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.
NMR and density study of Co site binding by polyelectrolytes   总被引:1,自引:0,他引:1  
The changes of density and chemical shifts of the water proton upon addition of CoCl2 to aqueous solutions of tetramethylammonium salts of seven polyelectrolytes (polyphosphate, maleic acid-methylvinylether alternated copolymer, polyacrylic acid, carboxymethylcellulose of substitution degree 0.98, 1.3, 2.1, 2.65) have been measured. Assuming a negligible contribution of pseudo-contact interaction to the water proton chemical shift and a constant hyperfine constant upon displacement of water molecules by other ligands, has permitted the calculation of (i) the number of water molecules released by Co2+ ions upon binding and the approximate fraction of Co2+ ions bound with loss of water, and (ii) the total volume change upon binding and the individual contributions of counterions and polyelectrolyte charged sites to this volume change. Our results are generally in agreement with those obtained using other methods.  相似文献   

2.
3.
BackgroundFragment-based ligand design is used for the development of novel ligands that target macromolecules, most notably proteins. Central to its success is the identification of fragment binding sites that are spatially adjacent such that fragments occupying those sites may be linked to create drug-like ligands. Current experimental and computational approaches that address this problem typically identify only a limited number of sites as well as use a limited number of fragment types.MethodsThe site-identification by ligand competitive saturation (SILCS) approach is extended to the identification of fragment bindings sites, with the method termed SILCS-Hotspots. The approach involves precomputation of the SILCS FragMaps following which the identification of Hotspots, performed by identifying of all possible fragment binding sites on the full 3D structure of the protein followed by spatial clustering.ResultsThe SILCS-Hotspots approach identifies a large number of sites on the target protein, including many sites not accessible in experimental structures due to low binding affinities and binding sites on the protein interior. The identified sites are shown to recapitulate the location of known drug-like molecules in both allosteric and orthosteric binding sites on seven proteins including the androgen receptor, the CDK2 and Erk5 kinases, PTP1B phosphatase and three GPCRs; the β2-adrenergic, GPR40 fatty-acid binding and M2-muscarinic receptors. Analysis indicates the importance of considering all possible fragment binding sites, and not just those accessible to experimental methods, when identifying novel binding sites and performing ligand design versus just considering the most favorable sites. The approach is shown to identify a larger number of known binding sites of drug-like molecules versus the commonly used FTMap and Fpocket methods.General significanceThe present results indicate the potential utility of the SILCS-Hotspots approach for fragment-based rational design of ligands, including allosteric modulators.  相似文献   

4.
The interaction between the dimer structure of ibuprofen drug (D-IB) and calf thymus DNA under simulative physiological conditions was investigated with the use of Hoechst 33258 and methylene blue dye as spectral probes by the methods of UV-visible absorption, fluorescence spectroscopy, circular dichroism spectroscopy and molecular modeling study.Using the Job's plot, a single class of binding sites for theD-IB on DNA was put in evidence. The Stern–Volmer analysis of fluorescence quenching data shows the presence of both the static and dynamic quenching mechanisms. The binding constants, Kb were calculated at different temperatures, and the thermodynamic parameters ?G°, ?H° and ?S° were given. The experimental results showed that D-IB molecules could bind with DNA via groove binding mode as evidenced by: I. DNA binding constant from spectrophotometric studies of the interaction of D-IB with DNA is comparable to groove binding drugs. II. Competitive fluorimetric studies with Hoechst 33258 have shown that D-IB exhibits the ability of this complex to displace with DNA-bounded Hoechst, indicating that it binds to DNA in strong competition with Hoechst for the groove binding. III. There is no significantly change in the absorption of the MB-DNA system upon adding the D-IB, indicates that MB molecules are not released from the DNA helix after addition of the D-IB and are indicative of a non-intercalative mode of binding. IV. Small changes in DNA viscosity in the presence of D-IB, indicating weak link to DNA, which is consistent with DNA groove binding. As well as, induced CD spectral changes, and the docking results revealed that groove mechanism is followed by D-IB to bind with DNA.  相似文献   

