PDF全文
Models of ligand binding are often based on four assumptions: (1) steric fit: that binding is determined mainly by shape complementarity; (2) native binding: that ligands mainly bind to native states; (3) locality: that ligands perturb protein structures mainly at the binding site; and (4) continuity: that small changes in ligand or protein structure lead to small changes in binding affinity. Using a generalization of the 2D HP lattice model, we study ligand binding and explore these assumptions. We first validate the model by showing that it reproduces typical binding behaviors. We observe ligand-induced denaturation, ANS and heme-like binding, and "lock-and-key" and "induced-fit" specific binding behaviors characterized by Michaelis-Menten or more cooperative types of binding isotherms. We then explore cases where the model predicts violations of the standard assumptions. For example, very different binding modes can result from two ligands of identical shape. Ligands can sometimes bind highly denatured states more tightly than native states and yet have Michaelis-Menten isotherms. Even low-population binding to denatured states can cause changes in global stability, hydrogen-exchange rates, and thermal B-factors, contrary to expectations, but in agreement with experiments. We conclude that ligand binding, similar to protein folding, may be better described in terms of energy landscapes than in terms of simpler mass-action models.  相似文献   

3.
Binding of nucleic acids to the prion protein (PrP) created a conundrum that required distinguishing between non-specific interactions and biologically important polynucleotides. In the process of developing selective ligands for PrP, we found using a single-stranded DNA thioaptamer library that the binding of thioaptamers to PrP occurs on at least two different sites on the protein. Selection against recombinant (rec) PrP of Syrian hamster (SHa) sequence 90-231 folded into an alpha-helical-rich conformation identified a 12-base consensus sequence within a series of 20 thioaptamers, all of which consist of 40 bases. Each thioaptamer was comprised of both normal and thio-dA modified bases. One thioaptamer designated 97 bound to recSHaPrP with affinity of 0.58(+/-0.1) nM; lower affinities for bovine (Bo), and human (Hu) were found, establishing that binding is dependent on the primary structure of PrP. High affinity binding of thioaptamer 97 to PrP was found to be mediated through the dodecyl sequence GACACAAGCCGA within the consensus region with five critical backbone modifications 5' to each dA residue. A control oligonucleotide with an equivalent number of phosphorothioates to thioaptamer 97 and a scrambled consensus sequence could not distinguish among the three PrP sequences. Control oligonucleotides bearing non-selected sequences bound to PrP at a sequence-independent DNA-binding site. In contrast, the high-affinity binding of thioaptamer 97 to PrP depends on (1) backbone modifications, (2) oligonucleotide sequence, and (3) PrP sequence.  相似文献   

4.
An experimental study of kinetics of peptide binding to MHC proteins [S. Witt, H. McConnell, Acc. Chem. Res. 26 (1993) 442 (and references therein)] showed an unusual phenomenon of the so-called `negative' t1/2 plots (where t1/2 is half time of reaching equilibrium concentration of peptide–protein complexes), which is the shorter t1/2 at lower added peptide concentrations. The bell-shaped curve for t1/2 as a function of peptide concentration is a seemingly peculiar effect, because in general, binding reactions go faster with an increase of reagent concentrations. It is shown that the suggested explanation of this phenomenon [S. Witt, H. McConnell, Acc. Chem. Res. 26 (1993) 442 (and references therein)] is misleading (and the numerical simulation used to support this explanation is inconsistent with the experimental data), so that the existence of `negative' t1/2 is in no way to be considered as an experimental indication of the two-step reaction of peptide binding to protein. This article gives a consistent explanation of the bell-shaped t1/2 plots for peptide–protein association and obtains the criteria for its existence, based exclusively on the formal chemical kinetics analysis of the accepted peptide–protein binding model and briefly discusses the experimental data, which really confirm the existence of the two-step mechanism of binding. The analysis of the maximum location of the half-time curves indicates a controversy between the prediction of the two-step binding model and the experimental data [S. Witt, H. McConnell, Acc. Chem. Res. 26 (1993) 442 (and references therein)]: either more complicated mechanism is involved in peptide–MHC binding or experimental data [S. Witt, H. McConnell, Acc. Chem. Res. 26 (1993) 442 (and references therein)] are not quite accurate.  相似文献   

