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 187 毫秒
1.
Human growth hormone binding sites from female rabbit kidney microsomes were solubilized by treatment with the nonionic detergent Triton X-100. The binding of 125I-labelled human growth hormone to the solubilized sites retains many of the properties observed in the particulate fraction, such as saturability, reversibility, high affinity and structural specificity. The association and the dissociation process are time- and temperature-dependent. The association rate constant, k1, is 1.6·107 mol?1·l·min?1 at 25°C, and the dissociation rate constant, k?1, is 2.8·10?4 min?1 at 25°C. Solubilization causes an increase in affinity as well as in binding capacity. Scatchard plots from saturation curves suggest the presence of a single class of binding site with a dissociation equilibrium constant, Kd, of 1.3·10?11 M and a binding capacity of 133 fmol/mg of protein. Similar results were obtained from competition experiments. Specificity studies revealed the lactogenic characteristics of the solubilized sites. The Stokes radii of the free binding sites and of the 125I-labelled human growth hormone-binding site complex, determined on a Sepharose CL-6B column, are 57 and 53 Å, respectively.  相似文献   

2.
The pH dependence of the oxidation of β-methyl-d-galactopyranoside by galactose oxidase at 1.33 mm O2 has been determined. The kcat exhibits a bell-shaped dependence on the ionization of at least two groups in the enzyme-substrate complex, pKb' = 6.3 and pKa' = 7.1, respectively. The pH-independent value for kcat at 1.33 mm O2 (nonsaturating) and saturating glycoside is 1435 s?; the pH optimum is 6.7. Galactose oxidase is inactivated rapidly by iodoacetamide. Although the reaction is much slower, iodoacetate also inactivates the enzyme. The inactivation by iodoacetamide obeys saturation kinetics; at pH 7.0 k3 = 2.19 min?1 and Ki = 5.1 mM; k3 but not Ki exhibits a bell-shaped pH dependence, with pKa values of 6.3 and 7.6, respectively. Labeling with [14C]iodoacetamide establishes that one carboxamidomethyl group is incorporated per enzyme molecule. This incorporation parallels the loss of enzymatic activity. Only N-3-carboxymethylhistidine is detected in chromatograms following hydrolysis of the labeled protein. The protein-bound copper is not lost as a consequence of alkylation. Apogalactose oxidase does not react with iodoacetamide. The alkylation is inhibited by the oxidation of an active center tryptophan residue (s) by N-bromosuccinimide. The fraction of residual enzyme activity remaining after tryptophan oxidation corresponds to the extent of labeling by [14C]iodoacetamide. Although alkylation causes little change in the spin Hamiltonian parameters of the Cu(II) atom, it nearly abolishes both the optical activity and optical absorbance of the metal. The native tryptophan fluorescence of the enzyme, which is a sensitive probe of its active site, is also markedly affected. Since binding of a substrate, β-methyl-d-galactopyranoside, reduces fluorescence as it does in the active enzyme and binding of CN? at the Cu(II) site as detected by electron spin resonance appears unaffected by the alkylation, the effect of alkylation is on catalysis, per se. Both a catalytic and a subtle conformational role for the active site histidine are inferred from the results.  相似文献   

3.
The binding of methyl isonitrile (CH3Nandz.tbnd;C) to hemoglobin β chains has been studied by measuring the 1H nuclear magnetic resonance transverse relaxation times for methyl isonitrile as a function of protein concentration, temperature and 14N decoupling. Binding of methyl isonitrile both at the heme iron and at a non-specific site (or sites) has an effect upon the measured nuclear spin relaxation times. The results yield a value of 57 ± 12 seconds?1 (20 °C) for the “off” rate constant K?1 for specific binding and an Arrhenius activation energy for k?1 of 14 ± 3 kcal mol?1.  相似文献   