5.
β-lactoglobulin (β-LG) is a member of lipocalin superfamily of transporters for small hydrophobic molecules such as retinoids. We located the binding sites of retinol and retinoic acid on β-LG in aqueous solution at physiological conditions, using FTIR, CD, fluorescence spectroscopic methods, and molecular modeling. The retinoid-binding sites and the binding constants as well as the effect of retinol and retinoic acid complexation on protein stability and secondary structure were determined. Structural analysis showed that retinoids bind strongly to β-LG via both hydrophilic and hydrophobic contacts with overall binding constants of K retinol- β -LG?=?6.4 (±?.6)?×?106?M?1 and K retinoic acid- β -LG?=?3.3 (±?.5)?×?106?M?1. The number of retinoid molecules bound per protein (n) is 1.1 (±?.2) for retinol and 1.5 (±?.3) for retinoic acid. Molecular modeling showed the participation of several amino acids in the retinoid–protein complexes with the free binding energy of ?8.11?kcal/mol for retinol and ?7.62?kcal/mol for retinoic acid. Protein conformation was altered with reduction of β-sheet from 59 (free protein) to 52–51% and a major increase in turn structure from 13 (free protein) to 24–22%, in the retinoid–β-LG complexes, indicating a partial protein destabilization.  相似文献   

6.
Methods were developed for obtaining highly viable mouse hepatocytes in single cell suspension and for maintaining the hepatocytes in adherent static culture. The characteristics of transferrin binding and iron uptake into these hepatocytes was investigated. (1) After attachment to culture dishes for 18–24 h hepatocytes displayed an accelerating rate of iron uptake with time. Immediately after isolation mouse hepatocytes in suspension exhibited a linear iron uptake rate of 1.14·105molecules/cell per min in 5 μM transferrin. Iron uptake also increased with increasing transferrin concentration both in suspension and adherent culture. Pinocytosis measured in isolated hepatocytes could account only for 10–20% of the total iron uptake. Iron uptake was completely inhibited at 4°C. (2) A transferrin binding component which saturated at 0.5 μM diferric transferrin was detected. The number of specific, saturable diferric transferrin binding sites on mouse hepatocytes was 4.4·104±1.9·104 for cells in suspension and 6.6·104±2.3·104 for adherent cultured cells. The apparent association constants were 1.23·107 1·mol?1 and 3.4·106 1·mol?1 for suspension and cultured cells respectively. (3) Mouse hepatocytes also displayed a large component of non-saturable transferrin binding sites. This binding increased linearly with transferrin concentration and appeared to contribute to iron uptake in mouse hepatocytes. Assuming that only saturable transferrin binding sites donate iron, the rate of iron uptake is about 2.5 molecules iron/receptor per min at 5 μM transferrin in both suspension and adherent cells and increases to 4 molecules iron/receptor per min at 10 μM transferrin in adherent cultured cells. These rates are considerably greater than the 0.5 molcules/receptor per min observed at 0.5 μM transferrin, the concentration at which the specific transferrin binding sites are fully occupied. The data suggest that either the non-saturable binding component donates some iron or that this component stimulates the saturable component to increase the rate of iron uptake. (4) During incubations at 4°C the majority of the transferrin bound to both saturable and nonsaturable binding sites lost one or more iron atoms. Incubations including 2 mM α,α′-dipyridyl (an Fe11 chelator) decreased the cell associated 59Fe at both 4 and 37°C while completely inhibiting iron uptake within 2–3 min of exposure at 37°C. These observations suggest that most if not all iron is loosened from transferrin upon interaction of transferrin with the hepatocyte membrane. There is also greater sensitivity of 59Fe uptake compared to transferrin binding to pronase digestion, suggesting that an iron acceptor moiety on the cell surface is available to proteolysis.  相似文献   

