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 interactions between cytochrome P450 2B4 (d-2B4), NADPH:cytochrome P450 reductase and cytochrome b5 have been investigated in the monomeric reconstituted P450 2B4-containing monooxygenase system in the presence of a substrate (7-pentoxyresorufin) and an electron donor, NADPH. Each partner was immobilized via its amino groups on the carboxymethyldextran biochip surface of the optical biosensor IAsys+. Such mode immobilization was not accompanied by any loss of activities of the immobilized proteins. The formation of binary d-Fp/d-2B4 complexes was registered. The association/dissociation rate constants (kon/koff) were (0.013 ± 0.005) × 106 M?1 s?1/0.05 ± 0.02 s?1, and dissociation constant (KD) was (0.26 ± 0.13) × 10?6 M. Comparison of kon, koff and KD values for d-Fp/d-2B4 complexes formed under hydroxylation (O-dealkylation) with corresponding constants obtained for the oxidized proteins of (0.10 ± 0.03) × 106 M?1 s?1/(0.14 ± 0.06) s?1, and (0.71 ± 0.37) × 10?6 M, respectively shows that the decrease in kon and an insignificant decrease in KD are associated with the increase of complex lifetime during transition from the oxidized to hydroxylation conditions. Complex formation between d-Fp and d-b5 was not registered in both hydroxylation conditions and in the case of oxidized forms of these proteins. In both cases formation of the ternary d-Fp/d-2B4/d-b5 complexes occurred.  相似文献   

2.
3.
Green tea is rich in several polyphenols, such as (?)-epicatechin-3-gallate (ECG), (?)-epigallocatechin (EGC), and (?)-epigallocatechin-3-gallate (EGCG). The biological importance of these polyphenols led us to study the major polyphenol EGCG with human serum albumin (HSA) in an earlier study. In this report, we have compared the binding of ECG, EGC, and EGCG and the Cu(II) complexes of EGCG and ECG with HSA. We observe that the gallate moiety of the polyphenols plays a crucial role in determining the mode of interaction with HSA. The binding constants obtained for the different systems are 5.86?±?0.72?×?104 M?1 (K ECG-HSA), 4.22?±?0.15?×?104 M?1 (K ECG-Cu(II)-HSA), and 9.51?±?0.31?×?104 M?1 (K EGCG-Cu(II)-HSA) at 293?K. Thermodynamic parameters thus obtained suggest that apart from an initial hydrophobic association, van der Waals interactions and hydrogen bonding are the major interactions which held together the polyphenols and HSA. However, thermodynamic parameters obtained from the interactions of the copper complexes with HSA are indicative of the involvement of the hydrophobic forces. Circular dichroism and the Fourier transform infrared spectroscopic measurements reveal changes in α-helical content of HSA after binding with the ligands. Data obtained by fluorescence spectroscopy, displacement experiments along with the docking studies suggested that the ligands bind to the residues located in site 1 (subdomains IIA), whereas EGC, that lacks the gallate moiety, binds to the other hydrophobic site 2 (subdomain IIIA) of the protein.  相似文献   

4.
We studied interaction of the lectin from the bark of Golden Rain shrub (Laburnum anagyroides, LABA) with a number of basic fucose-containing carbohydrate antigens by changes in its tryptophan fluorescence. The strongest LABA binding was observed for the trisaccharide H of type 6 [α-L-Fucp-(1-2)-β-D-Galp-(1-4)-D-Glc, K a = 4.2 × 103 M?1]. The following antigens were bound with a weaker affinity: H-disaccharide α-L-Fucp-(1-2)-D-Gal, a glucoanalogue of tetrasaccharide Ley α-L-Fucp-(1-2)-β-D-Galp-(1-4)-[α-L-Fucp-(1-3)]-D-Glc, and 6-fucosyl-N-acetylglucosamine, a fragment of core of the N-glycans family (K a 1.1?1.7 × 103 M?1). The lowest binding was observed for L-fucose (K a = 2.7 × 102 M?1) and trisaccharide Lea, (β-Galp-(1-3)-[α-L-Fucp-(1-4)]-GlcNAc (K a = 6.4 × 102 M?1). The Led, Lea, and Lex pentasaccharides and Leb hexasaccharide were not bound to LABA.  相似文献   

