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.
The binding of the monoiodinated alpha-neurotoxin I from Naja mossambica mossambica to the membrane-bound acetylcholine receptor from Torpedo marmorata was investigated using a new picomolar-sensitive microtitration assay. From equilibrium binding studies a non-linear Scatchard plot demonstrated two populations of binding sites characterized by the two dissociation constants Kd1 = 7 +/- 4 pM and Kd2 = 51 +/- 16 pM and having equal binding capacities. These two populations differed in their rate of dissociation (k-1.1 = 25 x 10(-6) s-1 and k-1.2 = 623 x 10(-6) s-1 respectively), but not in their rate of formation of the toxin-receptor complex (k + 1 = 11.7 x 10(6) M-1 s-1). From these rate constants the same two values of dissociation constant were deduced (Kd1 = 2 pM and Kd2 = 53 pM). All the specific binding was prevented by the cholinergic antagonists alpha-bungarotoxin and d-tubocurarine. In addition, a biphasic competition phenomenon allowed us to differentiate between two d-tubocurarine sites (Kda = 103 nM and Kdb = 13.7 microM respectively). Evidence is provided indicating that these two sites are shared by d-tubocurarine and alpha-neurotoxin I, with inverse affinities. Fairly conclusive agreement between our equilibrium, kinetic and competition data demonstrates that the two high-affinity binding sites for this short alpha-neurotoxin are selectively distinguishable.  相似文献   

2.
The binding of substrates and the herbicide N-(phosphonomethyl)glycine (glyphosate) to enolpyruvoylshikimate-3-phosphate (EPSP) synthase was evaluated by stopped-flow and equilibrium fluorescence measurements. Changes in protein fluorescence were observed upon the binding of EPSP and upon the formation of the enzyme-shikimate 3-phosphate-glyphosate ternary complex; no change was seen with either shikimate 3-phosphate (S3P) or glyphosate alone. By fluorescence titrations, the dissociation constants were determined for the formation of the enzyme binary complexes with S3P (Kd,S = 7 +/- 1.2 microM) and EPSP (Kd,EPSP = 1 +/- 0.01 microM). The dissociation constant for S3P was determined by competition with EPSP or by measurements in the presence of a low glyphosate concentration. At saturating concentrations of S3P, glyphosate bound to the enzyme--S3P binary complex with a dissociation constant of 0.16 +/- 0.02 microM. Glyphosate did not bind significantly to free enzyme, so the binding is ordered with S3P binding first: (formula; see text) where S refers to S3P, G refers to glyphosate, and E.S.G. represents the complex with altered fluorescence. The kinetics of binding were measured by stopped-flow fluorescence methods. The rate of glyphosate binding to the enzyme--S3P complex was k2 = (7.8 +/- 0.2) X 10(5) M-1 s-1, from which we calculated the dissociation rate k-2 = 0.12 +/- 0.02 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
A new chromium complex, (bis(hydroxyethyl)amino-tris(hydroxymethyl)methane)oxochromate(V), has been characterized by epr spectroscopy. The chromium(V) complex was formed by the ligand displacement reaction of bis(2-ethyl-2-hydroxybutanato) oxochromate(V) with bis(hydroxyethyl)amino-tris(hydroxy-methyl)methane (BT). Both epr and kinetic data indicate that the reaction proceeds through a chromium(V) intermediate. Kinetics of formation of the intermediate exhibit a rate saturation at higher [BT] (> 30 mM) indicating a rate law constituting an equilibrium between the parent Cr(V) complex and the bis-tris ligand followed by a pure first order process. The g-value of the intermediate is consistent with a Cr(V) complex in which the BT is coordinated in a bidentate fashion replacing a coordinated hydroxy butanoic acid ligand, affording a mixed ligand complex. The equilibrium step (K = 36 M-1) consists of monodentate coordination by the BT ligand and the limiting first order rate constant (1.9 x 10(-2) s-1) manifests the rate of chelation by the polydentate ligand. The intermediate is converted to the product upon further chelation through the complete displacement of the remaining 2-ethyl-2-hydroxy butanoic acid by a first order process (k = 0.023 s-1). The epr data support a pair of products that are in rapid equilibrium. In these products, BT functions either as a tetra or a penta-dentate ligand coordinating through four or five alkoxy sites. The enthalpy and entropy of activations related to the two chelation steps were found to be 32 +/- 2 kJ/mol and -(1.7 +/- 0.2) x 10(2) J/mol K for the intermediate, and 36 +/- 1 kJ/mol and -(1.5 +/- 0.2) x 10(2) J/mol K for the product. Our data support an associative mechanism for the chelation steps. The Cr(V)-BT product is more stable than the parent complex. The second order disproportionation rate constant for the Cr(V)-BT complex was evaluated to be 0.1 M-1 s-1 compared to 8.0 M-1 s-1 for the parent complex. This is the first example of a chromium(V) complex with a non-macrocyclic ligand coordinating through oxygen donor atoms which is stable in aqueous solution at neutral pH over a long period of time.  相似文献   