7.
Cultured Ehrlich ascites tumor cells equilibrate d-glucose via a carrier mechanism with a Km and V of 14 mM and 3 μmol/s per ml cells, respectively. Cytochalasin B competitively inhibits this carrier-mediated glycose transport with an inhibition constant (Ki) of approx. 5·10?7 M. Cytochalasin E does not inhibit this carrier function. With cytochalasin B concentrations up to 1·10?5 M, the range where the inhibition develops to practical completion, three discrete cytochalasin B binding sites, namely L, M and H, are distinguished. The cytochalasin B binding at L site shows a dissociation constant (Kd) of approx. 1·10-6 M, represents about 30% of the total cytochalasin B binding of the cell (8·106 molecules/cell), is sensitively displaced by cytochalasin E but not by d-glucose, and is located in cytosol. The cytochalasin B binding to M site shows a Kd of 4–6·10?7 M, represents approx. 60% of the total saturable binding (14·106 molecules/cell), is specifically displaced by d-glucose with a displacement constant of 15 mM, but not by l-glucose, and is insensitive to cytochalasin E. The sites are membrane-bound and extractable with Triton X-100 but not by EDTA in alkaline pH. The cytochalasin B binding at H site shows a Kd of 2–6 · 10?8 M, represents less than 10% of the total sites (2 · 106 molecules/cell), is not affected by either glucose or cytochalasin E and is of non-cytosol origin. It is concluded that the cytochalasin B binding at M site is responsible for the glucose carrier inhibition by cytochalasin B and the Ehrlich ascites cell is unique among other animal cells in its high content of this site. Approx. 16-fold purification of this site has been achieved.  相似文献   

8.
The influence of water-soluble cationic meso-tetra-(4?N-oxyethylpyridyl)porphyrin (H2TOEPyP4) and it’s metallocomplexes with Ni, Cu, Co, and Zn on hydrodynamic and spectral behavior of DNA solutions has been studied by UV/Vis absorption and viscosity measurement. It was shown that the presence of planar porphyrins such as H2TOEPyP4, NiTOEPyP4, and СuTOEPyP4 leads to an increase in viscosity at relatively small concentrations, and then decrease to stable values. Such behavior is explained by intercalation of these porphyrins in DNA structure because the intercalation mode involves the insertion of a planar molecule between DNA base pairs which results in a decrease in the DNA helical twist and lengthening of the DNA. Further decrease of viscosity is explained by the saturation intercalation sites and occurs outside the binding mode. But, in the case of porphyrins with axial ligands such as CoTOEPyP4 and ZnTOEPyP4, the hydrodynamic parameters decrease, which is explained by self-stacking of these porphyrins in DNA surface. This data are proved by spectral measurements. The results obtained from titration experiments were used for calculation of binding parameters: the binding constant K b and the number of binding sites per base pair n. Obtained data reveal that K b varies between 3.4 and 5.4?×?106?M?1 for a planar porphyrins, a range typical for intercalation mode interactions, and 5.6?×?105?M?1 and 1.8?×?106?M?1 for axial porphyrins. In addition, the exclusion parameter n also testifies that at intercalation, (n~2) the adjacent base pairs are removed to place the planar molecules, and for outside binders to pack on the surface needs too few places (n~0.5–1). It is apparent that the binding is somewhat stronger at intercalation. The viscometric and spectrophotometric measurements are in good agreement.  相似文献   

9.
Interactions between the polyamidoamine (PAMAM) dendrimer and drug molecules are of interest for their potential biomedical applications. The goal of this work is to examine the interaction of PAMAM‐C12 25% dendrimer with morin. The ultraviolet–visible, fluorescence spectroscopic methods as well as molecular modeling were used to analyze drug‐binding mode, binding constants and binding sites, etc. The experimental data showed that the binding constant of morin‐PAMAM‐C12 25% is about 105 L/mol. The interaction of morin with PAMAM‐C12 25% is mainly driven by the hydrophobic, electrostatic, hydrogen bonds and van der Waals forces. There are mainly three classes of binding site of morin at the interface of PAMAM‐C12 25%. These results provided some useful information for self‐assembling and disassembling the PAMAM dendrimer as well as efficient drug delivery and therapeutic applications. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