5.
The kinetics of uptake and retention of β-ecdysone by imaginal discs from late third instar larvae of Drosophila melanogaster correspond well with those of the first synthetic response of discs to hormone, an increase in RNA synthesis.Competition studies indicate the presence of two types of hormone binding sites, specific and non-specific. The specific sites are saturated at hormone concentrations which fully induce morphogenesis. Results are consistent with the hypothesis that analogs which induce morphogenesis at differing concentrations bind to the same sites. Experiments with the inhibitors N-ethylmaleimide, actinomycin d, and cycloheximide suggest that the binding sites are pre-existing in the cell and require functional sulfhydryl groups for binding.Specific binding, binding that is competed by excess unlabeled β-ecdysone, is saturable (70–80 nM). Kinetic rate constants for this specific binding were estimated to be ka = 1.5 × 105M?1 min?1, kd = 3 × 10?2 min?1. The equilibrium dissociation constant calculated from the kinetic rate constants was Keq = 2 × 10?7M compared to 1.7 × 10?7M β-ecdysone required to induce morphogenesis in vitro and 2.5 × 10?7M determined to be the in vivo concentration at the time of induction of morphogenesis.  相似文献   

6.
The DNA-binding and photonuclease activity of newly synthesized tetra-azamacrocyclic ligand L (C32H32N8O4) and its complexes of type [MLCl2] and [ML]Cl2 (where M = Co(II), Fe(II) and Cu(II); L = N,N′-[3-(4-{5-[(2-amino-ethylamino)-methyl]-isoxazol-3yl}-phenyl)-isoxazol-5-yl methyl-ethane-1,2-diamine] are specified. An octahedral geometry has been proposed for Fe(II) and Co(II) complexes, while the Cu(II) complex has a square planar environment. The absorption spectral results indicate that the complexes bind with the base pairs of DNA, with an intrinsic binding constant Kb of Fe(II), Co(II), and Cu(II) complexes found to be 3.2 × 104 M?1, 5.3 × 104 M?1, and 4.2 × 104 M?1, respectively, in 5 mM Tris-HCl/50 mM NaCl buffer at pH 7.2. The large enhancement in the relative viscosity of DNA on binding to the complexes supports the proposed DNA binding modes. The viscosity and thermal denaturation studies sustain the effective intercalation with DNA. The DNA photocleavage studies demonstrated that compounds exhibit significant photonuclease activity by a concentration dependent on singlet oxygen mediated mechanism.  相似文献   

7.
Abstract

Saffron is the red dried stigmas of Crocus sativus L. flowers and used both as a spice and as a drug in traditional therapeutic. The biological activity of saffron in modern medicine is in development. Its numerous applications as an anti-oxidant and anti-cancer agent are due to its secondary metabolites and their derivatives (safranal, crocins, crocetin, dimethylcrocetin). The aim of this study was to examine the interaction of transfer RNA with safranal, crocetin, and dimethylcrocetin in aqueous solution at physiological conditions. Constant tRNA concentration (6.25 mM) and various drug/tRNA (phosphate) molar ratios of 1/48 to 1/8 were used. FT-IR and UV-Visible difference spectroscopic methods have been applied to determine the drug binding mode, the binding constants and the effects of drug complexation on the stability and conformation of tRNA duplex. External binding mode was observed for safranal crocetin and dimethylcrocetin, with overall binding constants Ksafranal = 6.8 (± 0.34) × 103 M?1, KCRT = 1.4 (± 0.31) × 104 M?1, and KDMCRT = 3.4 (± 0.30) × 104 M?1. Transfer RNA remains in the A-family structure, upon safranal, crocetin and dimethylcrocetin complexation.  相似文献   