4.
Membrane-mediated assembly of the prothrombinase complex   总被引:1,自引:0,他引:1  
Prothrombinase assembly was studied on macroscopic planar bilayers consisting of 20% dioleoyl-phosphatidylserine (DOPS) and 80% dioleoyl-phosphatidylcholine (DOPC). The dissociation constant for the binding of factor Xa to the bilayer, measured by ellipsometry, was Kd = 47 +/- 8 nM (mean +/- S.D.) and this value was lowered to Kd = 2.2 +/- 0.3 pM by preadsorption of factor Va. This latter value was determined from direct measurement of steady-state thrombin production. A comparable value of Kd = 1.0 +/- 0.1 pM was found by repeating these experiments in suspensions of phospholipid vesicles, and it was verified that prothrombinase assembly was not influenced by the addition of prothrombin. Using a minute amount (0.094 fmol cm-2) of preadsorbed factor Va, it was found that the rate of prothrombinase assembly exceeds the rate of collisions between Xa molecules from the buffer and the sparse Va molecules on the bilayer. Apparently, factor Xa adsorbs first to the membrane and then associates rapidly with factor Va by lateral diffusion. The data indicate almost instantaneous equilibrium of this complex formation on the surface with a lower limit for the bimolecular rate constant of kon = 2.8 x 10(13) (mol/cm2)-1 s-1. In suspensions of small phospholipid vesicles, prothrombinase assembly is collisionally limited and the value of kon should be proportional to vesicle diameter. This was verified with a method for estimation of kon values from thrombin generation curves. Values of 0.36 x 10(9) and 1.6 x 10(9) M-1 s-1 were found for vesicles of 20-30- and 60-80-nm diameter, respectively.  相似文献   

5.
We have studied the kinetics of binding of the menaquinol analog 2-n-heptyl-4-hydroxyquinoline-N-oxide (HOQNO) by fumarate reductase (FrdABCD) using the stopped-flow method. The results show that the fluorescence of HOQNO is quenched when HOQNO binds to FrdABCD. The observed quenching of HOQNO fluorescence has two phases and it can be best fitted to a double exponential equation. A two-step equilibrium model is applied to describe the binding process in which HOQNO associates with FrdABCD by a fast bimolecular step to form a loosely bound complex; this is subsequently converted into a tightly bound complex by a slow unimolecular step. The rates of the forward and the reverse reactions for the first equilibrium (k1 and k2) are determined to be k1 = (1.1 +/- 0.1) x 10(7) M-1.s-1, and k2 = 6.0 +/- 0.6 s-1, respectively. The dissociation constants of the first equilibrium (Kd1 = k2/k1) is calculated to be about 550 nM. The overall dissociation constant for the two-step equilibrium, Kd overall = Kd1/[1+ (1/Kd2)], is estimated to be < or = 7 nM. Comparison of the kinetic parameters of HOQNO binding by FrdABCD and by dimethyl sulfoxide reductase provides important information on menaquinol binding by these two enzymes.  相似文献   

6.
Temperature jump relaxation kinetics of the P-450cam spin equilibrium   总被引:1,自引:0,他引:1  
M T Fisher  S G Sligar 《Biochemistry》1987,26(15):4797-4803
The ferric spin-state equilibrium and relaxation rate of cytochrome P-450 has been examined with temperature jump spectroscopy using a number of camphor analogues known to induce different mixed spin states in the substrate-bound complexes [Gould, P., Gelb, M., & Sligar, S. G. (1981) J. Biol. Chem. 256, 6686]. All temperature-induced spectral changes were monophasic, and the spin-state relaxation rate reached a limiting value at high substrate concentrations. The ferric spin equilibrium constant, Kspin, is defined in terms of the rate constants k1 and k-1 via Kspin = k1/k-1 = [P-450(HS)]/[P-450(LS)] where HS and LS represent high-spin (S = 5/2) and low-spin (S = 1/2) ferric iron, respectively, and the spectrally observed spin-state relaxation rate by kobsd = k1 + k-1. A strong correlation between the fraction of high-spin species and the rate constant, k-1, is observed. For a 3 degrees C temperature jump (from 10 to 13 degrees C), the 23% high-spin tetramethylcyclohexanone complex (Kd = 45 +/- 20 microM) is characterized by a ferric spin relaxation rate of kobsd = 1990 s-1, while the rates for the d-fenchone (41% high spin, Kd = 42 +/- 10 microM) and kobsd = 1990 s-1, while the rates for the d-fenchone (41% high spin, Kd = 42 +/- 10 microM) and camphoroquinone (75% high spin, Kd = 15 +/- 5 microM) complexes are 1430 and 346 s-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

