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.
Naphthylphthalamic acid (NPA), an inhibitor of polar auxin transport, binds with high affinity to membrane preparations from callus and cell suspension cultures derived from Nicotiana tabacum (K d approx. 2·10–9 M). The concentration of membrane-bound binding sites is higher in cell suspension than in callus cultures. The binding of NPA to these sites seems to be a simple process, in contrast to the binding of the synthetic auxin naphthylacetic acid (1-NAA) to membrane preparations from callus cultures, which is more complex (A.C. Maan et al., 1983, Planta 158, 10–15). Naphthylacetic acid, a number of structurally related compounds and the auxin-transport inhibitor triiodobenzoic acid were all able to compete with NPA for the same binding site with K d values ranging from 10–6 to 10–4 M. On the other hand, NPA was not able to displace detectable amounts of NAA from the NAA-binding site. A possible explantation is the existence of two different membrane-bound binding sites, one exclusively for auxins and one for NPA as well as auxins, that differ in concentration. The NPA-binding site is probably an auxin carrier.Abbreviations 1-NAA 1-Naphthylacetic acid - 2-NAA 2-Naphthylacetic acid - NPA N-1-Naphthylphthalamic acid  相似文献   

2.
Phosphopeptide pTyr-Glu-Glu-Ile (pYEEI) has been introduced as an optimal Src SH2 domain ligand. Peptides, Ac-K(IDA)pYEEIEK(IDA) (1), Ac-KpYEEIEK (2), Ac-K(IDA)pYEEIEK (3), and Ac-KpYEEIEK(IDA) (4), containing 0–2 iminodiacetate (IDA) groups at the N- and C-terminal lysine residues were synthesized and evaluated as the Src SH2 domain binding ligands. Fluorescence polarization assays showed that peptide 1 had a higher binding affinity (Kd = 0.6 μM) to the Src SH2 domain when compared with Ac-pYEEI (Kd = 1.7 μM), an optimal Src SH2 domain ligand, and peptides 24 (Kd = 2.9–52.7 μM). The binding affinity of peptide 1 to the SH2 domain was reduced by more than 2-fold (Kd = 1.6 μM) upon addition of Ni2+ (300 μM), possibly due to modest structural effect of Ni2+ on the protein as shown by circular dichroism experimental results. The binding affinity of 1 was restored in the presence of EDTA (300 μM) (Kd = 0.79 μM). These studies suggest that peptides containing IDA groups may be used for designing novel SH2 domain binding ligands.  相似文献   

3.
ATP and adenylylimidodiphosphate (AdoPP[NH]P) bind to (Na+ + K+)-ATPase in the absence of Mg2+ (EDTA present) with a homogeneous but 15-fold different affinity, the Kd values being 0.13 μM and 1.9 μM, respectively. The binding capacities of the two nucleotides are nearly equal and amount to 3.9 and 4 nmol/mg protein or 1.7 and 1.8 mol/mol (Na+ + K+)-ATPase, respectively. The Kd value for ATP is equal to the Km for phosphorylation by ATP (0.05–0.25 μM) and the binding capacity is equivalent to the phosphorylation capacity of 1.8 mol/mol (Na+ + K+)-ATPase. Hence, the enzyme contains two high-affinity nucleotide binding and phosphorylating sites per molecule, or one per α-subunit. Additional low-affinity nucleotide binding sites are elicited in the presence of Mg2+, as shown by binding studies with the non-phosphorylating (AdoPP[NH]P). The Kd and binding capacity for AdoPP[NH]P at these sites is dependent on the Mg2+ concentration. The Kd increases from 0.06 mM at 0.5 mM Mg2+ to a maximum of 0.26 mM at 2 mM Mg2+ and the binding capacity from 1.5 nmol/mg protein at 0.5 mM Mg2+ to 3.3 nmol/mg protein at 4 mM Mg2+. Extrapolation of a double reciprocal plot of binding capacity vs. total Mg2+ concentration yields a maximal binding capacity at infinite Mg2+ concentration of 3.8 nmol/mg protein or 1.7 mol/mol (Na+ + K+)-ATPase. The Kd for Mg2+ at the sites, where it exerts this effect, is 0.8 mM. The Kd for the high-affinity sites increases from 1.5–1.9 μM in the absence of Mg2+ to a maximum of 4.2 μM at 2 mM Mg2+ concentration. The binding capacity of these sites (1.8 mol/mol enzyme) is independent of the Mg2+ concentration. Hence, Mg2+ induces two low-affinity non-phosphorylating nucleotide binding sites per molecule (Na+ + K+)-ATPase in addition to the two high-affinity, phosphorylating nucleotide binding sites.  相似文献   

