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.
The ligand binding curve for a macromolecular system presents the average number or ligand molecules bound per macromolecule as a function of the chemical potential or the logarithm of the ligand concentration. We show that various observable properties of this curve, for example its asymptotes and derivatives, are expressible in terms of linear combinations of the mole fractions αi of macromolecules binding i molecules of ligand. Whenever enough such properties of the binding curve are known, the linear equations in αi can be solved to give the mole fractions of each of the various macromolecular species. An application of these results is that a Hill plot for hemoglobin-ligand equilibrium where the asymptotes approach unit slope can be made to yield the four Adair constants by a simple algebraic method. A second use is that a knowledge of the first and second derivatives of the binding curve at points along the curve can yield the species fractions as functions of the degree of saturation without direct knowledge of the ligand binding constants. These methods are illustrated by some numerical examples.  相似文献   

2.
Kinetic, thermodynamic, and structural properties of the aminoglycoside N3‐acetyltransferase‐VIa (AAC‐VIa) are determined. Among the aminoglycoside N3‐acetyltransferases, AAC‐VIa has one of the most limited substrate profiles. Kinetic studies showed that only five aminoglycosides are substrates for this enzyme with a range of fourfold difference in kcat values. Larger differences in KM (~40‐fold) resulted in ~30‐fold variation in kcat/KM. Binding of aminoglycosides to AAC‐VIa was enthalpically favored and entropically disfavored with a net result of favorable Gibbs energy (ΔG < 0). A net deprotonation of the enzyme, ligand, or both accompanied the formation of binary and ternary complexes. This is opposite of what was observed with several other aminoglycoside N3‐acetyltransferases, where ligand binding causes more protonation. The change in heat capacity (ΔCp) was different in H2O and D2O for the binary enzyme–sisomicin complex but remained the same in both solvents for the ternary enzyme–CoASH–sisomicin complex. Unlike, most other aminoglycoside‐modifying enzymes, the values of ΔCp were within the expected range of protein‐carbohydrate interactions. Solution behavior of AAC‐VIa was also different from the more promiscuous aminoglycoside N3‐acetyltransferases and showed a monomer‐dimer equilibrium as detected by analytical ultracentrifugation (AUC). Binding of ligands shifted the enzyme to monomeric state. Data also showed that polar interactions were the most dominant factor in dimer formation. Overall, thermodynamics of ligand‐protein interactions and differences in protein behavior in solution provide few clues on the limited substrate profile of this enzyme despite its >55% sequence similarity to the highly promiscuous aminoglycoside N3‐acetyltransferase. Proteins 2017; 85:1258–1265. © 2017 Wiley Periodicals, Inc.  相似文献   

3.
A simple method for the detection of sequence- and structural-selective ligand binding to nucleic acids is described. The method is based on the commonly used thermal denaturation method in which ligand binding is registered as an elevation in the nucleic acid melting temperature (Tm). The method can be extended to yield a new, higher -throughput, assay by the simple expediency of melting designed mixtures of polynucleotides (or oligonucleotides) with different sequences or structures of interest. Upon addition of ligand to such mixtures at low molar ratios, the Tm is shifted only for the nucleic acid containing the preferred sequence or structure. Proof of principle of the assay is provided using first a mixture of polynucleotides with different sequences and, second, with a mixture containing DNA, RNA and two types of DNA:RNA hybrid structures. Netropsin, ethidium, daunorubicin and actinomycin, ligands with known sequence preferences, were used to illustrate the method. The applicability of the approach to oligonucleotide systems is illustrated by the use of simple ternary and binary mixtures of defined sequence deoxyoligonucleotides challenged by the bisanthracycline WP631. The simple mixtures described here provide proof of principle of the assay and pave the way for the development of more sophisticated mixtures for rapidly screening the selectivity of new nucleic acid binding compounds.  相似文献   