8.
Kinetics and mechanism of bilirubin binding to human serum albumin   总被引:3,自引:0,他引:3  
The kinetics of bilirubin binding to human serum albumin at pH 7.40, 4 degrees C, was studied by monitoring changes in bilirubin absorbance. The time course of the absorbance change at 380 nm was complex: at least three kinetic events were detected including the bimolecular association (k1 = 3.8 +/- 2.0 X 10(7) M-1 S-1) and two relaxation steps (52 = 40.2 +/- 9.4 s-1 and k3 = 3.8 +/- 0.5 s-1). The presence of the two slow relaxations was confirmed under pseudo-first order conditions with excess albumin. Curve-fitting procedures allowed the assignment of absorption coefficients to the intermediate species. When the bilirubin-albumin binding kinetics was observed at 420 nm, only the two relaxations were seen; apparently the second order association step was isosbestic at this wavelength. The rate of albumin-bound bilirubin dissociation was measured by mixing the pre-equilibrated human albumin-bilirubin complex with bovine albumin. The rate constant for bilirubin dissociation measured at 485 nm was k-3 = 0.01 s-1 at 4 degrees C. A minimum value of the equilibrium constant for bilirubin binding to human albumin determined from the ratio k1/k-3 is therefore approximately 4 X 10(9) M-1.  相似文献   

9.
The diffusion coefficient (D) of 2,3-bisphosphoglycerate (DPG) was measured using pulsed-field gradient (PFG)-31P nuclear magnetic resonance spectroscopy in solutions containing 2.7-5.0 mM hemoglobin (Hb) and a range of DPG concentrations. The dependence of the measured values of D on the fraction of the total DPG in the sample that is bound to Hb enabled the estimation of the dissociation constants (Kd) of complexes of DPG with carbonmonoxygenated, oxygenated, and deoxygenated Hb; the values of Kd (mM), measured at 25 degrees C, pH 6.9 and in 100 mM bis Tris/50 mM KCl, were 1.98 +/- 0.26, 1.8 +/- 0.5 and 0.39 +/- 0.26, respectively. In intact erythrocytes the apparent diffusion coefficient, Dapp, of DPG was larger in oxygenated and carbonmonoxygenated cells (6.17 +/- 0.20 x 10(-11) m2s-1) than in deoxygenated cells (4.10 +/- 0.23 x 10(-11) m2s-1). Changes in intracellular DPG concentration (5-55 mM) in erythrocytes, brought about by incubation in a medium containing inosine and pyruvate, did not result in significant changes in the value of Dapp; this result supports the hypothesis that DPG binds to other sites in the erythrocyte. Monte Carlo simulations of diffusion in biconcave discs were used to test the adequacy of the values of Kd estimated in solution to describe the binding of DPG to Hb in oxygenated and deoxygenated erythrocytes. The results of the simulations implied that the value of Kd estimated for deoxygenated Hb-DPG was greater than expected from the experiments involving intact erythrocytes. This difference is surmised to be at least partly due to the difficulty of measuring D at low-ligand concentrations. Notwithstanding this shortcoming, the PFG method appears to be suitable for probing interactions between macromolecules and ligands when the Kd is in the millimolar range. It is one of the few techniques available in which these interactions can be studied in intact cells. In addition, the Monte Carlo simulations of the diffusion experiments highlighted important differences between theory and experiment relating to the nature of molecular motion inside the cells.  相似文献   

10.
Cooperativity with glucose is a key feature of human glucokinase (GK), allowing its crucial role as a glucose sensor in hepatic and pancreatic cells. We studied the changes in enzyme intrinsic tryptophan fluorescence induced by binding of different ligands to this monomeric enzyme using stopped-flow and equilibrium binding methods. Glucose binding data under pre-steady state conditions suggest that the free enzyme in solution is in a preexisting equilibrium between at least two conformers (super-open and open) which differ in their affinity for glucose (Kd* = 0.17 +/- 0.02 mM and Kd = 73 +/- 18 mM). Increasing the glucose concentration changes the ratio of the two conformers, thus yielding an apparent Kd of 3 mM (different from a Km of 7-10 mM). The rates of conformational transitions of free and GK complexed with sugar are slow and during catalysis are most likely affected by ATP binding, phosphate transfer, and product release steps to allow the kcat to be 60 s-1. The ATP analogue PNP-AMP binds to free GK (super-open) and GK-glucose (open) complexes with comparable affinities (Kd = 0.23 +/- 0.02 and 0.19 +/- 0.08 mM, respectively). However, cooperativity with PNP-AMP observed under equilibrium binding conditions in the presence of glucose (Hill slope of 1.6) is indicative of further complex tightening to the closed conformation. Another physiological modulator (inhibitor), palmitoyl-CoA, binds to GK with similar characteristics, suggesting that conformational changes induced upon ligand binding are not restricted by an active site ligand. In conclusion, our data support control of GK activity and Km through the ratio of distinct conformers (super-open, open, and closed) through either substrate or other ligand binding and/or dissociation.  相似文献   

