首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Simultaneous formulation binding and structure modification of the binding site leads to binding process that can be analyzed within the framework of the non-linear theory of dynamic systems. Such an approach allows us to obtain several properties of the binding center: plurality of stationary (stable and unstable) states at binding, recognition of bistable and hysteretic binding modes. It is also shown that adsorption centre deformation leads to a S-shaped adsorption curve.  相似文献   

2.
The adhesive properties of the mouse P388D1 macrophage-like line were explored. Cells were deposited in glass capillary tubes, and the kinetics of adhesion and spreading were studied. Binding involved the cell metabolism since it was decreased by cold, azide, or a divalent cation chelator. Glass-adherent cells were subjected to calibrated laminar shear flows with a highly viscous dextran solution. A tangential force of about 5×10−3 dyn/cell was required to achieve substantial detachment. The duration of application of the shearing force strongly influenced cell-substrate separation when this was varied from 1–10 s. Further, this treatment resulted in marked cell deformation, with the appearance of an elongated shape. Hence, cell-substrate separation is a progressive process, and binding strength is expected to be influenced by cell deformability. The minimum time required for adhesion was also investigated by making cells adhere under flow conditions. The maximum flow rate compatible with adhesion was about 1000-fold lower than that required to detach glass-bound cells. A simple model was devised to provide a quantitative interpretation for the experimental results of kinetic studies. It is concluded that cell-to-glass adhesion required a cell-substrate contact longer than a few seconds. This first step of adhesion was rapidly followed by a large (about 1000-fold) increase of adhesion strength. It is therefore emphasized that adhesion is heavily dependent on the duration of cell-to-cell encounter, as well as the force used to remove so-called unbound cells.  相似文献   

3.
4.
T F Wei  W Bujalowski  T M Lohman 《Biochemistry》1992,31(26):6166-6174
The Escherichia coli single-strand binding (SSB) protein is an essential protein involved in DNA replication, recombination, and repair processes. The tetrameric protein binds to ss nucleic acids in a number of different binding modes in vitro. These modes differ in the number of nucleotides occluded per SSB tetramer and in the type and degree of cooperative complexes that are formed with ss DNA. Although it is not yet known whether only one or all of these modes function in vivo, based on the dramatically different properties of the SSB tetramer in these different ss DNA binding modes, it has been suggested that the different modes may function selectively in replication, recombination, and/or repair. The transitions between these different modes are very sensitive to solution conditions, including salt (concentration, as well as cation and anion type), pH, and temperature. We have examined the effects of multivalent cations, principally the polyamine spermine, on the SSB-ss poly(dT) binding mode transitions and find that the transition from the (SSB)35 to the (SSB)56 binding mode can be induced by micromolar concentrations of polyamines as well as the inorganic cation Co(NH3)6(3+). Furthermore, these multivalent cations, as well as Mg2+, induce the binding mode transition by binding cooperatively to the SSB-poly(dT) complexes. These observations are interesting in light of the fact that polyamines, such as spermidine, are part of the ionic environment in E. coli and hence these cations are likely to affect the distribution of SSB-ss DNA binding modes in vivo.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Binding capacity is the homotropic second derivative of the binding potential with respect to the chemical potential of the ligand. It provides a measure of steepness of the binding isotherm and represents the extent of cooperativity. In the present study, the shape of the binding capacity curve for various systems was investigated and the relation between binding capacity and the extent of cooperativity examined. In this regard, a novel linear graphical method was introduced for binding data analysis. The stoichiometry of binding and the extent of cooperativity can be determined by this method. This method has been successfully applied to various systems such as binding of oxygen to hemoglobin, warfarin to human serum albumin and dodecyltrimethylammonium bromide to alpha-amylase.  相似文献   

