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 0 毫秒
1.
The identification of protein mutations that enhance binding affinity may be achieved by computational or experimental means, or by a combination of the two. Sources of affinity enhancement may include improvements to the net balance of binding interactions of residues forming intermolecular contacts at the binding interface, such as packing and hydrogen-bonding interactions. Here we identify noncontacting residues that make substantial contributions to binding affinity and that also provide opportunities for mutations that increase binding affinity of the TEM1 beta-lactamase (TEM1) to the beta-lactamase inhibitor protein (BLIP). A region of BLIP not on the direct TEM1-binding surface was identified for which changes in net charge result in particularly large increases in computed binding affinity. Some mutations to the region have previously been characterized, and our results are in good correspondence with this results of that study. In addition, we propose novel mutations to BLIP that were computed to improve binding significantly without contacting TEM1 directly. This class of noncontacting electrostatic interactions could have general utility in the design and tuning of binding interactions.  相似文献   

2.
A sensitive ESR method which allows a direct quantitative determination of nucleic acid binding affinities of proteins under physiologically relevant conditions has been applied to the gene 5 protein of bacteriophage fd. This was achieved with two spin-labeled nucleic acids, (ldT, dT)n and (lA,A)n, which served as macro-molecular spin probes in ESR competition experiments. With the two different macromolecular spin probes, it was possible to determine the relative apparent affinity constants, Kapp, over a large affinity domain. In 20 mM Tris X HCl (pH 8.1), 1 mM sodium EDTA, 0.1 mM dithiothreitol, 10% (w/v) glycerol, 0.05% Triton, and 125 mM NaCl, the following affinity relationship was observed: K(dT)napp = 10(3) KfdDNAapp = 2 X 10(4) K(A)napp = 6.6 X 10(4) KrRNAapp = 1.5 X 10(5) KR17RNAapp. Increasing the [NaCl] from 125 to 200 mM caused considerably less tight binding of gene 5 protein to (lA,A)n, and a typical cooperative binding isotherm was observed, whereas at the lower [NaCl] used for the competition experiments, the binding was essentially stoichiometric. A computer fit of the experimental titration data at 200 mM NaCl gave an intrinsic binding constant, Kint, of 1300 M-1 and a cooperativity factor, omega, of 60 (Kint omega = Kapp) for (lA,A)n.  相似文献   

3.
A theoretical relationship has been developed to allow the effect of free ligand concentration on the capacity of an affinity Chromatography matrix to be determined where the protein adsorbed shows co-operative binding. Computer simulations using literature values for association constants show that under optimal conditions resin capacity can be increased significantly in the presence of a small but finite concentration of free ligand. The model also allows prediction of the soluble ligand concentration required for biospecific elution. The results obtained suggest the possibility of a new elution technique, "reverse biospecific elution," that reduces the amount of free ligand required to effect elution.  相似文献   

4.
This study reports on various parameters which affect the binding stoichiometry for complexes of bacteriophage T4 gene 32 protein (P32) and single stranded polynucleotides (determined by UV absorbance and fluorescence quenching) and presents results of a quantitative electron spin resonance assay to determine physiologically effective binding affinity differences of nucleic acid binding proteins. The assay employs macromolecular spin probes (spin-labeled nucleic acids) which are used to determine the fraction of saturation in competition experiments with unlabeled nucleic acids. It was found that the fraction of complexed spin-labeled polynucleotides can be directly monitored by ESR with a two-component analysis approach when ligands such as poly(L-lysine), gene 5 protein (P5) of filamentous bacteriophage fd, and gene 32 protein (P32) of bacteriophage T4 are used. The ESR data unequivocally show that: 1) the binding stoichiometry for poly(L-lysine), P5 and P32 is nucleotide/lysine, 4 nucleotides/P5 monomer, and 10 nucleotides/P32 monomer, respectively; and 2) under physiologically relevant buffer conditions the relative affinity of P32 in the cooperative binding mode for polythymidylic acid is about 4 times greater than for polydeoxyinosinic acid and about 12 times greater than for polyinosinic acid, and the relative affinity of P32 for polydeoxyinosinic acid is about 3 times greater than for polyinosinic acid.  相似文献   