4.
The regulation of the rate of ATP hydrolysis by H-meromyosin   总被引:3,自引:0,他引:3  
The effect of N-ethylmaleimide on the ATPase activity and ADP binding of tryptic H-meromyosin was studied at 6 and 23 °C temperatures. The affinity constant of H-meromyosin for ADP with Mg as activator was increased by small concentrations of N-ethylmaleimide (2.25 moles per mole of enzyme) at both temperatures, accompanied by activation of ATP hydrolysis at 25 °C and inhibition at 6 °C. With higher N-ethylmaleimide concentrations, the ATPase activity was inhibited at both temperatures, without comparable inhibition of ADP binding. Rapid kinetic analysis of the rate of development of difference spectrum after the addition of ATP or ADP to H-meromyosin indicates, that blocking of the S1 and S2 SH groups of H-meromyosin decreases both the formation (k1) and the dissociation (k2) rate constants of H-meromyosin substrate complex. At 6 °C, in the presence of Mg, the value of k2 for ADP is similar to the turnover number of ATP hydrolysis, suggesting that dissociation of ADP from the active site may be the rate-limiting step of ATP hydrolysis. At 23 °C, the turnover number of Mg-moderated ATP hydrolysis is much smaller than k2, indicating that the rate limitation shifted so another, so far unidentified, step.  相似文献   

5.
The binding of cations of β-casein at pH 6.6 was considered previously. Available for three sodium concentiations, I = 0.04, 0.08, or 0.16 M are: [1] proton releases between I and [2] for each I, as calcium activity is increased, correlated sequences of monomer net charge, proton release, site bound calcium and protein Solvation- Models for ion binding are examined. Critical considerations are the intrinsic binding constants between hydrogen[H], calcium[Ca] and sodium[Na] ions and phosphate[P] and caiboxyIate[C] sites, and the effects of electrostatic interaction between sites as influenced by spatial fixed charge distribution, ionic strength and dielectric constant [D]. Anticipated intrinsic binding constants are kH,Po = 3 × 106, kCa,Po = 120, kNa,Po = 1, kH,Co = 7 × 104 and kCa,Co = 5.6Distributed charge models, either surface or volume, are inadequate since any reasonable monomer size yields fixed charge densities requiring kH,Po and kCa,Co which are too low when the maximum in D is 75. Also, with increasing calcium binding, calculated proton release is only 0.4 to 0.5 of that observed.Discrete charge models accept anticipated ko and yield calculated sequences of calcium binding and proton release which are in good agreement with those observed provided that: (1) using the known amino acid sequence of the phosphate-containing acidic peptide portion of the molecule, pep tide fixed charge is distributed at the lowest I so as to minimize electrostatic free energy; (2) in the region of fixed charge, D is approximately 5; (3) the distances between peptide fixed charges decrease with increasing ionic strength or calcium binding and (4) while protein is in solution, the acidic peptide and the remainder of the molecule are essentially electrostatically independent.  相似文献   

6.
Understanding enzyme catalysis through the analysis of natural enzymes is a daunting challenge—their active sites are complex and combine numerous interactions and catalytic forces that are finely coordinated. Study of more rudimentary (wo)man-made enzymes provides a unique opportunity for better understanding of enzymatic catalysis. KE07, a computationally designed Kemp eliminase that employs a glutamate side chain as the catalytic base for the critical proton abstraction step and an apolar binding site to guide substrate binding, was optimized by seven rounds of random mutagenesis and selection, resulting in a > 200-fold increase in catalytic efficiency. Here, we describe the directed evolution process in detail and the biophysical and crystallographic studies of the designed KE07 and its evolved variants. The optimization of KE07's activity to give a kcat/KM value of ∼ 2600 s− 1 M− 1 and an ∼ 106-fold rate acceleration (kcat/kuncat) involved the incorporation of up to eight mutations. These mutations led to a marked decrease in the overall thermodynamic stability of the evolved KE07s and in the configurational stability of their active sites. We identified two primary contributions of the mutations to KE07's improved activity: (i) the introduction of new salt bridges to correct a mistake in the original design that placed a lysine for leaving-group protonation without consideration of its “quenching” interactions with the catalytic glutamate, and (ii) the tuning of the environment, the pKa of the catalytic base, and its interactions with the substrate through the evolution of a network of hydrogen bonds consisting of several charged residues surrounding the active site.  相似文献   