4.
An analog of adenosine 5′-triphosphate (ATP) was synthesized in which the C4′---C5′---O---Pα system is replaced by a trans C4′---CH=CH---Pα system. In the form of 1:1 complexes with Mg, this analog and its counterpart with a C4′---CH2---CH2---Pα system were linear competitive inhibitors, with respect to MgATP, of the MAT-II (normal tissue) and MAT-T (hepatoma tissue) forms of rat ATP: -methionine-S-adenosyltransferase (MAT); Km(ATP)/Ki values ranged from 0.4 to 2.4. 2′-Deoxy-ATP was a weak substrate, Km(ATP)/Km = 0.035, of MAT-II and a weak competitive inhibitor, Km(ATP)/Ki = 0.07, of MAT-T. These findings, together with interactions of the MAT forms with other substrates and inhibitors, indicate that binding of ATP to these transferases is accompanied by little rotation about the C5′---O5′ bond, and that C4′ and Pα are in a trans-type relationship in enzyme-bound ATP.  相似文献   

5.
Summary The effects of several sulfamoyl benzoic acid derivatives on Na–K–Cl cotransport were investigated in winter flounder intestine. The relative efficacy (IC50 values) and order of potency of these derivatives were benzmetanide, 5×10–8 m> bumetanide 3×10–7 m>piretanide 3×10–6 m>furosemide 7×10–6 m> amino piretanide 1×10–5 3-amino-4-penoxy-5-sulfamoyl benzoic acid. Binding of [3H] bumetanide was studied in microsomal membranes from winter flounder intestine and compared to that in bovine kidney outer medulla. Binding was also studied in brush-border membranes from winter flounder intestine. The estimated values forK d and number of binding sites (n) were: bovine kidney,K d =1.6×10–7,n=10.5 pmol/mg protein; winter flounder intestine,K d 1.2×10–7,n=7.3 pmol/mg protein, and brush-border membranes from winter flounder,K d =5.3×10–7,n=20.4 pmol/mg protein. The estimatedK d for bumetamide binding to winter flounder brush-border membranes derived from association and dissociation kinetics was 6.8×10–7 m. The similarity in magnitudes of IC50 andK d for bumetanide suggests that the brush-border cotransporter is ordinarily rate-limiting for transmural salt absorption and that bumetanide specifically binds to the cotransporter. Measurement of bumetanide binding at various concentrations of Na, K and Cl showed that optimal binding required all three ions to be present at about 5mm concentrations. Higher Na and K concentrations did not diminish binding but higher Cl concentrations (up to 100mm Cl) inhibited bumetanide binding by as much as 50%. Still higher Cl concentrations (500 and 900mm) did not further inhibit bumetanide binding. Scatchard analysis of bumetanide binding at 5 and 100mm Cl concentrations showed that bothK d andn were lower at the higher Cl concentration (5mm Cl:K d =5.29×10–7 m,n=20.4 pmol/mg protein; 100mm Cl:K d =2.3×10–7 m,n=8.8 pmol/mg protein). These data suggest two possibilities: that bumetanide and Cl binding are not mutually exclusive (in contrast to pure competitive inhibition) and that they each bind to separate sites or that two distinct bumetanide binding sites exist, only one of which exhibits Cl inhibition of binding. This inhibition would then be consistent with a competitive interaction with Cl.  相似文献   

6.
The role of tryptophan (Trp17) in immunoreactivity of P1, the diagnostically relevant peptide from a major allergen/antigen of Aspergillus fumigatus, was evaluated by chemically modifying tryptophanyl residue of P1. In BIAcore kinetic studies, unmodified P1 showed a 100-fold higher binding with ABPA (Allergic Bronchopulmonary Aspergillosis) patients’ IgG [KD (equilibrium dissociation constant) = 2.74 e−8 ± 0.13 M] than the controls’ IgG (KD = 2.97 e−6± 0.14 M), whereas chemically-modified P1 showed similar binding [KD patients’ IgG = 3.25 e−7± 0.16 M, KD controls’ IgG = 3.86 e−7± 0.19 M] indicating loss of specific immunoreactivity of P1 on tryptophan modification. Modified P1 showed loss of specific binding to IgE and IgG antibodies of ABPA patients in ELISA (Enzyme-Linked Immunosorbent Assay). The study infers that tryptophan residue (Trp17) is essential for immunoreactivity of P1.  相似文献   