5.
Transferrins are two-domain proteins with a very strong site for iron binding located in each domain. Using ultrasensitive titration calorimetry, the binding of ferric ion (chelated with a 2-fold molar excess of nitrilotriacetate) to the two sites of ovotransferrin was studied in detail as well as the binding to the single site in the N- and C-terminal half-molecules. In the presence of excess bicarbonate ion, the binding occurs in two kinetic steps. The fast process of contact binding is instantaneous with respect to instrument response time, is strongly exothermic for the N site and less so for the C site, and corresponds to binding of the chelated ferric ion. The slower process of bicarbonate insertion with concomitant release of nitrilotriacetate occurs on a time scale of 2-20 min over the temperature range 7-37 degrees C and is endothermic for the N site and exothermic for the C site, with rates being significantly slower for insertion at the C site. The delta H of binding is strongly temperature-dependent for both sites, arising from a large negative delta Cp of binding which probably indicates removal of hydrophobic groups from contact with water. When bicarbonate ion is absent, only the fast process of contact binding is seen. Each site within a half-molecule is qualitatively similar to the same site in intact ovotransferrin, although quantitative differences were detected. It was shown that contact binding to ovotransferrin occurs reversibly with free exchange of Fe+3 between N and C sites, while the attachment to either site becomes essentially irreversible after bicarbonate insertion. The strong preference for the first ferric ion to bind to the N site is shown to be due to its larger contact binding constant and the faster rate of bicarbonate insertion, relative to the C site, and is not due to stronger thermodynamic binding after bicarbonate insertion. True equilibrium is achieved only over much longer periods of time. In another series of experiments, direct binding studies were carried out between the two half-molecules under different states of ligation with Fe+3 in the presence of bicarbonate. The results indicate that the two binding sites in ovotransferrin, separated by ca. 40 A, are not independent of one another but communicate as a result of ligand-dependent changes in the heats and free energies of domain-domain interactions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
Ma Y  Lieber MR 《Biochemistry》2001,40(32):9638-9646
Despite its central role in the nonhomologous DNA end joining process, we still have an incomplete picture of the interaction between Ku and DNA. Here we describe both kinetic (surface plasmon resonance or SPR) and equilibrium (electrophoretic mobility shift assay or EMSA) studies of Ku binding to linear double-stranded DNA. Ku interaction with 1-site DNA is noncooperative, as expected. Electrophoretic mobility shift assays indicate cooperativity in the binding of Ku molecules to DNA long enough for two Ku molecules to bind (2-site DNA). For the kinetic studies, we use surface plasmon resonance in which one end of the DNA molecules is linked to a surface while the other end is free to interact with Ku. We find that one Ku molecule dissociates from 1-site DNA with simple Langmuir (i.e., independent) kinetics. However, two Ku molecules associate and dissociate from 2-site DNA with a time course that cannot be described as a simple Langmuir interaction. On 3- and 4-site DNA, EMSA and SPR studies do not reveal any cooperativity, suggesting that the middle Ku does not exhibit cooperative interaction with the two Ku molecules bound at the DNA ends. These results indicate that Ku molecules can demonstrate cooperative interaction, and this is influenced by their positions along the DNA.  相似文献   

7.
The molecular interactions of laminin with several tumor cell lines and skin fibroblasts were investigated by radioligand binding studies and cell attachment assays using laminin, the laminin-nidogen complex, and laminin fragments as substrates and also domain-specific antibodies as inhibitors of cell attachment. The majority of cells showed a dual binding pattern for fragments 1 and 8 which originate from short-arm or long-arm structures of laminin, respectively. Both of these fragments in solution bind to suspended cells with high affinity (KD = 1-10 nM), with the receptor numbers for each fragment depending on the cell type. Competition studies and independent variation of receptor numbers demonstrated that the cell-binding structures on each fragment are different, implicating the existence of two distinct cellular receptors for laminin. The ability of these fragments to act as substrates for cell adhesion correlated with the presence of high affinity binding sites on the cells. However, only antibodies to fragment 8 were able to block cell adhesion to laminin, despite the presence of binding sites for fragment 1. A few cells had very low numbers of high affinity receptors for either fragment 1 or 8. The latter cell type was used to demonstrate that complex formation between laminin and nidogen, which binds to fragment 1 structures, reduces the potential of laminin for cell binding.  相似文献   