6.
Surface pressure measurements, external reflection-Fourier transform infrared spectroscopy, and neutron reflectivity have been used to investigate the lipid-binding behavior of three antimicrobial peptides: melittin, magainin II, and cecropin P1. As expected, all three cationic peptides were shown to interact more strongly with the anionic lipid, 1,2 dihexadecanoyl-sn-glycerol-3-(phosphor-rac-(1-glycerol)) (DPPG), compared to the zwitterionic lipid, 1,2 dihexadecanoyl-sn-glycerol-3-phosphocholine (DPPC). All three peptides have been shown to penetrate DPPC lipid layers by surface pressure, and this was confirmed for the melittin-DPPC interaction by neutron reflectivity measurements. Adsorption of peptide was, however, minimal, with a maximum of 0.4 mg m(-2) seen for melittin adsorption compared to 2.1 mg m(-2) for adsorption to DPPG (from 0.7 microM solution). The mode of binding to DPPG was shown to depend on the distribution of basic residues within the peptide alpha-helix, although in all cases adsorption below the lipid layer was shown to dominate over insertion within the layer. Melittin adsorption to DPPG altered the lipid layer structure observed through changes in the external reflection-Fourier transform infrared lipid spectra and neutron reflectivity. This lipid disruption was not observed for magainin or cecropin. In addition, melittin binding to both lipids was shown to be 50% greater than for either magainin or cecropin. Adsorption to the bare air-water interface was also investigated and surface activity followed the trend melittin>magainin>cecropin. External reflection-Fourier transform infrared amide spectra revealed that melittin adopted a helical structure only in the presence of lipid, whereas magainin and cecropin adopted helical structure also at an air-water interface. This behavior has been related to the different charge distributions on the peptide amino acid sequences.  相似文献   

7.
Prediction of RNA-RNA interaction is a key to elucidating possible functions of small non-coding RNAs, and a number of computational methods have been proposed to analyze interacting RNA secondary structures. In this article, we focus on predicting binding sites of target RNAs that are expected to interact with regulatory antisense RNAs in a general form of interaction. For this purpose, we propose bistaRNA, a novel method for predicting multiple binding sites of target RNAs. bistaRNA employs binding profiles that represent scores for hybridized structures, leading to reducing the computational cost for interaction prediction. bistaRNA considers an ensemble of equilibrium interacting structures and seeks to maximize expected accuracy using dynamic programming. Experimental results on real interaction data validate good accuracy and fast computation time of bistaRNA as compared with several competitive methods. Moreover, we aim to find new targets given specific antisense RNAs, which provides interesting insights into antisense RNA regulation. bistaRNA is implemented in C++. The program and Supplementary Material are available at http://rna.naist.jp/program/bistarna/.  相似文献   

8.
Nearest-neighbor cooperative binding of a ligand covering n sites and binding with equilibrium constant K and cooperativity factor omega to a large molecule with m binding sites (m much greater than n omega, n/omega) can be approximately described by a Gaussian distribution P(q-qmax), where q is the number of ligands bound and qmax the most probable value of q. The variance of the Gaussian is equal to the derivative dqmax/d ln(L), where L is the free ligand concentration. This variance, sigma 2, is a complicated function of qmax. However, in the limits of very large cooperativity, omega much greater than 1, very large anticooperativity, omega much less than 1, or noncooperativity, omega = 1, simpler expressions for sigma 2 can be given. For qmax = m/(n + 1), where the most probable number of bound ligands equals the number of free binding sites, sigma 2 has a particularly simple form: sigma 2 = 2m omega 1/2/(n + 1)3. The Gaussian and the infinite lattice approximations for the average number of ligands bound are good approximations only if sigma is much smaller than the number of binding sites. The variance may therefore provide an easy check on the validity of the infinite lattice approximation, which is commonly used to analyze experimental binding data.  相似文献   

9.
10.
Calcium binding to complexes of calmodulin and calmodulin binding proteins   总被引:12,自引:0,他引:12  
B B Olwin  D R Storm 《Biochemistry》1985,24(27):8081-8086
The free energy of coupling for binding of Ca2+ and the calmodulin-sensitive phosphodiesterase to calmodulin was determined and compared to coupling energies for two other calmodulin binding proteins, troponin I and myosin light chain kinase. Free energies of coupling were determined by quantitating binding of Ca2+ to calmodulin complexed to calmodulin binding proteins with Quin 2 to monitor free Ca2+ concentrations. The geometric means of the dissociation constants (-Kd) for Ca2+ binding to calmodulin in the presence of equimolar rabbit skeletal muscle troponin I, rabbit skeletal muscle myosin light chain kinase, and bovine heart calmodulin sensitive phosphodiesterase were 2.1, 1.1, and 0.55 microM. The free-energy couplings for the binding of four Ca2+ and these proteins to calmodulin were -4.48, -6.00, and -7.64 kcal, respectively. The Ca2+-independent Kd for binding of the phosphodiesterase to calmodulin was estimated at 80 mM, indicating that complexes between calmodulin and this enzyme would not exist within the cell under low Ca2+ conditions. The large free-energy coupling values reflect the increase in Ca2+ affinity of calmodulin when it is complexed to calmodulin binding proteins and define the apparent positive cooperativity for Ca2+ binding expected for each system. These data suggest that in vitro differences in free-energy coupling for various calmodulin-regulated enzymes may lead to differing Ca2+ sensitivities of the enzymes.  相似文献   