7.
Affinity capillary electrophoresis was used to detect a shift in mobility when a zinc ion binds to the highly basic nucleocapsid protein (NCp7) of HIV-1. NCp7 contains two Cys-X2- Cys-X4-His-X4-Cys zinc fingers. With constant concentrations of NCp7 as a receptor and various concentrations of zinc as a ligand in the sample buffer and the electrophoresis buffer, we observed changes in electrophoretic mobilities of NCp7 protein when complexes were formed with zinc. Scatchard analysis of the mobility indicates the presence of at least two types of binding sites for zinc. At pH 6.0, one site is shown to bind zinc strongly with a binding constantKb= 3.25 × 105M−1and the second site has aKb= 1.8 × 105M−1. The binding of zinc to the first zinc finger decreased the affinity of zinc for the second zinc finger approximately twofold. The Hill coefficient for this negative cooperativity is 0.9. A series of NCp7 mutants were also examined in the assay to determine their ability to bind zinc. This assay affords a quick method to observe a zinc ion binding to NCp7 and to calculate binding constants.  相似文献   

8.
Epinephrine (EPI) is thought to act by stimulating adenylyl cyclase (ACase) and cAMP production through β-adrenoceptors in the liver of more primitive vertebrates. Recent observations, however, point to an involvement of α1-adrenoceptors in EPI action, at least in some fish species. The role of the α1- and β-adrenergic transduction pathways has been investigated in rainbow trout (Oncorhynchus mykiss) hepatic tissue. Radioligand-binding assays with the β-adrenergic antagonist 3H-CGP-12177 using hepatic membranes purified on a discontinuous sucrose gradient confirmed the presence of β-adrenoceptors (Kd0.36 nM, Bmax 8.61 fmol · mg−1 protein). We provide the first demonstration of α1-adrenoceptors in these same membranes; analysis of binding data with the α1-adrenergic antagonist 3H-prazosin demonstrated a single class of binding sites with a Kdof 15.4 nM and a Bmax of 75.2 fmol · mg−1 protein. There is a straight correlation between β-adrenoceptor occupancy, ACase activation and cAMP production. On the contrary, the role of inositol 1,4,5-trisphosphate (IP3) has to be elucidated; in fact, despite the presence of specific microsomal binding sites for IP3 (Kd 6.03 nM, Bmax 90.2 fmol · mg−1 protein), its cytosolic concentration was not modulated by EPI. On the other hand, we have previously shown in American eel and bullhead hepatocytes that α1-adrenergic agonists are able to increase intracellular concentrations of IP3 and Ca2+ and to activate glycogenolysis. These data suggest a marked variation in the liver of different fish both in terms of α1-binding sites affinity and of α1-adrenoceptor/IP3/Ca2+ transduction systems.  相似文献   

9.
Summary An L1210 cell line (JT-1), which can grow in medium supplemented with 1nm folate, has been isolated. These cells exhibit a slower growth rate than folate-replete parental cells and have a lower ability to transport folate or methotrexate via the reduced folate transport system. Measurements at nanomolar concentrations of folate revealed that the adapted cells have acquired a high-affinity folate-binding protein. Binding to this component at 37°C was rapid and reached a maximum value after 30 min which corresponded in amount to 0.23±0.3 pmol/mg protein, and excess unlabeled folate added 30 min subsequent to the [3H]folate led to a rapid release of the bound substrate. Radioactivity bound to or released from the cells after 30 min at 37°C remained as unmetabolized folic acid. Binding was also rapid at 0°C but uptake at the plateau was only one-half the value obtained at 37°C. Half-maximal saturation of the binding component (K D) occurred at a folate concentration of 0.065nm at pH 7.4, while the affinity for folate decreased 30-fold when the pH was reduced to 6.2 (K D=2.0nm). 5-Methyltetrahydrofolate was also bound by this component (K i=13nm at pH 7.4) but with a much lower affinity than for folate, while progressively weaker interactions were observed with 5-formyltetrahydrofolate (K i=45nm) and methotrexate (K i=325nm). When the same adaptation procedure was performed with limiting amounts of 5-formyltetrahydrofolate, two additional cell lines, JT-2 and JT-3, were isolated which expressed elevated levels of the folate-binding protein. The binding activity of the latter cells was 0.46 and 1.4 pmol/mg protein, respectively. When the level of binding protein was compared in cells grown at different concentrations of folate, an increase in medium folate from 1 to 500nm caused a sevenfold reduction in binding activity in the JT-3 cell line, while these same growth conditions had no effect on binding by the other cells. These results indicate that L1210 cells adapted to low concentrations of folate or 5-formyltetrahydrofolate contain elevated levels of a high-affinity binding protein and that this protein is able to mediate the intracellular accumulation of folate compounds. L1210 cells thus appear to have two potential uptake routes for folate compounds, the previously characterized anion-exchange system and a second route mediated by a high-affinity binding protein. An additional low-affinity, high-capacity transport system for folate that had been proposed previously was not observed under a variety of experimental conditions in either the adapted or parental cells.  相似文献   