7.
The DNA binding properties of the araC protein in the absence of l-arabinose have been studied in Escherichia coli using the nitrocellulose membrane filter technique. Equilibrium competition experiments demonstrate that the araC protein binds specifically to the ara operator. The apparent Km of the interaction is 1 × 10?12m at 20 °C. The rates of association and dissociation of the complex have also been determined. A ka of 2 × 109m?1 s?1at 20 °C is calculated assuming binding to a single site. The half-life of the complex is three minutes. The equilibrium constant calculated from the ratio of ka to kd is 2.8 × 10?12m at 20 °C. The good agreement between the equilibrium and kinetic determinations of the equilibrium constant suggest that the kinetic studies are providing true rate constants. It is calculated that about 1% of the purified araC protein is active with respect to operator binding activity.  相似文献   

8.
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   

9.
DmmA is a haloalkane dehalogenase (HLD) identified and characterized from the metagenomic DNA of a marine microbial consortium. Dehalogenase activity was detected with 1,3-dibromopropane as substrate, with steady-state kinetic parameters typical of HLDs (Km = 0.24 ± 0.05 mM, kcat = 2.4 ± 0.1 s−1). The 2.2-Å crystal structure of DmmA revealed a fold and active site similar to other HLDs, but with a substantially larger active site binding pocket, suggestive of an ability to act on bulky substrates. This enhanced cavity was shown to accept a range of linear and cyclic substrates, suggesting that DmmA will contribute to the expanding industrial applications of HLDs.  相似文献   

10.
Structural studies of caspase-1 reveal that the dimeric thiol protease can exist in two states: in an on-state, when the active site is occupied, or in an off-state, when the active site is empty or when the enzyme is bound by a synthetic allosteric ligand at the dimer interface ∼ 15 Å from the active site. A network of 21 hydrogen bonds from nine side chains connecting the active and allosteric sites change partners when going between the on-state and the off-state. Alanine-scanning mutagenesis of these nine side chains shows that only two of them—Arg286 and Glu390, which form a salt bridge—have major effects, causing 100- to 200-fold reductions in catalytic efficiency (kcat/Km). Two neighbors, Ser332 and Ser339, have minor effects, causing 4- to 7-fold reductions. A more detailed mutational analysis reveals that the enzyme is especially sensitive to substitutions of the salt bridge: even a homologous R286K substitution causes a 150-fold reduction in kcat/Km. X-ray crystal structures of these variants suggest the importance of both the salt bridge interaction and the coordination of solvent water molecules near the allosteric binding pocket. Thus, only a small subset of side chains from the larger hydrogen bonding network is critical for activity. These form a contiguous set of interactions that run from one active site through the allosteric site at the dimer interface and onto the second active site. This subset constitutes a functional allosteric circuit or “hot wire” that promotes site-to-site coupling.  相似文献   

11.
Cytochrome P450 27C1 (P450 27C1) is a retinoid desaturase expressed in the skin that catalyzes the formation of 3,4-dehydroretinoids from all-trans retinoids. Within the skin, retinoids are important regulators of proliferation and differentiation. In vivo, retinoids are bound to cellular retinol-binding proteins (CRBPs) and cellular retinoic acid–binding proteins (CRABPs). Interaction with these binding proteins is a defining characteristic of physiologically relevant enzymes in retinoid metabolism. Previous studies that characterized the catalytic activity of human P450 27C1 utilized a reconstituted in vitro system with free retinoids. However, it was unknown whether P450 27C1 could directly interact with holo-retinoid-binding proteins to receive all-trans retinoid substrates. To assess this, steady-state kinetic assays were conducted with free all-trans retinoids and holo-CRBP-1, holo-CRABP-1, and holo-CRABP-2. For holo-CRBP-1 and holo-CRABP-2, the kcat/Km values either decreased 5-fold or were equal to the respective free retinoid values. The kcat/Km value for holo-CRABP-1, however, decreased ∼65-fold in comparison with reactions with free all-trans retinoic acid. These results suggest that P450 27C1 directly accepts all-trans retinol and retinaldehyde from CRBP-1 and all-trans retinoic acid from CRABP-2, but not from CRABP-1. A difference in substrate channeling between CRABP-1 and CRABP-2 was also supported by isotope dilution experiments. Analysis of retinoid transfer from holo-CRABPs to P450 27C1 suggests that the decrease in kcat observed in steady-state kinetic assays is due to retinoid transfer becoming rate-limiting in the P450 27C1 catalytic cycle. Overall, these results illustrate that, like the CYP26 enzymes involved in retinoic acid metabolism, P450 27C1 interacts with cellular retinoid-binding proteins.  相似文献   