5.
The kinetics of haem binding to human serum albumin and haemopexin were studied by means of the stopped flow technique. The reaction could be divided into three kinetically clearly distinguished steps: (1) extremely fast reaction of haem with nonspecific binding sites on the surface of the apoprotein molecule; this type of haem binding site seems to exist in proteins in general; (2) by meaas of equilibrium with its monomer, haem is transferred to the specific binding site; this second order reaction takes about 1–2 s, the reaction rate constant amounts to ≈106 l mol?1 s?1 both for albumin and haemopexin: (3) conformational changes of haemoprotein molecule, accompanied by changes of absorption spectra in the Soret region; this series of slow monomolecular reactions takes about 20 min. These results are discussed in connection with the mechanism of haem transport from blood to liver cells.  相似文献   

6.
Noncovalent binding of the synthetic peptide RS20 to calmodulin in the presence of calcium was confirmed by electrospray ionization coupled with Fourier transform ion cyclotron resonance mass spectrometry to form a complex with a 1:1:4 calmodulin/RS20/calcium stoichiometry. There was no evidence for formation of a calmodulin-RS20-Ca(2) species. The absence of calmodulin-RS20-Ca(2) would be consistent with models in which the two globular domains are coupled functionally. There was evidence that calmodulin, RS20-calmodulin without associated calcium, and calmodulin-RS20-Ca(4) existed together in solution, whereas calmodulin-calcium complexes were absent. It is proposed that calcium binding to form the calmodulin-RS20-Ca(4) complex occurs after an initial RS20-calmodulin binding event, and serves to secure the target within the calmodulin structure. The binding of more than one RS20 molecule to calmodulin was observed to induce unfolding of calmodulin.  相似文献   

7.
This study describes the establishment of a peptide-binding assay for purified, detergent-solubilized DR molecules. For each of the DR specificities and peptides studied, a unique pattern of interaction was observed. Excellent correlation was detected between the DR1-, 2-, 5-, and 52a-binding capacities and the known DR restrictions of a panel of synthetic peptides. This supports the immunologic relevance of the binding assay, and emphasizes the importance of determinant selection in defining the immune response of individuals. We have also examined the capacity of a panel of DR-restricted peptides to compete with one another for binding to DR1. The results obtained are compatible with a single peptide-binding site on DR molecules. The peptide-binding capacity of the four different DR types (DR1, DR2, DR5, and DR52a) has been further examined by testing a collection of 133 different peptides. This collection is unbiased with respect to previously known DR binding and restrictions, and includes peptides of eukaryotic, bacterial, and viral origins. It was found that: 1) approximately 15 to 35% of the peptides tested bound any given DR type; 2) DR-binding capacities appeared to correlate with each other, suggesting that different alleles of the DR isotype may recognize related structures on an Ag molecule; and 3) despite the statistical correlation between binding capacity of different DR types, approximately 50% of the peptides that were positive binders still were specific in that they could bind only one of the four DR molecules tested. Degenerate binding (i.e., binding to most or all the DR molecules tested) was detected in only a minority of the cases analyzed (approximately 25%).  相似文献   

8.
The quantitative binding of a phenothiazine drug to calmodulin, calmodulin fragments, and structurally related calcium binding proteins was measured under conditions of thermodynamic equilibrium by using a gel filtration method. Plant and animal calmodulins, troponin C, S100 alpha, and S100 beta bind chlorpromazine in a calcium-dependent manner with different stoichiometries and affinities for the drug. The interaction between calmodulin and chlorpromazine appears to be a complex, calcium-dependent phenomenon. Bovine brain calmodulin bound approximately 5 mol of drug per mol of protein with apparent half-maximal binding at 17 microM drug. Large fragments of calmodulin had limited ability to bind chlorpromazine. The largest fragment, containing residues 1-90, retained only 5% of the drug binding activity of the intact protein. A reinvestigation of the chlorpromazine inhibition of calmodulin stimulation of cyclic nucleotide phosphodiesterase further indicated a complex, multiple equilibrium among the reaction components and demonstrated that the order of addition of components to the reaction altered the drug concentration required for half-maximal inhibition of the activity over a 10-fold range. These results confirm previous observations using immobilized phenothiazines [Marshak, D.R., Watterson, D.M., & Van Eldik, L.J. (1981) Proc. Natl. Acad. Sci. U.S.A. 78, 6793-6797] that indicated a subclass of calcium-modulated proteins bound phenothiazines in a calcium-dependent manner, demonstrate that the interaction between phenothiazines and calmodulin is more complex than previously assumed, and suggest that extended regions of the calmodulin molecule capable of forming the appropriate conformation are required for specific, high-affinity, calcium-dependent drug binding activity.  相似文献   