8.
Four complexes [Pd(L)(bipy)Cl]·4H2O (1), [Pd(L)(phen)Cl]·4H2O (2), [Pt(L)(bipy)Cl]·4H2O (3), and [Pt(L)(phen)Cl]·4H2O (4), where L = quinolinic acid, bipy = 2,2’-bipyridyl, and phen = 1,10-phenanthroline, have been synthesized and characterized using IR, 1H NMR, elemental analysis, and single-crystal X-ray diffractometry. The binding of the complexes to FS-DNA was investigated by electronic absorption titration and fluorescence spectroscopy. The results indicate that the complexes bind to FS-DNA in an intercalative mode and the intrinsic binding constants K of the title complexes with FS-DNA are about 3.5?×?104 M?1, 3.9?×?104 M?1, 6.1?×?104 M?1, and 1.4?×?105 M?1, respectively. Also, the four complexes bind to DNA with different binding affinities, in descending order: complex 4, complex 3, complex 2, complex 1. Gel electrophoresis assay demonstrated the ability of the Pt(II) complexes to cleave pBR322 plasmid DNA.  相似文献   

9.
The electron transfer reactions of horse heart cytochrome c with a series of amino acid-pentacyanoferrate(II) complexes have been studied by the stopped-flow technique, at 25°C, μ = 0.100, pH 7 (phosphate buffer). A second-order behavior was observed in the case of the Fe(CN)5 (histidine)3? complex, with k = 2.8 x 105 M?1 sec?1. For the Fe(CN)5 (alanine)4? and Fe(CN)5(L-glutamate)5? complexes, only a minor deviation of the second-order behavior, close to the experimental error (k = 3.2 × 105 and 1.6 x 105 M?1 sec?1, respectively) was noted at high concentrations of the reactants (e.g., 6 × 10?4 M). The results are in accord with recent work on the Fe(CN)64?/cytochrome c system demonstrating weak association of the reactants. The calculated self-exchange rate constants including electrostatic interactions for the imidazole,L -histidine, 4-aminopyridine, glycinate, β-alaninate, andL-glutamate pentacyanoferrate(II) complexes were 3.3 × 105, 3.3 × 105, 2.8 × 106,4.1 × 102,5.5 × 102, and 6.0 M?1 sec?1, respectively. Marcus theory calculations for the cytochrome c reactions were interpreted in terms of two nonequivalent binding sites for the complexes, with the metalloprotein self-exchange rate constants varying from 104 M?1 sec?1 (histidine, imidazole, and 4-aminopyridine complexes) to 106 M?1 sec ?1 (glycinate, β-alaninate, and L-glutamate complexes).  相似文献   

10.
Abstract

Most of the drugs binding to human serum albumin (HSA) are transported to various parts of the body. Here, we have studied the molecular interaction between HSA and synthesized uridine derivatives, 1-[(3R, 4S, 5?R)-2-methyl-3, 4-dihydroxy-5-(hydroxymethyl)oxolan-2-yl]pyrimidine-2,4-dion.)(C-MU); [(2R,3R,4R,5R)-5-(2,4-dioxo-3,4-dihydropyrimidin-1(2H)-yl)-3,4-dihydroxy-4-methyl-tetrahydrofuran-2-yl] methyl methyl phosphochloridate (CM-MU) and [(2R,3S,4R,5R)-5-(2,4-dioxopyrimidin-1-yl)-2-methyl-3,4-dihydroxyoxolan-2-yl] methyl dihydrogen phosphate (P-MU). Cytotoxic studies of these synthesized compounds with mouse macrophages (RAW 246.7) and HeLa cells (human cervical cancer cells) and binding mechanism of these uridine derivatives with HSA were performed. Subsequently, fluorescence quenching was observed upon titration of uridine derivatives with HSA via static mode of quenching, and the binding constants (K2-C-MU = 4?±?0.03?×?104M?1, K5-CM-MU = 1.95?±?0.03?×?104 M?1 and K5-P-MU =1.56?±?0.03?×?104 M?1) were found to be in sync with the computational results. Further, molecular displacement and molecular docking data revealed that all the derivatives are binding in the subdomain IIA and IIB regions of HSA. The protein secondary structure of complexes was determined by circular dichroism, indicating partial unfolding of the protein upon addition of the uridine derivatives. Furthermore, atomic force microscopy data reveal the change in topology upon binding of 2-C-MU, 5-CM-MU and 5-P-MU with HSA, indicating change in the microenvironment around tryptophan region. Additionally, cytotoxicity studies on HeLa and Raw Cell lines suggested that these molecules have significant anti-proliferative and anti-inflammatory properties. Hence, the study may be of help for development of new drugs based on uridine derivatives which may be helpful for combating various potential diseases.