11.
Aptamer-based drugs represent an attractive approach in pharmacological therapy. The most studied aptamer, thrombin binding aptamer (TBA), folds into a well-defined quadruplex structure and binds to its target with good specificity and affinity. Modified aptamers with improved biophysical properties could constitute a new class of therapeutic aptamers. In this study we show that the modified thrombin binding aptamer (mTBA), 3′GGT5′-5′TGGTGTGGTTGG3′, which also folds into a quadruplex structure, is more stable than its unmodified counterpart and shows a higher thrombin affinity. The stability of the modified aptamer was investigated using differential scanning calorimetry, and the energetics of mTBA and TBA binding to thrombin was characterized by means of isothermal titration calorimetry (ITC). ITC data revealed that TBA/thrombin and mTBA/thrombin binding stoichiometry is 1:2 for both interactions. Structural models of the two complexes of thrombin with TBA and with mTBA were also obtained and subjected to molecular dynamics simulations in explicit water. Analysis of the models led to an improvement of the understanding of the aptamer-thrombin recognition at a molecular level.  相似文献   

12.
Selective binding of cholesterol by recombinant fatty acid binding proteins   总被引:3,自引:0,他引:3  
The sterol binding specificity of rat recombinant liver fatty acid binding protein (L-FABP) and intestinal fatty acid binding protein (I-FABP) was characterized with [3H]cholesterol and a fluorescent sterol analog dehydroergosterol. Ligand binding analysis, fluorescence spectroscopy, and activation of microsomal acyl-CoA:cholesterol acyltransferase activity showed that L-FABP-bound sterols. 1) Lipidex-1000 assay showed a dissociation constant Kd = 0.78 +/- 0.18 microM and stoichiometry of 0.47 +/- 0.16 mol/mol for [3H]cholesterol binding to L-PABP. 2) With [3H]cholesterol/phosphatidylcholine liposomes, the cholesterol binding parameters for L-FABP were Kd = 1.53 +/- 0.28 microM and stoichiometry 0.83 +/- 0.07 mol/mol. 3) L-FABP interaction with dehydroergosterol altered the fluorescence intensity and polarization of dehydroergosterol. Dehydroergosterol bound to L-FABP with Kd = 0.37 microM and a stoichiometry of 0.83 mol/mol. 4) Cholesterol and dehydroergosterol decreased L-FABP tyrosine lifetime. Dehydroergosterol binding produced sensitized emission of bound dehydroergosterol with longer lifetime.5) L-FABP bound two cis-parinaric acid molecules/molecule of protein. Cholesterol displaced one of these bound cis-parinaric acids. 6) L-FABP enhanced acyl-CoA:cholesterol acyltransferase in a concentration-dependent manner. In contrast, these assays indicated that I-FABP did not bind sterols. Thus, L-FABP appears able to bind 1 mol of cholesterol/mol of L-FABP, the L-FABP sterol binding site is equivalent to one of the two fatty acid binding sites, and L-FABP stimulates acyl-CoA:cholesterol acyltransferase by transfer of cholesterol.  相似文献   

13.
Detecting cis-regulatory binding sites for cooperatively binding proteins   总被引:1,自引:0,他引:1  
Several methods are available to predict cis-regulatory modules in DNA based on position weight matrices. However, the performance of these methods generally depends on a number of additional parameters that cannot be derived from sequences and are difficult to estimate because they have no physical meaning. As the best way to detect cis-regulatory modules is the way in which the proteins recognize them, we developed a new scoring method that utilizes the underlying physical binding model. This method requires no additional parameter to account for multiple binding sites; and the only necessary parameters to model homotypic cooperative interactions are the distances between adjacent protein binding sites in basepairs, and the corresponding cooperative binding constants. The heterotypic cooperative binding model requires one more parameter per cooperatively binding protein, which is the concentration multiplied by the partition function of this protein. In a case study on the bacterial ferric uptake regulator, we show that our scoring method for homotypic cooperatively binding proteins significantly outperforms other PWM-based methods where biophysical cooperativity is not taken into account.  相似文献   