4.
Caffeine (CAF) and other methylxanthines (MTX) may interact directly with several aromatic, intercalating ligands through mixed stacking aggregation. Formation of such stacking hetero-complexes may decrease their free form concentration and, in consequence, diminish their biological activity, which is often related to their direct interaction with DNA. In this paper interactions of acridine mutagen (ICR191) with DNA in the presence of three MTX: caffeine (CAF), pentoxifylline (PTX) and theophylline (TH) are investigated. Several mathematical models are used to calculate all association constant values and every component concentration in each analyzed mixture. Model McGhee–von Hippel is used to analyze ligand–DNA interaction, and model Zdunek et al.—to analyze ligand–MTX interactions. Finally, two distinct mathematical models are employed to analyze three-component mixture containing ligand, MTX and DNA molecules. The first model describes possible interactions of ligand with DNA and MTX, and rejects direct MTX interactions with DNA. The second model describes all interactions mentioned above and, additionally, allows MTX to interact directly with DNA. Results obtained using these models are similar. However, correspondence of theoretical results to experimental data is better for the first model than the second one. In this paper possible interactions of ICR191 with eukaryotic cell chromatin are also analyzed, showing that CAF reduces acridine mutagen potential to interact directly with cell chromatin. Additionally, it is demonstrated that MTX inhibit mutagenic activity of ICR191 in a dose-dependent manner. Furthermore, biological activity of ICR191–MTX mixtures corresponds with concentration of free mutagen form calculated using appropriate mathematical model.  相似文献   

5.
Abstract

The association (k+1) and dissociation (k-1) rate constants, and the equilibrium thermodynamic binding parameters (ΔG°, ΔH° and ΔS°) of the β-adrenergic ligand [125Iodo]cyanopindolol (ICYP) were studied in a crude lung membrane preparation of rats of different ages. There was no difference in k+1-values for the different age groups, while the k-1-values were in all cases difficult to measure: almost no dissociation of ICYP from its binding site occurs. The thermodynamic properties were not affected by age. It is concluded that, in these experimental conditions, age has no effect on the kinetic parameters of the binding of ICYP to the β-adrenoceptors in rat lung.  相似文献   

6.
The dissociation of insulin from human insulin antibodies has been investigated using a technique that is rapid and does not require addition of excess unlabelled insulin. A slow (k1 = 2·1?3 min?1 and a fast (k2 = 4·10?2 min?1) dissociating antibody component were identified in all studies. These have been shown to correspond, respectively, to the high and low affinity antibody components of equilibrium binding studies. The range of k1 and k2 values and their response to temperature change is small. Insulin resistance and stability of diabetes are not related to properties of antibody dissociation. Dissociation is faster in the presence of high (6–850 nM) insulin concentration due to increased binding to the fast dissociating component without change in the dissociation rate constants. When incubation time is increased beyond achivement of maximal binding there is a time-dependent rise in binding to the slow dissociating component, with a concomitant fall in k1. The traditional concept that equilibrium is established at maximum binding requires further examination.  相似文献   

7.
Herein, we report the metal complexation properties of a macrocyclic ligand (L) that contains three pyridines as well as three urea groups. Linear and strand like ligands are typically used to afford helical coordination polymer. Our reported macrocyclic ligand (L) has remarkable flexibility and can twist upon dative bond formation. Two macrocyclic ligands complex with three cadmium atoms to form a helicate monomeric structure [Cd3L2(H2O)6(CH3CN)2]6+, which extends to a 1D polymeric structure via hydrogen-bonding. We also investigated the binding property of this new ligand in solution by NMR and UV-Vis spectroscopy. These results together with diffusion NMR studies suggest that in solution this ligand also forms an oligomeric complex with cadmium.  相似文献   

8.

Background

The combinatorial library strategy of using multiple candidate ligands in mixtures as library members is ideal in terms of cost and efficiency, but needs special screening methods to estimate the affinities of candidate ligands in such mixtures. Herein, a new method to screen candidate ligands present in unknown molar quantities in mixtures was investigated.