10.
Identifying and characterizing protein interactions are fundamental steps toward understanding and modeling biological networks. Methods that detect protein interactions in intact cells rather than buffered solutions are likely more relevant to natural systems since molecular crowding events in the cytosol can influence the diffusion and reactivity of individual proteins. One in vivo, imaging-based method relies on the colocalization of two proteins of interest fused to DivIVA, a cell division protein from Bacillus subtilis, and green fluorescent protein (GFP). We have modified this imaging-based assay to facilitate rapid cloning by constructing new vectors encoding N- and C-terminal DivIVA or GFP molecular tag fusions based on site-specific recombination technology. The sensitivity of the assay was defined using a well-characterized protein interaction system involving the eukaryotic nuclear import receptor subunit, Importin α (Impα), and variant nuclear localization signals (NLS) representing a range of binding affinities. These data demonstrate that the modified colocalization assay is sensitive enough to detect protein interactions with Kd values that span over four orders of magnitude (1 nM to 15 μM). Lastly, this assay was used to confirm numerous protein interactions identified from mass spectrometry-based analyses of affinity isolates as part of an interactome mapping project in Rhodopseudomonas palustris.  相似文献   

11.
Designed protein receptors hold diagnostic and therapeutic promise. We now report the design of five consensus leucine‐rich repeat proteins (CLRR4–8) based on the LRR domain of nucleotide‐binding oligomerization domain (NOD)‐like receptors involved in the innate immune system. The CLRRs bind muramyl dipeptide (MDP), a bacterial cell wall component, with micromolar affinity. The overall K d app values ranged from 1.0 to 57 μM as measured by fluorescence quenching experiments. Biphasic fluorescence quenching curves were observed in all CLRRs, with higher affinity K d1 values ranging from 0.04 to 4.5 μM, and lower affinity K d2 values ranging from 3.1 to 227 μM. These biphasic binding curves, along with the docking studies of MDP binding to CLRR4, suggest that at least two MDPs bind to each protein. Previously, only single MDP binding was reported. This high‐capacity binding of MDP promises small, soluble, stable CLRR scaffolds as candidates for the future design of pathogen biosensors.  相似文献   

12.
N-Methylpurine-DNA glycosylase (MPG), a ubiquitous DNA repair enzyme, initiates excision repair of several N-alkylpurine adducts, deaminated and lipid peroxidation-induced purine adducts. MPG from human and mouse has previously been cloned and expressed. However, due to the poor expression level in Escherichia coli (E. coli) and multi-step purification process of full-length MPG, most successful attempts have been limited by extremely poor yield and stability. Here, we have optimized the codons within the first five residues of human MPG (hMPG) to the best used codons for E. coli and expressed full-length hMPG in large amounts. This high expression level in conjunction with a strikingly high isoelectric point (9.65) of hMPG, in fact, helped purify the enzyme in a single step. A previously well-characterized monoclonal antibody having an epitope in the N-terminal tail could detect this codon-optimized hMPG protein. Surface plasmon resonance studies showed an equilibrium binding constant (KD) of 0.25 nM. Steady-state enzyme kinetics showed an apparent Km of 5.3 nM and kcat of 0.2 min−1 of MPG for the hypoxanthine (Hx) cleavage reaction. Moreover, hMPG had an optimal activity at pH 7.5 and 100 mM KCl. Unlike the previous reports by others, this newly purified full-length hMPG is appreciably stable at high temperature, such as 50 °C. Thus, this study indicates that this improved expression and purification system will facilitate large scale production and purification of a stable human MPG protein for further biochemical, biophysical and structure–function analysis.  相似文献   