12.
There is much confusion and error in published treatments of data for multiple binding of ligands (e.g., substrates) by proteins (e.g., enzymes). There is a widespread impression that if the equilibrium binding, r, of ligand, A, by a protein with n sites can be fitted to an equation with two hyperbolic terms, i.e., r=nαkα(A)1+kα(A)+nβkβ(A)1+kβ(A) (nα+nβ=n) then kβ and kβ are always the intrinsic binding constants for two sets of sites. Such a conclusion is often incorrect. For example, in many cases, the protein is constituted of identical protomers with initially identical sites for binding ligands, and yet graphical representations of the binding data appear to behave as if the sites are partitioned between two classes. Although the use of a linear combination of hyperbolic terms to represent binding of ligands by macromolecules always yields empirical parameters kα, kβkλ, they cannot correspond to site binding constants when there are interactions between sites. In some circumstances their values may even be imaginary, complex numbers. On the other hand, stoichiometric binding constants can be assigned unambiguously without making any assumption regarding the nature of the interactions among binding sites. These principles are illustrated concretely by analyses of binding measurements for several different proteins containing two to six sites.  相似文献   

13.
Ricin inhibits protein synthesis by depurinating the α-sarcin/ricin loop (SRL). Ricin holotoxin does not inhibit translation unless the disulfide bond between the A (RTA) and B (RTB) subunits is reduced. Ricin holotoxin did not bind ribosomes or depurinate them but could depurinate free RNA. When RTA is separated from RTB, arginine residues located at the interface are exposed to the solvent. Because this positively charged region, but not the active site, is blocked by RTB, we mutated arginine residues at or near the interface of RTB to determine if they are critical for ribosome binding. These variants were structurally similar to wild type RTA but could not bind ribosomes. Their Km values and catalytic rates (kcat) for an SRL mimic RNA were similar to those of wild type, indicating that their activity was not altered. However, they showed an up to 5-fold increase in Km and up to 38-fold decrease in kcat toward ribosomes. These results suggest that the stalk binding stimulates the catalysis of ribosome depurination by RTA. The mutated arginines have side chains behind the active site cleft, indicating that the ribosome binding surface of RTA is on the opposite side of the surface that interacts with the SRL. We propose that stalk binding stimulates the catalysis of ribosome depurination by orienting the active site of RTA toward the SRL and thereby allows docking of the target adenine into the active site. This model may apply to the translation factors that interact with the stalk.  相似文献   

14.
A comparison of the structures and kinetic properties of human 20α-hydroxysteroid dehydrogenase (AKR1C1) and its mutant enzymes (Leu308Val and Leu308Ala) indicates that Leu308 is a selectivity determinant for substrate binding. While the Leu308Val mutation improved the catalytic efficiency (kcat/Km) of AKR1C1 towards the two substrates 5α-pregnane-3α,20α-diol (PregA) and 5β-pregnan-3α-ol-20-one (PregB), the Leu308Ala mutation rendered the enzyme inactive. In the docked model of PregA the conformation of the steroid molecule was similar to that of 20α-hydroxyprogesterone in the crystal structure of the AKR1C1 complex where the steroid did not interact with the catalytic residues Tyr55 and His117. In the case of PregB the steroid interacted with the catalytic residue His117 and formed close contacts with Leu308, suggesting that the binding mechanism of 3α-hydroxysteroids in the active site of AKR1C1 is different from that of 20α-hydroxysteroids.  相似文献   