14.
G Walker  R Yabkowitz  D R Burgess 《Biochemistry》1991,30(42):10206-10210
The way in which actin and myosin II become localized to the contractile ring of dividing cells resulting in cleavage furrow formation and cytokinesis is unknown. While much is known about actin binding proteins and actin localization, little is known about myosin localization. A 53-kDa (53K) polypeptide present in the sea urchin egg binds to myosin II in a nucleotide-dependent manner and mediates its solubility in vitro [Yabkowitz, R., & Burgess, D.R. (1987) J. Cell Biol. 105, 927-936]. The binding site of 53K on the myosin molecule was examined in an effort to understand the mechanism of 53K-induced myosin solubility and its potential function in myosin regulation. Blot overlay and chemical cross-linking techniques utilizing myosin proteolytic fragments indicate that 53K binds to fragments proximal to the head-rod junction of myosin. Fragments distal to the head-rod junction do not bind 53K. In addition, the binding of 53K to myosin largely inhibits protease digestion that produces the head and rod fragments. The binding of 53K to the head-rod domain of myosin may be critical in regulation of myosin conformation, localization, assembly, and ATPase activity.  相似文献   

15.
S-shaped binding curves often characterize interactions of ligands with nucleic acid molecules as analyzed by different physico-chemical and biophysical techniques. S-shaped experimental binding curves are usually interpreted as indicative of the positive cooperative interactions between the bound ligand molecules. This paper demonstrates that S-shaped binding curves may occur as a result of the "mixed mode" of DNA binding by the same ligand molecule. Mixed mode of the ligand-DNA binding can occur, for example, due to 1) isomerization or dimerization of the ligands in solution or on the DNA lattice, 2) their ability to intercalate the DNA and to bind it within the minor groove in different orientations. DNA-ligand complexes are characterized by the length of the ligand binding site on the DNA lattice (so-called "multiple-contact" model). We show here that if two or more complexes with different lengths of the ligand binding sites could be produced by the same ligand, the dependence of the concentration of the complex with the shorter length of binding site on the total concentration of ligand should be S-shaped. Our theoretical model is confirmed by comparison of the calculated and experimental CD binding curves for bis-netropsin binding to poly(dA-dT) poly(dA-dT). Bis-netropsin forms two types of DNA complexes due to its ability to interact with the DNA as monomers and trimers. Experimental S-shaped bis-netropsin-DNA binding curve is shown to be in good correlation with those calculated on the basis of our theoretical model. The present work provides new insight into the analysis of ligand-DNA binding curves.  相似文献   

16.
The binding of serotonin to a soluble, high affinity binding protein, present in synaptosomes and associated with serotonergic tracts, has now been studied for the effects of metallic ions and various drugs. At optimal concentration (10-4 M) of Fe2+ the enhancement of binding was close to 20-fold. A much smaller effect was noted with Cu2+. With other ions (Fe3+, Mn2+, Co2+, Ni2+, Cr3+, Mg2+, Ca2+) little or no effect was seen. For the effect with Fe2+. preincubation was required (10 min, 25°C) and concentrations higher than 10-4M were inhibitory. Studies based on equilibrium dialysis show that the effect of Fe2+ was on the affinity of the binding of serotonin to the protein, rather than on the binding capacity. In polydcrylamide gels at pH 8.6 the migratory properties of thc serotonin-protein complex formed in the presence of Fe2+ differ from those of the complex formed without Fe2+. Nucleotides (ATP, GTP, ADP, AMP) inhibited thc binding. The effects of several classes of drugs (inhibitors of biogenic amine storage and uptake, psychotomimetics, MAO) inhibitors and drugs binding to contractile proteins) were also studied. The only effective inhibitors of serotonin binding were reserpine, vinblastine and CZ-74, which caused 50% inhibition at 2 × 10-6 M, 7.5 × 10-6 M and 0.2 × 10-6M respectively.  相似文献   