Results

The proposed method involves preparing a processed-mixture-for-screening (PMFS) with each mixture sample and an exogenous reference ligand, initiating competitive binding among ligands from the PMFS to a target immobilized on magnetic particles, recovering target-ligand complexes in equilibrium by magnetic force, extracting and concentrating bound ligands, and analyzing ligands in the PMFS and the concentrated extract by chromatography. The relative affinity of each candidate ligand to its reference ligand is estimated via an approximation equation assuming (a) the candidate ligand and its reference ligand bind to the same site(s) on the target, (b) their chromatographic peak areas are over five times their intercepts of linear response but within their linear ranges, (c) their binding ratios are below 10%. These prerequisites are met by optimizing primarily the quantity of the target used and the PMFS composition ratio. The new method was tested using the competitive binding of biotin derivatives from mixtures to streptavidin immobilized on magnetic particles as a model. Each mixture sample containing a limited number of candidate biotin derivatives with moderate differences in their molar quantities were prepared via parallel-combinatorial-synthesis (PCS) without purification, or via the pooling of individual compounds. Some purified biotin derivatives were used as reference ligands. This method showed resistance to variations in chromatographic quantification sensitivity and concentration ratios; optimized conditions to validate the approximation equation could be applied to different mixture samples. Relative affinities of candidate biotin derivatives with unknown molar quantities in each mixture sample were consistent with those estimated by a homogenous method using their purified counterparts as samples.

Conclusions

This new method is robust and effective for each mixture possessing a limited number of candidate ligands whose molar quantities have moderate differences, and its integration with PCS has promise to routinely practice the mixture-based library strategy.  相似文献   

9.
Fluorescence anisotropy assay was implemented for characterization of ligand binding dynamics to melanocortin 4 (MC4) receptors. This approach enables on-line monitoring of reactions that is essential for estimation of more correct binding parameters, understanding of ligand binding and its regulation mechanisms, and design of new drugs with desirable properties. Two different red-shifted fluorophore-labeled peptide ligands, Cy3B-NDP-α-MSH and TAMRA-NDP-α-MSH, were used and compared in assays that monitored their binding to MC4 receptors in membranes of Sf9 insect cells. The Cy3B dye-labeled ligand exhibited improved performance in assays when compared with the TAMRA-labeled ligand, having higher photostability, insensitivity to buffer properties, and better signal/noise ratio. The binding of both ligands to membranes of Sf9 cells expressing MC4 receptors was saturable and with high affinity. All studied MC4 receptor-specific nonlabeled ligands displaced fluoroligands’ binding in a concentration-dependent manner with potencies in agreement with their pharmacological activities. On-line monitoring of the reactions revealed that equilibrium of peptide binding was not reached even after 3 h. Real-time monitoring of ligand binding dynamics enabled us to find optimal experimental conditions for each particular ligand and an improved estimate of their binding parameters.  相似文献   

10.
Computer simulations of equilibrium binding studies of a mixture of two labeled ligands binding competitively to a single class of identical and independent sites (receptors) were performed to investigate how ligand heterogeneity affects the observed data in such studies. The simulated data are presented in Scatchard plots. Ligand heterogeneity was generally found to be indistinguishable from the case of a homogeneous ligand when usual experimental conditions applied (that is, Scatchard plots of the data were straight lines). Some factors that increased the probability of recognizing heterogeneity in the system were identified, however. These are 1) a large difference between the dissociation constants of the two ligands, 2) a high concentration of receptors relative to the dissociation constant of the higher-affinity ligand, 3) a high concentration of the lower-affinity ligand relative to that of the higher-affinity ligand, 4) a high specific activity of the lower-affinity ligand relative to that of the higher-affinity ligand, and 5) lack of experimental error. When ligand heterogeneity (under certain conditions) did cause curvilinearity in the Scatchard plot, the curve formed was always concave-downwards. Thus, ligand heterogeneity may occasionally mimic positive cooperativity, but never mimics negative cooperativity or multiple classes of binding sites. Implications of these findings for equilibrium binding studies involving lipoproteins (which are generally isolated as heterogeneous mixtures of particles) are discussed in detail. These findings are also relevant to equilibrium binding studies using ligands which are mixtures of stereoisomers or which contain chemical or radiochemical impurities.  相似文献   

11.
In this report, we applied site-specifically deuterated N-stearoylsphingomyelins (SSMs) to raft-exhibiting ternary mixtures containing SSM, 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC), and cholesterol (Chol) and successfully acquired deuterium quadrupole coupling profiles of SSM from liquid-ordered (Lo) and liquid-disordered (Ld) domains. To our knowledge, this is the first report that shows detailed lipid chain dynamics separately and simultaneously obtained from coexisting Lo and Ld domains. We also found that the quadrupole profile of the Lo phase in the ternary system was almost identical to that in the SSM-Chol binary mixture, suggesting that the order profile of the binary system is essentially applicable to more complicated membrane systems in terms of the acyl chain order. We also demonstrated that 2H NMR spectroscopy, in combination with organic synthesis of deuterated components, could be used to reveal the accurate mole fractions of each component distributed in the Lo and Ld domains. As compared with the reported tie-line analysis of phase diagrams, the merit of our 2H NMR analysis is that the domain-specific compositional fractions are directly attainable without experimental complexity and ambiguity. The accurate compositional distributions as well as lipid order profiles in ternary mixtures are relevant to understanding the molecular mechanism of lipid raft formation.  相似文献   