9.
The effects of mastoparan and compound 48/80 on the activities of alpha beta gamma-trimeric GTP-binding proteins (G proteins) were studied with purified Go and Gi-1 which had been reconstituted into phospholipid vesicles. Pertussis toxin-catalyzed ADP-ribosylation of Go or Gi-1 was inhibited by mastoparan or compound 48/80, suggesting that the G proteins were dissociated into their constituent alpha- and beta gamma-subunits in the presence of these compounds. The steady-state rate of GTP hydrolysis catalyzed by Go or Gi-1 was stimulated by the two compounds. Both the stimulations were due to increases in the rate of the GDP-GTP exchange reaction occurring on the G proteins. However, the modes stimulation of the GTPase activity depended on the type of G protein used, and the stimulations caused by the two compounds were differently affected by pertussis toxin-catalyzed ADP-ribosylation of G proteins. Moreover, the mastoparan-induced stimulation of the GTPase activity was partially inhibited by compound 48/80. Thus, the two histamine secretagogues mastoparan and compound 48/80 appear to activate G proteins differently, though they interact with the signal-transducing proteins, at least partly, at a common binding site.  相似文献   

10.
Host binding proteins and bacterial adhesion: ecology and binding model   总被引:1,自引:0,他引:1  
Defining the involvement of specific recognition and (or) adhesion molecules in the precise association formed between cells of an organism during development or between bacteria and specific host tissues has become a focus of extensive research. The possibility that the same molecules responsible for cellular adhesion in the host may also play a major role in determining host-bacterial interactions is now becoming more evident. The following review looks at the interaction of a group of host binding proteins, including lectins, fibronectin, and laminin, with respect to their specific association with bacteria. This information is dealt with both from the perspective of the ecology of the host and its autochthonous and pathogenic bacterial populations, as well as in terms of the difficulties in defining the nature of ligand associations even in the more simplified bacterial-host interaction.  相似文献   

11.
A series of synthetic peptides have been studied as models for non-specific protein-DNA interactions. In an alpha-helical conformation, the charged amino acid residues of the N-terminal 24 residues of RecA protein are asymmetrically distributed; at neutral pH there is a +4 charge on one face of the helix and a -3 charge on the other face. Modeling suggests that the positive face of the helix can bind five DNA phosphate groups by electrostatic interactions. Circular dichroism (c.d.) spectra indicate that the analogous peptide, Rec24 (AIDENKQKALAAALGQIEKQFGKG-amide), is largely unstructured in water but becomes highly helical in the presence of DNA. Peptide titrations of fluorescent etheno-DNA confirm that the changes in the c.d. spectrum of the peptide are associated with binding, although a dependence of the c.d. signal on the degree of DNA saturation is observed, indicating that peptide can be bound in more than one conformation. At saturation the peptide binds to 5.0(+/- 0.5) DNA phosphate groups as predicted and the electrostatic nature of the binding is confirmed by a strong dependence on salt concentration. A "mutant" peptide where an acidic glutamate residue replaces an alanine on the basic face of the Rec24 helix exhibits weaker binding to single-stranded DNA, also consistent with the electrostatic nature of the proposed peptide-DNA interaction. Extending Rec24 by ten amino acid residues, where the additional residues do not participate in the helical motif, does not noticeably affect binding. Thus, we show experimentally that an asymmetric charge distribution on an alpha-helix can represent an important element for binding nucleic acids.  相似文献   