13.
The effects of brain ischemia on the maximum binding capacity (Bmax) and affinity (Kd) of A1 receptors were studied in the rat cerebral cortex, with an in vitro approach. The results were correlated with changes in 3H-adenosine release, studied under identical experimental conditions. Fifteen minutes of in vitro ‘ischemia’ (hypoxic, glucose-free medium) induced a significant increase in both Bmax (2398±132 fmol/mg protein, 151% of the control, P<0.05) and in Kd (2.43±0.12 nM, 161% of the control, P<0.01). At the same time, an increase in tritium efflux from [3H]-adenosine labeled cerebral cortex slices to 324% of the control was observed. A trend toward normalization was evident 5–15 min after ‘reoxygenation’ (restoring normal medium), but the binding parameters were still altered after 60 min (Bmax 2110±82 fmol/mg protein, Kd 2.26±0.14 nM, P<0.01 vs the corresponding control) as was adenosine release (196% of the control). These findings suggest that the increased availability of adenosine and its receptors may be a defense mechanism against ischemic injury, while the reduced affinity of A1 receptors, possibly due to desensitization, may be a sign of ischemia-induced cellular damage.  相似文献   

14.
Summary Glutamine uptake by rat renal brushborder vesicles occurred via two distinct saturable processes withK m values of 0.145 and 8.5 mM which were stimulated by both ionic and sodium gradients with a pH optimum of 6.8–7.1 Glutamic acid uptake also occurred by a two-component system withK m values of 0.016 and 3.60 mM. Both components were stimulated specifically by a sodium gradient. The lowK m system for glutamic acid had a pH optimum of 7.2–7.4. Glutamine entry at 0.06 mM was inhibited by a variety of amino acids at 3 mM, including dibasic amino acids, glycine, valine, and phenylalanine. Glutamic acid entry at 0.06 mM was inhibited 20–30% by 3 mM phenylalanine, valine, -aminoisobutyric acid, and glutamine. No metabolic alteration of glutamic acid was observed on incubation with membrane vesicles, but glutamine was significantly hydrolyzed to glutamic acid upon prolonged incubation. Hydrolysis of glutamine was negligible at 15 sec incubation which was employed for measurement of initial rate of entry. These studies provide support for the existence of an uptake system in the brushborder of the renal proximal tubule cell capable of handling the reabsorption of glutamine normally present in glomerular filtrate.  相似文献   

15.
16.
A low-Mr tight binding proteinase inhibitor was purified from bovine muscle by alkaline denaturation of cysteine proteinases, gel filtration on Sexphadex G-75 and affinity chromatography on carboxymethyl-papain-Sepharose. Chromatofocusing separated three isoforms which are similar in their Mr of about 14 000, their stability with heating at 80°C and their inhibitory activity towards cathepsin H, cathepsin B and papain. The equilibrium constants (Ki) were determined for these three cysteine proteinases but for cathepsin H, association (kass) and dissociation (kdiss) rate constants were also evaluated. Ki values of 56 nM and 8.4 nM were found for cathepsin B and cathepsin H, respectively. For papain, Ki was in the range of 0.1–1 nM. The kinetic features of enzyme-inhibitor binding suggest a possible role for this low-Mr protein inhibitor in controlling ‘in vivo’ cathepsin H proteolytic activity. With regard to cathepsin B, such a physiological role was less evident.  相似文献   

17.
The ability of methyllycaconitine (MLA) to inhibit the binding of [125I]α-bungarotoxin to rat brain membranes, frog and human muscle extracts and the human muscle cell line TE671 has been measured. MLA showed a markedly higher affinity for the rat brain site (Ki 1.4 × 10−9 M) than for the muscle receptors (Ki; 10−5-10−6 M). Structure modelling techniques were used to fit the structure of MLA to a nicotinic pharmacophore model. MLA is the first low molecular weight ligand to be shown to discriminate between muscle nicotinic receptors and their α-bungarotoxinbinding counterpart in the brain, and as such may be a useful structural probe for pursuing the structural and functional properties of the neuronal protein.  相似文献   