15.
The vaccinia viral protein A27 in mature viruses specifically interacts with heparan sulfate for cell surface attachment. In addition, A27 associates with the viral membrane protein A17 to anchor to the viral membrane; however, the specific interaction between A27 and A17 remains largely unclear. To uncover the active binding sites and the underlying binding mechanism, we expressed and purified the N-terminal (18–50 residues) and C-terminal (162–203 residues) fragments of A17, which are denoted A17-N and A17-C. Through surface plasmon resonance, the binding affinity of A27/A17-N (KA = 3.40 × 108 m−1) was determined to be approximately 3 orders of magnitude stronger than that of A27/A17-C (KA = 3.40 × 105 m−1), indicating that A27 prefers to interact with A17-N rather than A17-C. Despite the disordered nature of A17-N, the A27-A17 interaction is mediated by a specific and cooperative binding mechanism that includes two active binding sites, namely 32SFMPK36 (denoted as F1 binding) and 20LDKDLFTEEQ29 (F2). Further analysis showed that F1 has stronger binding affinity and is more resistant to acidic conditions than is F2. Furthermore, A27 mutant proteins that retained partial activity to interact with the F1 and F2 sites of the A17 protein were packaged into mature virus particles at a reduced level, demonstrating that the F1/F2 interaction plays a critical role in vivo. Using these results in combination with site-directed mutagenesis data, we established a computer model to explain the specific A27-A17 binding mechanism.  相似文献   

16.
Abstract

Saturation experiments were performed on intact human peripheral mononuclear leucocytes (MNL) and MNL membranes with (-)125Iodocyanopindolol (125ICYP) over a large concentration range (1.5-600pmol/l). The corresponding Scatchard plots were curvilinear suggesting two saturable classes of binding sites: A high affinity binding site (Bmax1=1000±400 sites/cell, Kd1= 2.1±0.9 pmol/l for intact MNL and Bmax1=550±190 sites/cell, Kd1=4.1±0.9 pmol/l for MNL membranes)and a low affinity binding site (Bmax2=9150±3590 binding sites/cell, Kd2=440±50 pmol/l for intact MNL and Bmax2=11560±4690 sites/cell, Kd2=410±70 pmol/l for MNL membranes). Dissociation of (-)125ICYP from MNL was biphasic consisting of a slow dissociating component (dissociation rate constant k-1=(0.5±0.2)x10?3 min?1 for intact MNL and k-1=(1.0±0.1)x10?3min?1 for MNL membranes) and a fast dissociating component (k-2=(80±20)x10?3min?1 for intact MNL and k-2=(60±10)x10?3min?1 for MNL membranes). In dissociation experiments started after equilibration with various (-)125ICYP concentrations k-1 and k-2 were independent of the equilibrium concentration, whereas the percentual occupancy of the slow and the fast dissociating component varied and was similar to the estimated fractional occupancy of either binding site at the same (-)125ICYP concentrations in saturation experiments. The association rate constant was in the same order of magnitude for both binding sites. These results suggest two independent classes of binding sites for (-)125ICYP on MNL.  相似文献   

17.
High-mobility-group proteins HMG-1 and HMG-I/Y bind to multiple sites within a 268 bp A/T-rich enhancer element of the pea plastocyanin gene (PetE). Within a 31 bp region of the enhancer, the binding site for HMG-1 overlaps with the binding site for HMG-I/Y. The kinetics of binding and the affinities of HMG-1 and HMG-I/Y for the 31 bp DNA were determined using surface plasmon resonance. Due to very high non-specific interactions of the HMG proteins with a carboxymethyl–dextran matrix, a novel method using a cholesterol tag to anchor the DNA in a supported lipid monolayer on a thin gold film was devised. The phosphatidylcholine monolayer produced a surface that reduced background interactions to a minimum and permitted the measurement of highly reproducible protein–DNA interactions. The association rate constant (ka) of HMG-I/Y with the 31 bp DNA was ~5-fold higher than the rate constant for HMG-1, whereas the dissociation constant (KD) for HMG-I/Y (3.1 nM) was ~7-fold lower than that for HMG-1 (20.1 nM). This suggests that HMG-I/Y should bind preferentially at the overlapping binding site within this region of the PetE enhancer.  相似文献   