8.
Escherichia coli apomanganese superoxide dismutase, prepared by removing the native metal ion under denaturing conditions, exhibits thermally triggered metal uptake behavior previously observed for thermophilic and hyperthermophilic superoxide dismutases but over a lower temperature range. Differential scanning calorimetry of aposuperoxide dismutase and metalated superoxide dismutase unfolding transitions has provided quantitative estimates of the metal binding affinities for manganese superoxide dismutase. The binding constant for Mn(II) (K(Mn(II)) = 3.2 x 10(8) m(-1)) is surprisingly low in light of the essentially irreversible metal binding characteristic of this family of proteins and indicates that metal binding and release processes are dominated by kinetic, rather than thermodynamic, constraints. The kinetic stability of the metalloprotein complex can be traced to stabilization by elements of the protein that are independent of the presence or absence of the metal ion reflected in the thermally triggered metalation characteristic of these proteins. Binding constants for Mn(III), Fe(II), and Fe(III) complexes were estimated using quasireversible values for the unfolding enthalpy and DeltaC(p) for apo-Mn superoxide dismutase and the observed T(m) values for unfolding the metalated species in the absence of denaturants. For manganese and iron complexes, an oxidation state-dependent binding affinity reflects the protein perturbation of the metal redox potential.  相似文献   

9.
This short review reports on progresses in the study of drug-melanin interactions using the technique of affinity chromatography. Melanins are natural or synthetic pigments derived from the oxidation and polymerization of various precursors including L-dopa, tyrosine and cystein. Accumulation of toxic compounds, drugs, and metal ions in pigmented tissues through reversible binding to melanin has been linked to chronic toxicity. Affinity chromatography using chromatographic stationary phases based on physically adsorbed or chemically bonded melanin provides a useful tool for studying the interactions of small molecules and metal ions with melanin  相似文献   

10.
11.
12.
Calorimetric studies of cytochrome oxidase-phospholipid interactions   总被引:1,自引:0,他引:1  
Thermotropic phase transitions in phospholipid vesicles reconstituted with mitochondrial cytochrome oxidase (EC 1.9.3.1) were studied using differential scanning calorimetry. Both dimyristoylphosphatidylcholine (DMPC) and mixtures of DMPC and cardiolipin were used at different lipid-to-protein ratios. The incorporated protein reduces the energy absorbed during phase transitions of DMPC vesicles, and causes a small decrease in the transition temperature (tm). delta H depends on the amount of protein in the vesicles. This dependence indicates that about 72 DMPC molecules are influenced per cytochrome alpha alpha 3 monomer. The transition parameters remain unaffected by changes in ionic strength or by reduction of the enzyme. Incorporation of cytochrome oxidase depleted of subunit III into DMPC liposomes resulted in a larger decrease of tm, but the amount of perturbed phospholipids remains similar to that in the case of the intact enzyme. Incorporation of cytochrome oxidase into DMPC/cardiolipin vesicles counteracts the effect of cardiolipin in decreasing the enthalpy of the DMPC transition. Thus cytochrome oxidase segregates the phospholipids by attracting cardiolipin from the bulk lipid. Cytochrome c does not significantly affect this apparent cardiolipin 'shell' around membranous cytochrome oxidase.  相似文献   

13.
Circular dichroism (CD) is a useful technique for an assessment of DNA-binding mode, being a more accessible, low-resolution complement to NMR and X-ray diffraction methods. Ligand-DNA interactions can be studied by virtue of the interpretation of induced ligand CD signals resulting from the coupling of electric transition moments of the ligand and DNA bases within the asymmetric DNA environment. This protocol outlines methods to determine the binding mode and affinity of ligand-DNA interactions and takes approximately 7.5 h.  相似文献   