10.
Hydration of protein cavities influences protein stability, dynamics, and function. Protein active sites usually contain water molecules that, upon ligand binding, are either displaced into bulk solvent or retained to mediate protein–ligand interactions. The contribution of water molecules to ligand binding must be accounted for to compute accurate values of binding affinities. This requires estimation of the extent of hydration of the binding site. However, it is often difficult to identify the water molecules involved in the binding process when ligands bind on the surface of a protein. Cytochrome P450cam is, therefore, an ideal model system because its substrate binds in a buried active site, displacing partially disordered solvent, and the protein is well characterized experimentally. We calculated the free energy differences for having five to eight water molecules in the active site cavity of the unliganded enzyme from molecular dynamics simulations by thermodynamic integration employing a three-stage perturbation scheme. The computed free energy differences between the hydration states are small (within 12 kJ mol−1) but distinct. Consistent with the crystallographic determination and studies employing hydrostatic pressure, we calculated that, although ten water molecules could in principle occupy the volume of the active site, occupation by five to six water molecules is thermodynamically most favorable. Proteins 32:381–396, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

11.
The binding of concanavalin A (ConA) to zona-free unfertilized and fertilized mouse eggs has been investigated using tritiated ConA. At low lectin concentrations (1–5 μg ml?1) the fertilized egg shows a higher affinity for [3H]ConA than does the unfertilized egg. In saturation conditions, however, unfertilized and fertilized eggs show the same binding capacity (1.55 × 108 ConA molecules/egg). The results indicate that ConA-binding sites change qualitatively following fertilization; possible connections between this change and other fertilization-induced changes in the egg surface are discussed.  相似文献   

12.
Abstract

[3H]Phenytoin binding to rat cortical membrane was significantly enhanced in the presence of diazepam. This binding is saturable, reversible and displacable by unlabelled phenytoin. Analyses of the binding data either by the Scatchard plot or by the displacement curve revealed a high and a low affinity sites with Kd values of 32 ± 5 nM and 8.5 ± 1.1 μM, respectively. Similar enhancement of [3H]phenytoin binding was observed when diazepam was replaced by Ro 5–4864 (4″-chlorodiazepam) which is selective for the ‘peripheral’ type benzodiazepine binding sites. In contrast, neither the ‘central’ type receptor selective agonist clonazepam nor the antagonist Ro 15–1788 enhanced [3H]phenytoin binding. Therefore, it seems that these phenytoin binding sites in rat cerebral cortex are associated with a benzodiazepine site similar to the ‘peripheral’ type binding site for its selective affinity for Ro 5–4864. However, judging from the micromolar concentrations required for the enhancement of [3H]phenytoin binding, they appear unlikely to be the same ‘peripheral’ type binding sites as measured by [3H]Ro 5–4864 binding (Kd approx. 1 nM). The micromolar affinity benzodiazepine recognition sites are a possibility, if they indeed exist.  相似文献   

13.
This work deals with the synthesis of 6-methyl-3-[(4′-methylphenyl)imino]methyl-4H-chromen-4-one (MMPIMC), its binding to β-cyclodextrin, and the influence of the cyclodextrin complexation on the compound’s binding to bovine serum albumin (BSA). The 1:2 stoichiometry for the complexation of MMPIMC with β-cyclodextrin is determined with the binding constant of 1.90 × 104 M?2. The structure of host–guest complex plays a role in protein binding of MMPIMC. One- and two-dimensional NMR spectra are used to determine the mode of binding of the guest to β-cyclodextrin cavity and the structure of the inclusion complex is proposed. The binding of MMPIMC with BSA in the absence and the presence of β-cyclodextrin is studied. The binding strengths of MMPIMC–BSA (1.73 × 105 M?1) and β-cyclodextrin-complexed MMPIMC–BSA (9.0 × 104 M?1) show difference in magnitude. The Förster Resonance Energy Transfer efficiency and the proximity of the donor and acceptor molecules, are modulated by β-cyclodextrin. Molecular modeling is used to optimize the sites and mode of binding of MMPIMC with bovine serum albumin.  相似文献   