17.
Directed binding     
We propose a novel physical mechanism to describe the mode of processive propagation of twoheaded kinesin motor proteins along microtubule (MT) filaments. Binding and unbinding of the kinesin heads to and from the MT filament play a crucial role in producing movement. The chemical energy of adenosine triphosphate hydrolysis is used in large part for the unbinding process of kinesin from the MT filament. Importantly, in our model, the binding of each head is to be directionally oriented to the MT filament. Therefore, we treat the two motor domains (heads) as extended objects that are connected with each other by a neck region that contains the kinesin dimerization domain. The head domains recognize tubulin binding sites by feeling the two-dimensional periodic potential from the MT surface and are also subjected to thermal noise. Using experimentally determined results regarding physical parameters of the walk, we develop a simple mathematical and mechanical model in which directed binding of the heads to tubulin results in a directed twist of the molecule, probably in the neck linker region, away from its relaxed state. Unbinding of the head from the filament relaxes the twist and defines the propagation direction. We showed that there must be at least two torsional springs (one for every head) involved that can store elastic energy. Consequently, in our model, it is the internal structure both of the relaxed and tensed-up state and the transition mode between them that define the walking direction of kinesin. We present calculations based on the model that are in good quantitative agreement with experimental observations for kinesin.  相似文献   

18.
Formulas for the free energy of binding to a macromolecule are obtained by thermodynamic and statistical mechanical methods, and it is shown that the free energy of binding is intimately related with the binding polynomial and Wyman's binding potential. The expression for the free energy of binding is applied to a number of cases of ligand-induced conformational changes, cooperativity, association reactions, etc.  相似文献   

19.
Apparently homogeneous odorant binding protein purified from pig nasal mucosa (pOBP) exhibited subunit molecular masses of 17 223, 17 447, and 17 689 (major component) Da as estimated by ESI/MS. According to gel filtration, this protein, its truncated forms, and/or its variants are homodimeric under physiologic conditions (pH 6-7, 0.1 M NaCl). The dimer if monomer equilibrium shifts toward a prevalent monomeric form at pH <4.5. Velocity sedimentation reveals a monomeric state of OBP at both pH 7.2 and 3.5, indicating a pressure-induced dissociation of the homodimer. High-sensitivity differential scanning calorimetry (HS-DSC) shows that the unfolding transition of pOBP is reversible at neutral pH. It is characterized by the transition temperature of 69.23 degrees C and an enthalpy of 391.1 kJ/mol per monomer. The transition heat capacity curve of pOBP is well-approximated by the two-state model on the level of subunit, indicating that the two monomers behave independently. Isothermal titration calorimetry (ITC) shows that at physiological pH pOBP binds 2-isobutyl-3-methoxypyrazine (IBMP) and 3,7-dimethyloctan-1-ol (DMO) with association constants of 3.19 x 10(6) and 4.94 x 10(6) M(-)(1) and enthalpies of -97.2 and -87.8 kJ/mol, respectively. The binding stoichiometry of both ligands is nearly one molecule of ligand per homodimer of pOBP. The interaction of pOBP with both ligands is enthalpically driven with an unfavorable change of entropy. The binding affinity of pOBP with IBMP does not change significantly at acidic pH, while the binding stoichiometry is nearly halved. According to HS-DSC data, the interaction with IBMP and DMO leads to a substantial stabilization of the pOBP folded structure, which is manifested by the increase in the unfolding temperature and enthalpy. The calorimetric data allow us to conclude that the mechanism of binding of the studied odorants to pOBP is not dominated by a hydrophobic effect related to any change in the hydration state of protein and ligand groups but, most likely, is driven by polar and van der Waals interactions.  相似文献   

20.
The binding of retinoic acid to mutants of Cellular Retinoic Acid Binding Protein II (CRABPII) was evaluated to better understand the importance of the direct protein/ligand interactions. The important role of Arg111 for the correct structure and function of the protein was verified and other residues that directly affect retinoic acid binding have been identified. Furthermore, retinoic acid binding to CRABPII mutants that lack all previously identified interacting amino acids was rescued by providing a carboxylic acid dimer partner in the form of a Glu residue. Proteins 2009. © 2008 Wiley‐Liss, Inc.  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