12.
The interactions of long chain fatty acids (FA) with wild type (WT) fatty acid binding proteins (FABP) and engineered FABP mutants have been monitored to determine the equilibrium binding constants as well as the rate constants for binding and dissociation. These measurements have been done using the fluorescent probes, ADIFAB and ADIFAB2, that allow the determination of the free fatty acid (FFA) concentration in the reaction of FA with proteins and membranes. The results of these studies indicate that for WT proteins from adipocyte, heart, intestine, and liver, Kd values are in the nM range and affinities decrease with increasing aqueous solubility of the FA. Binding affinities for heart and liver are generally greater than those for adipocyte and intestine. Moreover, measurements of the rate constants indicate that binding equilibrium at 37øC is achieved within seconds for all FA and FABPs. These results, together with the level of serum (unbound) FFA, suggests a buffering action of FABPs that helps to maintain the intracellular concentration of FFA so that the flux of FFA between serum and cells occurs down a concentration gradient. Measurements of the temperature dependence of binding reveal that the free energy is predominately enthalpic and that the enthalpy of the reaction results from FA-FABP interactions within the binding cavity. The nature of these interactions were investigated by determining the thermodynamics of binding to engineered point mutants of the intestinal FABP. These measurements showed that binding affinities did not report accurately the changes in protein-FA interactions because changes in the binding entropy and enthalpy tend to compensate. For example, an alanine substitution for arginine 106 yields a 30 fold increase in binding affinity, because the loss in enthalpy due to the elimination of the favorable interaction between the FA carboxylate and Arg106, is more than compensated for by an increase in entropy. Thus understanding the effects of amino acid replacements on FA-FABP interactions requires measurements of enthalpy and entropy, in addition to affinity.  相似文献   

13.
The pregnane X receptor (PXR), a member of the nuclear receptor superfamily, regulates the expression of drug-metabolizing enzymes in a ligand-dependent manner. The conventional view of nuclear receptor action is that ligand binding enhances the receptor's affinity for coactivator proteins, while decreasing its affinity for corepressors. To date, however, no known rigorous biophysical studies have been conducted to investigate the interaction among PXR, its coregulators, and ligands. In this work, steady-state total internal reflection fluorescence microscopy (TIRFM) and total internal reflection with fluorescence recovery after photobleaching were used to measure the thermodynamics and kinetics of the interaction between the PXR ligand binding domain and a peptide fragment of the steroid receptor coactivator-1 (SRC-1) in the presence and absence of the established PXR agonist, rifampicin. Equilibrium dissociation and dissociation rate constants of ~5 μM and ~2 s(-1), respectively, were obtained in the presence and absence of rifampicin, indicating that the ligand does not enhance the affinity of the PXR and SRC-1 fragments. Additionally, TIRFM was used to examine the interaction between PXR and a peptide fragment of the corepressor protein, the silencing mediator for retinoid and thyroid receptors (SMRT). An equilibrium dissociation constant of ~70 μM was obtained for SMRT in the presence and absence of rifampicin. These results strongly suggest that the mechanism of ligand-dependent activation in PXR differs significantly from that seen in many other nuclear receptors.  相似文献   

14.
15.
16.
In this work, the behavior of the neurohypophyseal hormones and their selected analogs was studied in the presence of membrane models in an attempt to correlate their activities with a distinct behavior at a level of peptide-lipid interactions. The influence of the peptides studied on the lipid acyl chain order was determined using FTIR spectroscopy. Conformational changes in the peptides upon binding to liposomes were examined using CD spectra. Attempts were also made to determine the binding parameters of the peptides to lipids using isothermal titration calorimetry (ITC). The results show unambiguously that the neurohyphophyseal hormone-like peptides interact with lipids, being a model of a eukaryotic cell membrane. Moreover, hydrophobic interactions between the peptides and liposomes are likely to determine the overall conformation of the peptide, especially below the temperature of the main phase transition (T(m)). Thus, the bulky and hydrophobic nature of the residues incorporated into the N-terminal part of neurohyphophyseal hormones is an important factor for both restriction of peptide mobility and the interaction of the analog with biomembrane. In turn, above T(m), the electrostatic interactions become also relevant for the conformation of the acyclic tail of the AVP-like peptides.  相似文献   