Communicated by Ramaswamy H. Sarma  相似文献   

11.
Abstract

(±)125 I-cyanopindolol (±) I CYP) was used to characterize β-adrenoceptors on rat lung and cerebral cortex membranes. The affinity of (±) ICYP was higher for lung (Kd = 64.3 pM) at 37°C. The association reaction of (±) ICYP was faster with lung (k+1 = 1.52 × 109 M?1.min?1) than with cerebral cortex β-adrenoceptors (k+1 = 1.75 × 108 M?1.min?1). In both tissues, the dissociation reaction followed a biphasic process with a fast (t ½ = 15.4 min and 5.6 min for lung and cerebral cortex respectively) and a slow component (t ½ = 474 min and 255 min for lung and cerebral cortex respectively). The thermodynamic parameters for (±) ICYP - β-adrenoceptors binding have been determined from kinetics and equilibrium studies, for the two tissues, at several temperatures between 0° and 44° C. For lung and cerebral cortex, Arrhenius plots were linear with different energies of activation. Van't Hoff plot was not linear for lung and the standard enthalpy and entropy changes of (±) ICYP - β-adrenoceptors interaction decreased linearly with temperature : the binding occured with a negative heat capacity change (ΔCp° = -368.9 cal. moles?1. K?1) at 25° C. Thermodynamic and kinetic results show that binding of (±) ICYP to lung β-adrenoceptors could involve two successive equilibria with a conformational change of the β-adrenergic receptor.  相似文献   

12.
The three-dimensional model of the CtCBM35 (Cthe 2811), i.e. the family 35 carbohydrate binding module (CBM) from the Clostridium thermocellum family 26 glycoside hydrolase (GH) β-mannanase, generated by Modeller9v8 displayed predominance of β-sheets arranged as β-sandwich fold. Multiple sequence alignment of CtCBM35 with other CBM35s showed a conserved signature sequence motif Trp-Gly-Tyr, which is probably a specific determinant for mannan binding. Cloned CtCBM35 from Clostridium thermocellum ATCC 27405 was a homogenous, soluble 16 kDa protein. Ligand binding analysis of CtCBM35 by affinity electrophoresis displayed higher binding affinity against konjac glucomannan (K a = 2.5 × 105 M?1) than carob galactomannan (K a = 1.4 × 105 M?1). The presence of Ca2+ ions imparted slightly higher binding affinity of CtCBM35 against carob galactomannan and konjac glucomannan than without Ca2+ ion additive. However, CtCBM35 exhibited a low ligand-binding affinity K a = 2.5 × 10?5 M?1 with insoluble ivory nut mannan. Ligand binding study by fluorescence spectroscopy showed K a against konjac glucomannan and carob galactomannan, 2.4 × 105 M?1 and 1.44 × 105 M?1, and ΔG of binding ?27.0 and ?25.0 kJ/mol, respectively, substantiating the findings of affinity electrophoresis. Ca2+ ions escalated the thermostability of CtCBM35 and its melting temperature was shifted to 70°C from initial 55°C. Therefore thermostable CtCBM35 targets more β-(1,4)-manno-configured ligands from plant cell wall hemicellulosic reservoir. Thus a non-catalytic CtCBM35 of multienzyme cellulosomal enzymes may gain interest in the biofuel and food industry in the form of released sugars by targeting plant cell wall polysaccharides.  相似文献   