14.
When two proteins associate they form a molecular interface that is a structural and energetic mosaic. Within such interfaces, individual amino acid residues contribute distinct binding energies to the complex. In combination, these energies are not necessarily additive, and significant positive or negative cooperative effects often exist. The basis of reliable algorithms to predict the specificities and energies of protein-protein interactions depends critically on a quantitative understanding of this cooperativity. We have used a model protein-protein system defined by an affinity maturation pathway, comprising variants of a T cell receptor Vbeta domain that exhibit an overall affinity range of approximately 1500-fold for binding to the superantigen staphylococcal enterotoxin C3, in order to dissect the cooperative and additive energetic contributions of residues within an interface. This molecular interaction has been well characterized previously both structurally, by x-ray crystallographic analysis, and energetically, by scanning alanine mutagenesis. Through analysis of group and individual maturation and reversion mutations using surface plasmon resonance spectroscopy, we have identified energetically important interfacial residues, determined their cooperative and additive energetic properties, and elucidated the kinetic and thermodynamic bases for molecular evolution in this system. The summation of the binding free energy changes associated with the individual mutations that define this affinity maturation pathway is greater than that of the fully matured variant, even though the affinity gap between the end point variants is relatively large. Two mutations in particular, both located in the complementarity determining region 2 loop of the Vbeta domain, exhibit negative cooperativity.  相似文献   

15.
We demonstrate here that the Saccharomyces cerevisiae Mlh1-Pms1 heterodimer required for DNA mismatch repair and other cellular processes is a DNA binding protein. Binding was evaluated using a variety of single and double-stranded DNA molecules. Mlh1-Pms1 bound short substrates with low affinity and showed a slight preference for single-stranded DNA. In contrast, Mlh1-Pms1 exhibited a much higher affinity for long DNA molecules, suggesting that binding is cooperative. High affinity binding required a duplex DNA length greater than 241 base-pairs. The rate of association with DNA was rapid and dissociation of protein-DNA complexes following extensive dilution was very slow. However, in competition experiments, we observed a rapid active transfer of Mlh1-Pms1 from labeled to unlabeled DNA. Binding was non-sequence specific and highly sensitive to salt type and concentration, suggesting that Mlh1-Pms1 primarily interacts with the DNA backbone via ionic contacts. Cooperative binding was observed visually by atomic force microscopy as long, continuous tracts of Mlh1-Pms1 protein bound to duplex DNA. These images also showed that Mlh1-Pms1 simultaneously interacts with two different regions of duplex DNA. Taken together, the atomic force microscope images and DNA binding assays provide strong evidence that Mlh1-Pms1 binds duplex DNA with positive cooperativity and that there is more than one DNA binding site on the heterodimer. These DNA binding properties of Mlh1-Pms1 may be relevant to its participation in DNA mismatch repair, recombination and cellular responses to DNA damage.  相似文献   

16.
The influence of H1 and H5 histones proteins upon the accessibility of ethidium bromide into chromatin is studied by steady-state fluorescence anisotropy in the range of r-values ([Dye]/[Phosphate]) smaller than 0.01. This corresponds to the very strong binding process. When H1 and H5 are present, the DNA segment which contains the binding sites is 25–30 base pairs long, even if H1 and H5 are digested by trypsin or by natural proteolysis, but presumably still interacting with the DNA chromatin. On the contrary, when H1 or H5 are separated from chromatin by an increase of the ionic strength, ethidium binds to a segment of DNA about 55–60 base pairs long. We may explain the results by assuming that the ethidium sites are located on a continuous segment constituting about one half of the linker, the other half interacting with H1 and H5. When chromatin is depleted from these proteins, the high affinity sites are distributed all along the linker.  相似文献   

17.
To clarify the interplay between the binding affinity and kinetics of protein–protein interactions, and the possible role of intrinsically disordered proteins in such interactions, molecular simulations were carried out on 20 protein complexes. With bias potential and reweighting techniques, the free energy profiles were obtained under physiological affinities, which showed that the bound‐state valley is deep with a barrier height of 12 ? 33 RT. From the dependence of the affinity on interface interactions, the entropic contribution to the binding affinity is approximated to be proportional to the interface area. The extracted dissociation rates based on the Arrhenius law correlate reasonably well with the experimental values (Pearson correlation coefficient R = 0.79). For each protein complex, a linear free energy relationship between binding affinity and the dissociation rate was confirmed, but the distribution of the slopes for intrinsically disordered proteins showed no essential difference with that observed for ordered proteins. A comparison with protein folding was also performed. Proteins 2016; 84:920–933. © 2016 Wiley Periodicals, Inc.  相似文献   