12.
The energies of binding of seven ligands by p-hydroxybenzoate hydroxylase (PHBH) were calculated theoretically. Direct enzyme–ligand interaction energies were calculated using the ab initio quantum mechanical model assembly of the active site at the 3-21G level. Solvation energies of the ligands needed in the evaluation of the binding energies were calculated with the semiempirical AM1–SM2 method and the long-range electrostatic interaction energies between the ligands and the protein matrix classically using the static charge distributions of the ligands and the protein. Energies for proton-transfer between the ligands OH or SH substituent at position 4 and the active-site tyrosine within the ab initio model assemblies were calculated and compared to the corresponding pKas in aqueous solution. Excluding 3,4-dihydroxybenzoate, the natural product of PHBH, a linear relationship between the calculated binding energies and the experimental binding free energies was found with a correlation coefficient of 0.90. Contributions of the direct enzyme–ligand interaction energies, solvation energies and the long-range electrostatic interaction energies to the calculated binding energies were analyzed. The proton-transfer energies of the ligands with substituents ortho to the ionized OH were found to be perturbed less in the model calculations than the energies of their meta isomers as deduced from the corresponding pKas. © 1995 Wiley-Liss, Inc.  相似文献   

13.
Molecular recognition force spectroscopy, a biosensing atomic force microscopy technique allows to characterise the dissociation of ligand–receptor complexes at the molecular level. Here, we used molecular recognition force spectroscopy to study the binding capability of recently developed testosterone binders. The two avidin‐based proteins called sbAvd‐1 and sbAvd‐2 are expected to bind both testosterone and biotin but differ in their binding behaviour towards these ligands. To explore the ligand binding and dissociation energy landscape of these proteins, we tethered biotin or testosterone to the atomic force microscopy probe while the testosterone‐binding protein was immobilized on the surface. Repeated formation and rupture of the ligand–receptor complex at different pulling velocities allowed determination of the loading rate dependence of the complex‐rupturing force. In this way, we obtained the molecular dissociation rate (koff) and energy landscape distances (xβ) of the four possible complexes: sbAvd‐1‐biotin, sbAvd‐1‐testosterone, sbAvd‐2‐biotin and sbAvd‐2‐testosterone. It was found that the kinetic off‐rates for both proteins and both ligands are similar. In contrast, the xβ values, as well as the probability of complex formations, varied considerably. In addition, competitive binding experiments with biotin and testosterone in solution differ significantly for the two testosterone‐binding proteins, implying a decreased cross‐reactivity of sbAvd‐2. Unravelling the binding behaviour of the investigated testosterone‐binding proteins is expected to improve their usability for possible sensing applications. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

14.
Abstract

The screening of combinatorial libraries requires a deconvolution procedure to obtain, in fine, the most active compound of the starting library. the standard screening assays used in regular molecular pharmacology, have been poorly assessed when transposed to combinatorial chemistry-related experiments, particularly those involving large numbers of chemicals in a single assay. One key issue is the effect of the inactive analogs on the identification of the active ligand in mixtures. We chose melatonin receptors to measure the apparent affinity of a single ligand when tested alone or in mixtures of non-peptide low molecular weight compounds. Using ligands with IC50 from the micro- to the picomolar range, mixed with increasingly complex mixtures of 5 to 20 or 25 inactive compounds, we analyzed the displacements from the mt1 and MT2 melatonin receptor subtypes of the radioligand 2-iodomelatonin (Kd= 25pmol/l and 200pmol/l, respectively). the behavior of equimolar mixtures in displacement curves led to the conclusion that the observed binding affinity reflects the dilution effect of mixing the active component with inactive compounds but does not reveal noticeable interactions which would interfere with the binding process. From the practical point of view, the concentrations of the active species in the binding assay should be large enough to displace significantly the radioligand, a requirement which may be limited by the solubility of the ligand mixtures. in contrast, previous observations with peptide libraries report that the dilution effect is often compensated by additive or synergic action of structurally related analogs, thus making possible the deconvolution of very large (typically up to 107 compounds) peptide libraries.  相似文献   