13.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

14.
Measurements of the equilibrium and temperature-jump u.v., visible, and induced c.d. spectra of Methyl Orange (MO) in the presence of cyclomalto-octaose (γ-cyclodextrin, γ-CD) have been carried out. Three mechanistic steps were detected through the temperature-jump data (25.0°):
where K1, K2, and K3 are 45 (±7), 2.0 (±1.1) × 106, and 6.1 (±2.5) × 103 dm3.mol?1, respectively, k2 = 9.4 (±5.1) × 109 dm3.mol?1.s?1, and k?2 = 4.8 (±0.8) × 103 s?1. The equilibrium u.v./visible data are also consistent with this reaction scheme. The high stability of the dimer inclusion complex (MO)2 · γ-CD compared to that of the monomer inclusion complex MO · γ-CD appears to be related to the annular diameter of γ-CD and demonstrates a degree of selectivity in cyclodextrin inclusion complexes. The (MO)2 · (γ-CD)2 complex also contains a dimer, included by both γ-CD molecules.  相似文献   

15.
Studies of the binding of Ni2+ to adenylyl-3',5'-adenosine (ApA) at pH 6-0 by ultraviolet spectrophotometry indicate the formation of a 1:1 complex in the presence of a large excess of metal ion. At 25 °C. and ionic strength μ = 0.5 M, the stability constant of Ni(ApA) is evaluated to be K = 2.6 (±0.6) M?1. The low stability is taken as evidence that the predominant complex species is one in which the ApA acts as a monodentate ligand, mainly through the adenine group. The rate constants for complex formation and dissociation, kf = 1430 M?1 s?1 and kb = 665 s?1 (25°C. μ = 0.5M). determined by the temperature-jump relaxation technique, are consistent with this interpretation. The binding strength of Ni2+ to poly(adenylic acid) [poly(A)] has been studied at pH 7.0 using murexide as an indicator of the concentration of free Ni2+. Within the concentration range [Ni2+ = 1 × 10?5 × 10?3 M the data can be represented in the form of a linear Scatchard plot. i.e., the process can be described as the binding of Ni2+ to one class of independent binding sites. The number of binding sites per monomer is 0.26, and the stability constant K = 8.2×103 M?1 (25°C μ = 0.1 M). In kinetic studies of the reaction of Ni2+ with poly(A), two relaxation effects due to complex formation were detected, one with a concentration-independent time constant of about 0.4 ms, the other with a concentration-dependent time constant in the millisecond range. The concentration dependence of the longer relaxation time can be accounted for by a three-step mechanism which consists of a fast second-order association reaction followed by two first-order steps. There is evidence, however, that the overall process is more complicated than expressed by the three-step mechanism.  相似文献   

16.
The binding of[Co(CN)6]3?, and that of[Fe(CN)6]3? and [Ru(CN)6]4? using a competitive method, to horse cytochrome c has been studied by 59 Co NMR spectroscopy. At I = 0.07 M, without added salt and in 2H2O at ph* 7.3 (measured in 2H2O) and 25°C, there are at least two binding sites on ferricytochrome c and ferrocytochrome c for [Co(CN)6]3?. Association constants were determined to be 2.0 ± 0.6 × 103M?1 and 1.5 ± 0.5 × 102M?1 respectively. with no effect of the oxidation state of the cytochrome. At higher ionic strength (I = 0.12 M adjusted with KCl the binding markedly decreased, and, although it was not possible to determine the precise binding stoichiometry and magnitude of association constants, it is clear that the association constants are ≤ 1.5 × 10tM?1 The binding of [Ru(CN)6]4? at I = 0.07, without added salt and in 2H2O at pH 1.3 and 23°C, was not precisely defined, but its binding strength relative to that of [Fe(CN)6]3? was determined. Extrapolating this to I = 0.12 (KCl) suggests that under these conditions the association constant for [Ru(CN)6]4? binding to ferricytochrome c is ≤ 3 × 102M?1.  相似文献   