18.
Peptidyl arms extending from one protein domain to another protein domain mediate many important interactions in biology. A well-studied example of this type of protein-protein interaction occurs between the yeast homeodomain proteins, MAT alpha2 and MAT a1, which form a high-affinity heterodimer on DNA. The carboxyl-terminal arm extending from MAT alpha2 to MAT a1 has been proposed to produce an allosteric conformational change in the a1 protein that generates a very large increase in the DNA binding affinity of a1. Although early studies lent some support to this model, a more recent crystal structure determination of the free a1 protein argues against any allosteric change. This note presents a thermodynamic argument that accounts for the proteins' binding behavior, so that allosteric conformational changes are not required to explain the large affinity increase. The analysis presented here should be useful in analyzing binding behavior in other systems involving arm interactions.  相似文献   

19.
The trp RNA-binding Attenuation Protein (TRAP) from Bacillus subtilis is an 11-subunit protein that binds a series of 11 GAG and UAG repeats separated by two to three-spacer nucleosides in trp leader mRNA. The structure of TRAP bound to an RNA containing 11 GAG repeats shows that the RNA wraps around the outside of the protein ring with each GAG interacting with the protein in nearly identical fashion. The only direct hydrogen bond interactions between the protein and the RNA backbone are to the 2'-hydroxyl groups on the third G of each repeat. Replacing all 11 of these guanosines with deoxyriboguanosine eliminates measurable binding to TRAP. In contrast, a single riboguanosine in an otherwise entirely DNA oligonucleotide dramatically stabilizes TRAP binding, and facilitates the interaction of the remaining all-DNA portion with the protein. Studies of TRAP binding to RNAs with between 2 and 11 GAGs, UAGs, AAGs, or CAGs showed that the stability of a TRAP-RNA complex is not directly proportional to the number of repeats in the RNA. These studies also showed that the effect of the identity of the residue in the first position of the triplet, with regard to binding to TRAP, is dependent on the number of repeats in the RNA. Together these data support a model in which TRAP binds to RNA by first forming an initial complex with a small subset of the repeats followed by a cooperative interaction with the remaining triplets.  相似文献   

20.
The first biochemical and spectroscopic characterization of a purified membrane transporter for riboflavin (vitamin B(2)) is presented. The riboflavin transporter RibU from the bacterium Lactococcus lactis was overexpressed, solubilized, and purified. The purified transporter was bright yellow when the cells had been cultured in rich medium. We used a detergent-compatible matrix-assisted laser desorption ionization time-of-flight mass spectrometry method (Cadene, M., and Chait, B. T. (2000) Anal. Chem. 72, 5655-5658) to show that the source of the yellow color was riboflavin that had been co-purified with the transporter. The method appears generally applicable for substrate identification of purified membrane proteins. Substrate-free RibU was produced by expressing the protein in cells cultured in chemically defined medium. Riboflavin, FMN, and roseoflavin bound to RibU with high affinity and 1:1 stoichiometry (K(d) for riboflavin is 0.6 nM), but FAD did not bind to the transporter. The absorption spectrum of riboflavin changed dramatically when the substrate bound to RibU. Well resolved bands appeared at 441, 464, and 486 nm, indicating a hydrophobic binding pocket. The fluorescence of riboflavin was almost completely quenched upon binding to RibU, and also the tryptophan fluorescence of the transporter was quenched when flavins bound. The results indicate that riboflavin is stacked with one or more tryptophan residues in the binding pocket of RibU. Mutagenesis experiments showed that Trp-68 was involved directly in the riboflavin binding. The structural properties of the binding site and mechanistic consequences of the exceptionally high affinity of RibU for its substrate are discussed in relation to soluble riboflavin-binding proteins of known structure.  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