14.
The molecular forces involved in protein-nucleic acid interaction are electrostatic, stacking and hydrogen-bonding. These interactions have a certain amount of specificity due to the directional nature of such interactions and the spatial contributions of the steric effects of different substituent groups. Quantum chemical calculations on these interactions have been reported which clearly bring out such features. While the binding energies for electrostatic interactions are an order of magnitude higher, the differences in interaction energies for structures stabilised by hydrogen-bonding and stacking are relatively small. Thus, the molecular interactions alone cannot explain the highly specific nature of binding observed in certain segments of proteins and nucleic acids. It is therefore logical to assume that the sequence dependent three dimensional structures of these molecules help to place the functional groups in the correct geometry for a favourable interaction between the two molecules. We have carried out 2D-FT nuclear magnetic resonance studies on the oligonucleotide d-GGATCCGGATCC. This oligonucleotide sequence has two binding sites for the restriction enzyme Bam H1. Our studies indicate that the conformation of this DNA fragment is predominantly B-type except near the binding sites where the ribose ring prefers a3E conformation. This interesting finding raises the general question about the presence of specificity in the inherent backbone structures of proteins and nucleic acids as opposed to specific intermolecular interactions which may induce conformational changes to facilitate such binding.  相似文献   

15.
Due to Ca2+‐dependent binding and the sequence diversity of Calmodulin (CaM) binding proteins, identifying CaM interactions and binding sites in the wet‐lab is tedious and costly. Therefore, computational methods for this purpose are crucial to the design of such wet‐lab experiments. We present an algorithm suite called CaMELS (CalModulin intEraction Learning System) for predicting proteins that interact with CaM as well as their binding sites using sequence information alone. CaMELS offers state of the art accuracy for both CaM interaction and binding site prediction and can aid biologists in studying CaM binding proteins. For CaM interaction prediction, CaMELS uses protein sequence features coupled with a large‐margin classifier. CaMELS models the binding site prediction problem using multiple instance machine learning with a custom optimization algorithm which allows more effective learning over imprecisely annotated CaM‐binding sites during training. CaMELS has been extensively benchmarked using a variety of data sets, mutagenic studies, proteome‐wide Gene Ontology enrichment analyses and protein structures. Our experiments indicate that CaMELS outperforms simple motif‐based search and other existing methods for interaction and binding site prediction. We have also found that the whole sequence of a protein, rather than just its binding site, is important for predicting its interaction with CaM. Using the machine learning model in CaMELS, we have identified important features of protein sequences for CaM interaction prediction as well as characteristic amino acid sub‐sequences and their relative position for identifying CaM binding sites. Python code for training and evaluating CaMELS together with a webserver implementation is available at the URL: http://faculty.pieas.edu.pk/fayyaz/software.html#camels .  相似文献   

16.
An increasing number of medically important proteins are challenging drug targets because their binding sites are too shallow or too polar, are cryptic and thus not detectable without a bound ligand or located in a protein–protein interface. While such proteins may not bind druglike small molecules with sufficiently high affinity, they are frequently druggable using novel therapeutic modalities. The need for such modalities can be determined by experimental or computational fragment based methods. Computational mapping by mixed solvent molecular dynamics simulations or the FTMap server can be used to determine binding hot spots. The strength and location of the hot spots provide very useful information for selecting potentially successful approaches to drug discovery.  相似文献   