17.
Although interactions of proteins with glycosaminoglycans (GAGs), such as heparin and heparan sulphate, are of great biological importance, structural requirements for protein-GAG binding have not been well-characterised. Ionic interactions are important in promoting protein-GAG binding. Polyelectrolyte theory suggests that much of the free energy of binding comes from entropically favourable release of cations from GAG chains. Despite their identical charges, arginine residues bind more tightly to GAGs than lysine residues. The spacing of these residues may determine protein-GAG affinity and specificity. Consensus sequences such as XBBBXXBX, XBBXBX and a critical 20 Å spacing of basic residues are found in some protein sites that bind GAG. A new consensus sequence TXXBXXTBXXXTBB is described, where turns bring basic interacting amino acid residues into proximity. Clearly, protein-GAG interactions play a prominent role in cell-cell interaction and cell growth. Pathogens including virus particles might target GAG-binding sites in envelope proteins leading to infection. BioEssays 20:156–167, 1998. © 1998 John Wiley & Sons, Inc.  相似文献   

18.
Several binding scaffolds that are not based on immunoglobulins have been designed as alternatives to traditional monoclonal antibodies. Many of them have been developed to bind to folded proteins, yet cellular networks for signaling and protein trafficking often depend on binding to unfolded regions of proteins. This type of binding can thus be well described as a peptide–protein interaction. In this review, we compare different peptide-binding scaffolds, highlighting that armadillo repeat proteins (ArmRP) offer an attractive modular system, as they bind a stretch of extended peptide in a repeat-wise manner. Instead of generating each new binding molecule by an independent selection, preselected repeats – each complementary to a piece of the target peptide – could be designed and assembled on demand into a new protein, which then binds the prescribed complete peptide. Stacked armadillo repeats (ArmR), each typically consisting of 42 amino acids arranged in three α-helices, build an elongated superhelical structure which enables binding of peptides in extended conformation. A consensus-based design approach, complemented with molecular dynamics simulations and rational engineering, resulted in well-expressed monomeric proteins with high stability. Peptide binders were selected and several structures were determined, forming the basis for the future development of modular peptide-binding scaffolds.  相似文献   

19.
Thyrotropin-releasing hormone, TRH (< Glu-His-Proamide), and [N tau-Me-His]TRH (MeTRH) are present as neutral and positively charged forms at physiologic pH, and it was possible that they bind to the TRH receptor (TRH-R) as charged (protonated) species. Binding affinities of TRH and MeTRH to endogenous rat TRH-Rs and to transfected wild type mouse TRH-Rs decreased below pH 7.1. Half-maximal decreases in binding occurred at the approximate pK alpha values of these ligands. Asp to Ala mutations in extracellular loop 1, TM-4, and TM-5 did not decrease binding affinity, but an Asp to Ala mutation in TM-2 caused the affinity to decrease 8-fold. The pH dependences of binding of MeTRH, however, were similar in wild type and all mutant receptors and were consistent with the protonated form of MeTRH binding less well. Thus, the binding of TRH to its receptor does not involve ionic interactions and may be a prototype for binding of neutral peptide ligands to G protein-coupled receptors.  相似文献   

20.
Protein‐protein interactions control a large range of biological processes and their identification is essential to understand the underlying biological mechanisms. To complement experimental approaches, in silico methods are available to investigate protein‐protein interactions. Cross‐docking methods, in particular, can be used to predict protein binding sites. However, proteins can interact with numerous partners and can present multiple binding sites on their surface, which may alter the binding site prediction quality. We evaluate the binding site predictions obtained using complete cross‐docking simulations of 358 proteins with 2 different scoring schemes accounting for multiple binding sites. Despite overall good binding site prediction performances, 68 cases were still associated with very low prediction quality, presenting individual area under the specificity‐sensitivity ROC curve (AUC) values below the random AUC threshold of 0.5, since cross‐docking calculations can lead to the identification of alternate protein binding sites (that are different from the reference experimental sites). For the large majority of these proteins, we show that the predicted alternate binding sites correspond to interaction sites with hidden partners, that is, partners not included in the original cross‐docking dataset. Among those new partners, we find proteins, but also nucleic acid molecules. Finally, for proteins with multiple binding sites on their surface, we investigated the structural determinants associated with the binding sites the most targeted by the docking partners.  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