15.
16.
This protocol describes a method for identifying unknown target proteins from a mixture of biomolecules for a given drug or a lead compound. This method is based on a combination of chromatography and isothermal titration calorimetry (ITC) where ITC is used as a tracking tool. The first step involves the use of ITC to confirm the binding of ligand to a component in the biomolecular mixture. Subsequently, the biomolecular mixture is fractionated by chromatography, and the binding of the ligand with individual fractions (or subfractions) is verified by ITC. The iteration of chromatographic purification on the fractions combined with ITC results in identifying the target protein. This method is useful when the target protein or ligand is unknown and/or not amenable to labeling, chemical modification or immobilization. This protocol has been successfully used by our team and by others to identify both low-abundance and highly abundant target proteins present in biomolecular mixtures. With this protocol, it takes approximately 3-5 d to identify the target protein from a mixture.  相似文献   

17.
18.
Upon refluxing 2:1 mixtures of guanine (guH) and MnCl2, FeCl2 or NiCl2 in a 7:3 (v/v) mixture of ethanol and triethyl orthoformate for 1–2 weeks, partial substitution of gu? for Cl? groups occurs, and solid complexes of the M(gu)Cl·2ROH (R = C2H5 for M = Mn; R = H for M = Fe, Ni) type are obtained. The new complexes are pentacoordinated and appear to be linear chainlike polymeric species, involving a single-bridged
n backbone. Coordination number five is attained by the presence of one terminal chloro and two terminal ROH ligands per metal ion. Most probable binding sites of bidentate bridging gu? are the N(7) and N(9) imidazole ring nitrogens. IR evidence rules out the possibility of coordination of gu? through any of the exocyclic potential ligand sites (O(6) oxygen or N(2) nitrogen) [1].  相似文献   

19.
The kinetics of antibody–antigen interactions are reviewed in terms of general trends observed in both polyclonal and monoclonal antibody populations. Anti-fluorescein antibodies are featured in the review as model proteins to explore fluorescence-based kinetic measurements. Since the fluorescence of the fluorescein ligand is significantly quenched upon interaction with both polyclonal and monoclonal anti-fluorescein antibodies, the quenching parameter can be advantageously employed in measuring the rates of association (k1) and dissociation (k2). The near diffusion-limited k1 rates and the prolonged k2 rates are discussed in terms of antibody affinity and mechanisms involved in ligand binding. Specific prolongation effects of reagents, such as anti-metatype antibodies, on the dissociation rate are discussed in terms of antibody dynamics and conformational substates.  相似文献   

20.
Serum albumin, a protein naturally abundant in blood plasma, shows remarkable ligand binding properties of numerous endogenous and exogenous compounds. Most of serum albumin binding sites are able to interact with more than one class of ligands. Determining the protein‐ligand interactions among mammalian serum albumins is essential for understanding the complexity of this transporter. We present three crystal structures of serum albumins in complexes with naproxen (NPS): bovine (BSA‐NPS), equine (ESA‐NPS), and leporine (LSA‐NPS) determined to 2.58 Å (C2), 2.42 Å (P61), and 2.73 Å (P212121) resolutions, respectively. A comparison of the structurally investigated complexes with the analogous complex of human serum albumin (HSA‐NPS) revealed surprising differences in the number and distribution of naproxen binding sites. Bovine and leporine serum albumins possess three NPS binding sites, but ESA has only two. All three complexes of albumins studied here have two common naproxen locations, but BSA and LSA differ in the third NPS binding site. None of these binding sites coincides with the naproxen location in the HSA‐NPS complex, which was obtained in the presence of other ligands besides naproxen. Even small differences in sequences of serum albumins from various species, especially in the area of the binding pockets, influence the affinity and the binding mode of naproxen to this transport protein. Proteins 2014; 82:2199–2208. © 2014 Wiley Periodicals, 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