18.
Typical physico-chemical studies of metal binding proteins are usually aimed at determination of the metal binding constant K for a native protein (K n), while the significance of the K value for the thermally denatured protein (K u) is usually underestimated. Meanwhile, metal binding induced shift of thermal denaturation transition of a single site metal binding protein is defined by K n to K u ratio, implying that knowledge of both K values is required for full characterization of the system. In the present work, the most universal approach to the studies of single site metal binding proteins, namely construction of a protein “phase diagram” in coordinates of free metal ion concentration – temperature, is considered in detail. The detailed algorithm of construction of the phase diagrams along with underlying mathematic procedures developed here may be of use for studies of other simple protein-target type systems, where target represents low molecular weight ligand. Analysis of the simplest protein-ligand system reveals that thermodynamic properties of apo-protein dictate the maximal possible increase of its affinity to any simple ligand upon thermal denaturation of the protein. Experimental and general problems coupled with the use of the phase diagrams are discussed.  相似文献   

19.
Summary The fluorochrome AMHA (3-amino-6-methoxy-9-(2-hydroxyethylamino)acridine) stains the nuclear chromatin and the chromosomes of living HeLa cells. At relatively low dye concentrations C F10–4 M and short incubation periods t I2 h cell growth is not affected by the drug. But at higher C F and longer t I the population doubling time of the cell cultures rapidly increases, and finally the cells die.In vital staining experiments the dye AMHA preferentially binds to the DNA of the nuclei and to the chromosomes of the cells, respectively. The dye binding to DNA has been proved by the absorption and emission microspectra of the stained cells, and by the comparison with authentic spectra of AMHA bound to DNA in aqueous solutions. Within the limits of experimental errors both types of spectra are identical. The spectra of DNA-bound AMHA show a characteristic gap of ca. 3500 cm–1 between the 0-0-transitions of the long wave length 1 L a absorption and the fluorescence. AMHA molecules dissolved in the polar solvent water have a gap of even 4100 cm–1. This energy gap shows that the electron distribution of AMHA is strongly changed by light absorption and emission.Finally, using absorption spectroscopy, we investigated the binding of AMHA to DNA in aqueous solutions over a wide range of concentrations of the dye, of nuceleic acid (calf thymus), and of the competitor NaCl respectively. The Scatchard binding isotherms were determined. With the method of competitive salt effect three different bonds of AMHA to DNA can be distinguished even at low dye concentrations: The intercalation 1 of the fluorochrome F, binding constant K F1=1,1·105 M –1, binding parameter n 1=0,15; the pre-intercalative or external binding 2, K F2=6,9·105 M –1, n 2=0,21; the external binding 3, K F3=2,8·105 M –1, n 3=0,55. Externally bound dye molecules 2 and 3 occupy two phosphodiester residues of the DNA. A detailed discussion of the data and the competitive salt effect shows that in living cells only intercalated and small amounts of pre-intercalatively bound molecules 1 and 2 exist. The binding constant K F1=1,1·105 M –1 of AMHA is unusual high in comparison with the constants of intercalation of other dyes, K F1=(1–4)·104 M –1. Therefore, the amount of intercalated AMHA is also relatively high, and it is possible to visualize the DNA-bound fluorochrome in the nuclei and chromosomes of the living cells under the fluorescence microscope.  相似文献   

20.
Chemical shift perturbations (CSPs) in NMR spectra provide useful information about the interaction of a protein with its ligands. However, in a multiple‐ligand‐binding system, determining quantitative parameters such as a dissociation constant (Kd) is difficult. Here, we used a method we named CS‐PCA, a principal component analysis (PCA) of chemical shift (CS) data, to analyze the interaction between bovine β‐lactoglobulin (βLG) and 1‐anilinonaphthalene‐8‐sulfonate (ANS), which is a multiple‐ligand‐binding system. The CSP on the binding of ANS involved contributions from two distinct binding sites. PCA of the titration data successfully separated the CSP pattern into contributions from each site. Docking simulations based on the separated CSP patterns provided the structures of βLG–ANS complexes for each binding site. In addition, we determined the Kd values as 3.42 × 10−4M2 and 2.51 × 10−3M for Sites 1 and 2, respectively. In contrast, it was difficult to obtain reliable Kd values for respective sites from the isothermal titration calorimetry experiments. Two ANS molecules were found to bind at Site 1 simultaneously, suggesting that the binding occurs cooperatively with a partial unfolding of the βLG structure. On the other hand, the binding of ANS to Site 2 was a simple attachment without a significant conformational change. From the present results, CS‐PCA was confirmed to provide not only the positions and the Kd values of binding sites but also information about the binding mechanism. Thus, it is anticipated to be a general method to investigate protein–ligand interactions. Proteins 2013. © 2012 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