17.
We determined the binding sites of curcumin (cur), resveratrol (res), and genistein (gen) with milk β-lactoglobulin (β-LG) at physiological conditions. Fourier transform infrared spectroscopy, circular dichroism, and fluorescence spectroscopic methods as well as molecular modeling were used to determine the binding of polyphenol–protein complexes. Structural analysis showed that polyphenols bind β-LG via both hydrophilic and hydrophobic contacts with overall binding constants of Kcurcumin–β-LG?=?4.4 (±?.4)?×?104 M?1, Kresveratrol–β-LG?=?4.2 (±?.2)?×?104 M?1, and Kgenistein–β-LG?=?1.2 (±?.2)?×?104?M?1. The number of polyphenol molecules bound per protein (n) was 1 (cur), 1.1 (res), and 1 (gen). Molecular modeling showed the participation of several amino acid residues in polyphenol–protein complexation with the free binding energy of ?12.67 (curcumin–β-LG), ?12.60 (resveratrol–β-LG), and ?10.68?kcal/mol (genistein–β-LG). The order of binding was cur?>?res?>?gen. Alteration of the protein conformation was observed in the presence of polyphenol with a major reduction of β-sheet and an increase in turn structure, causing a partial protein structural destabilization. β-LG might act as a carrier to transport polyphenol in vitro.  相似文献   

18.
Abstract

In this study, forskolin-loaded human serum albumin nanoparticles (FR-HSANPs) were successfully prepared by incorporation and affinity-binding methods. FR-HSANPs were characterized by transmission electron microscope that most of them are circular in shape and size is around 340?nm. The drug loading was more than 88% and further sustained release profiles were observed as it is 77.5% in 24?h time. Additionally, the cytotoxicity results with HepG2 cells indicated that FR-HSANPs showed significantly higher cytotoxicity and lower cell viability as compared to free forskolin (FR). Furthermore, to understand the binding mechanism of human serum albumin (HSA) with forskolin resulted from fluorescence quenching as a static mechanism and the binding constant is 6.26?±?0.1?×?104 M?1, indicating a strong binding affinity. Further, association and dissociation kinetics of forskolin–HSA was calculated from surface plasmon resonance spectroscopy and the binding constant found to be Kforskolin = 3.4?±?0.24?×?104 M?1 and also fast dissociation was observed. Further, we used circular dichroism and molecular dynamics simulations to elucidate the possible structural changes including local conformational changes and rigidity of the residues of both HSA and HSA–forskolin complexes.

Communicated by Ramaswamy H. Sarma  相似文献   

19.
Triplet flash photolysis techniques, coupled with quenching of the triplets by molecular oxygen, are utilized as probes of the microenvironment of polycyclic aromatic molecules bound covalently and non-covalently to DNA. The triplet-oxygen quenching properties of the following adducts in aqueous solutions at 25±1°C were investigated: covalent adducts derived from the reaction of (±)-7β,8α-dihydroxy-9α,10α-epoxy -7,8,9,10-tetrahydrobenzo[a]pyrene (BaPDE) and of (±)-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BaPE) with DNA, and non-covalent intercalation complexes of acridine orange (AO) and DNA. In all cases the quenching follows the Stern-Volmer quenching law with a quenching constant of KO2T≈109 M?1·s?1 for the covalent BaPDE-DNA and BaPE-DNA complexes in aqueous solution. This value of KO2T is characteristic of free molecules (not bound to DNA) and indicates that the pyrene chromophore is totally accessible to oxygen, and is thus not located at an intercalation-type of binding site in these covalent adducts. In contrast, the AO-DNA complexes are characterized by values of KO2T≈108 M?1·s?1 indicating that the intercalated AO molecules are about ten times less accessible to molecular oxygen than free AO molecules. The KO2T values for the covalent BaPDE-DNA and BaPE-DNA adducts decrease when the DNA concentration is increased in the 1·10?4?3·10?3 M range (expressed in nucleotide concentration). This effect is attributed to intermolecular DNA-DNA interactions in which segments of adjacent DNA molecules tend to cover the pyrene chromophores on other strands, thus decreasing their accessibility to oxygen. In contrast the values of KO2T for the non-covalent AO-DNA intercalation complexes are independent of DNA concentration, as expected for interior binding sites.  相似文献   

20.
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   

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