18.
Quaterpyridyneiron (III) complex ions anchored to partially ordered poly (L-glutamate) or poly (D-glutamate) were used as (enantiomeric) catalysts for the H2O2-oxidation of L(+) ascorbic acid at pH 7. When the α-helical fraction of polypeptide matrices was low, the configuration dissymmetry of the active sites was unable to impart any stereoselective effect in the catalysis, i.e. k = 3.66 x 103 M?1?sec?1 (25.9°C) with both catalysts. On the contrary, by increasing the amount of α-helix in the polymeric supports the stereoselectivity increases, the second-order rate constants kFeD being definitely higher than kFeL.Implications of the role played by the conformational dissymmetry of the active sites in the stereospecificity of the process are briefly discussed.  相似文献   

19.
Closely related penta- and octaheme nitrite reductases catalyze the reduction of nitrite, nitric oxide, and hydroxylamine to ammonium and of sulfite to sulfide. NrfA pentaheme nitrite reductase plays the key role in anaerobic nitrate respiration and the protection of bacterial cells from stresses caused by nitrogen oxides and hydrogen peroxide. Octaheme nitrite reductases from bacteria of the Thioalkalivibrio genus are less studied, and their function in the cell is unknown. In order to estimate the possible role of octaheme nitrite reductases in the cell resistance to oxidative stress, the peroxidase activity of the enzyme from T. nitratireducens (TvNiR) has been studied in detail. Comparative analysis of the active site structure of TvNiR and cytochrome c peroxidases has shown some common features, such as a five-coordinated catalytic heme and identical catalytic residues in active sites. A model of the possible productive binding of peroxide at the active site of TvNiR has been proposed. The peroxidase activity has been measured for TvNiR hexamers and trimers under different conditions (pH, buffers, the addition of CaCl2 and EDTA). The maximum peroxidase activity of TvNiR with ABTS as a substrate (k cat = 17 s–1; k cat/K m = 855 mM–1 s–1) has been 100–300 times lower than the activity of natural peroxidases. The different activities of TvNiR trimers and hexamers indicate that the rate-limiting stage of the reaction is not the catalytic event at the active site but the electron transfer along the heme c electron-transport chain.  相似文献   

20.
The rate of Fe3+ release from horse spleen ferritin (HoSF) was measured using the Fe3+-specific chelator desferoxamine (DES). The reaction consists of two kinetic phases. The first is a rapid non-linear reaction followed by a slower linear reaction. The overall two-phase reaction was resolved into three kinetic events: 1) a rapid first-order reaction in HoSF (k1); 2) a second slower first-order reaction in HoSF (k2); and 3) a zero-order slow reaction in HoSF (k3). The zero-order reaction was independent of DES concentration. The two first-order reactions had a near zero-order dependence on DES concentration and were independent of pH from 6.8 to 8.2. The two first-order reactions accounted for 6-9 rapidly reacting Fe3+ ions. Activation energies of 10.5 ± 0.8, 13.5 ± 2.0 and 62.4 ± 2.1 kJ/mol were calculated for the kinetic events associated with k1, k2, and k3, respectively. Iron release occurs by: 1) a slow zero-order rate-limiting reaction governed by k3 and corresponding to the dissociation of Fe3+ ions from the FeOOH core that bind to an Fe3+ binding site designated as site 1 (proposed to be within the 3-fold channel); 2) transfer of Fe3+ from site 1 to site 2 (a second binding site in the 3-fold channel) (k2); and 3) rapid iron loss from site 2 to DES (k1).  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