17.
One approach to study the structure of promoter regions1–3 is to isolate RNA polymerase binding sites. Several attempts have been made to isolate such sites as protected DNA fragments4–10, but so far the DNA isolated has not definitely shown to be only one or a few specific sites. We report here the isolation of a single RNA polymerase binding site from the replicative form (RF) DNA of bacteriophage fd. The site as isolated is a short double-stranded DNA with a unique nucleotide sequence.  相似文献   

18.
Lactogenic hormones from the placenta and pituitary are primarily responsible for the growth and function of the mammary gland during pregnancy and lactation. In the present study we describe the optimal conditions for the measurement of 125I-labeled ovine prolactin binding to mammary gland slices of pregnant and lactating rats. Prolactin binding is saturable (Kd approx. 2.36 · 10?9 M), hormone specific and destroyed by proteases. The hormonal environments of pregnancy and lacation dramatically influence the availability and measurement of prolactin binding sites. Whereas binding consistently appears to be low in mammary glands removed from rats during pregnancy, binding levels rise 7–8-fold shortly after birth and remain high during the 22 days of lactation. However, the removal of the ovaries and gravid uteri at specific times during pregnancy results in prompt 3–6-fold increase in prolactin binding. Elevated levels in potential prolactin binding capacity appear in mammary tissue coincident with the reported rise in serum rat placental lactogen between the eight and eleventh days. We suggest that high levels of this lactogenic hormone promote the appearance of prolactin binding sites during pregnancy and mask the sites such that they are not available for measurement in vitro.  相似文献   

19.
Protein and RNA molecules interact and form complexes in many biological processes. However, it is still unclear how they can find the correct docking direction before forming complex. In this paper, we study preorientation of RNA and protein separated at a distance of 5–7?Å just before they form contacts and interact with each other only through pure electrostatic interaction when neglecting the influence of other molecules and complicated environment. Since geometric complementary has no meaning at such a distance, this is not a docking problem and so the conventional docking methods, like FTDock, are inapplicable. However, like the usual docking problem, we need to sample all the positions and orientations of RNA surrounding the protein to find the lowest energy orientations between RNA and protein. Therefore, we propose a long-range electrostatic docking-like method using Fast Fourier Transform-based sampling, LEDock, to study this problem. Our results show that the electrostatically induced orientations between RNA and protein at a distance of 5–7?Å are very different from the random ones and are much closer to those in their native complexes. Meanwhile, electrostatic funnels are found around the RNA-binding sites of the proteins in 62 out of 78 bound protein–RNA complexes. We also tried to use LEDock to find RNA-binding residues and it seems to perform slightly better than BindN Server for 23 unbound protein–RNA complexes.  相似文献   

20.

Sulfur-containing sites in proteins are of great importance for both protein structure and function, including enzymatic catalysis, signaling pathways, and recognition of ligands and protein partners. Selenium-77 is an NMR active spin-1/2 nucleus that shares many physiochemical properties with sulfur and can be readily introduced into proteins at sulfur sites without significant perturbations to the protein structure. The sulfur-containing amino acid methionine is commonly found at protein–protein or protein–ligand binding sites. Its selenium-containing counterpart, selenomethionine, has a broad chemical shift dispersion useful for NMR-based studies of complex systems. Methods such as (1H)-77Se-13C double cross polarization or {77Se}-13C REDOR could be valuable to map the local environment around selenium sites in proteins but have not been demonstrated to date. In this work, we explore these dipolar transfer mechanisms for structural characterization of the GB1 V39SeM variant of the model protein GB1 and demonstrate that 77Se-13C based correlations can be used to map the local environment around selenium sites in proteins. We have found that the general detection limit is?~?5 Å, but longer range distances up to?~?7 Å can be observed as well. This study establishes a framework for the future characterization of selenium sites at protein–protein or protein–ligand binding interfaces.

  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