11.
p-Hydroxyphenylacetate hydroxylase (HPAH) from Acinetobacter baumannii catalyzes the hydroxylation of p-hydroxyphenylacetate (HPA) to form 3,4-dihydroxyphenylacetate (DHPA). The enzyme system is composed of two proteins: an FMN reductase (C1) and an oxygenase that uses FMNH- (C2). We report detailed transient kinetics studies at 4 degrees C of the reaction mechanism of C2.C2 binds rapidly and tightly to reduced FMN (Kd, 1.2 +/- 0.2 microm), but less tightly to oxidized FMN (Kd, 250 +/- 50 microm). The complex of C -FMNH-2 reacted with oxygen to form C(4a)-hydroperoxy-FMN at 1.1 +/- 0.1 x 10(6) m(-1) s(-1), whereas the C -FMNH-2 -HPA complex reacted with oxygen to form C(4a)-hydroperoxy-FMN-HPA more slowly (k = 4.8 +/- 0.2 x 10(4) m(-1) s(-1)). The kinetic mechanism of C2 was shown to be a preferential random order type, in which HPA or oxygen can initially bind to the C -FMNH-2 complex, but the preferred path was oxygen reacting with C -FMNH-2 to form the C(4a)-hydroperoxy-FMN intermediate prior to HPA binding. Hydroxylation occurs from the ternary complex with a rate constant of 20 s(-1) to form the C2-C(4a)-hydroxy-FMN-DHPA complex. At high HPA concentrations (>0.5 mm), HPA formed a dead end complex with the C2-C(4a)-hydroxy-FMN intermediate (similar to single component flavoprotein hydroxylases), thus inhibiting the bound flavin from returning to the oxidized form. When FADH- was used, C(4a)-hydroperoxy-FAD, C(4a)-hydroxy-FAD, and product were formed at rates similar to those with FMNH-. Thus, C2 has the unusual ability to use both common flavin cofactors in catalysis.  相似文献   

12.
Lee SK  Lipscomb JD 《Biochemistry》1999,38(14):4423-4432
The effects of solvent pH and deuteration on the transient kinetics of the key intermediates of the dioxygen activation process catalyzed by the soluble form of methane monooxygenase (MMO) isolated from Methylosinus trichosporium OB3b have been studied. MMO consists of hydroxylase (MMOH), reductase, and "B" (MMOB) components. MMOH contains a carboxylate- and oxygen-bridged binuclear iron cluster that catalyzes O2 activation and insertion chemistry. The diferrous MMOH-MMOB complex reacts with O2 to form a diferrous intermediate compound O (O) and subsequently a diferric intermediate compound P (P), presumed to be a peroxy adduct. The O decay reaction was found to be pH-independent within error at 4 degrees C (kobs = 22 +/- 2 s-1 at pH 7.7; kobs = 26 +/- 2 s-1 at pH 7.0). In contrast, the P formation rate was found to decrease sharply with increasing pH to near zero at pH 8.6; the observed rate constants fit to a single deprotonation event with a pKa = 7.6 and a maximal formation rate at 4 degrees C of kP = 9.1 +/- 0.9 s-1 achieved near pH 6.5. The formation of P was slower than the disappearance of O, indicating that at least one other undetected intermediate (P) must form in between. P decays spontaneously to the highly chromophoric intermediate, compound Q (Q). The decay rate of P matched the formation rate of Q, and both rates decreased sharply with increasing pH to near zero at pH 8.6; the observed rate constants fit to a single deprotonation event with a pKa = 7.6 and a maximal formation rate at 4 degrees C of kQ = 2.6 +/- 0.1 s-1 achieved near pH 6.5. No pH dependence was observed for the decay of Q. The formation and decay rates of P and the formation rate of Q decreased linearly with mole fraction of D2O in the reaction mixture. Kinetic solvent isotope effect values of kH/kD = 1.3 +/- 0.1 (P formation) and kH/kD = 1.4 +/- 0.1 (P decay and Q formation) were observed at 5 degrees C. The linearity of the proton inventory plots suggests that only a single proton is transferred in the transition state of the formation reaction for each intermediate. If these protons are transferred to the bound oxygen molecule, as formally required by the reaction stoichiometry, the data are consistent with a model in which water is formed concurrently with the formation of the reactive bis mu-oxo-binuclear Fe(IV) species, Q.  相似文献   

13.
The reactions of Fe3+-EDTA and Fe2+-EDTA with O2- and CO2- were investigated in the pH range 3.8--11.8. Around neutral pH O2- reduces Fe3+-EDTA with a rate constant which is pH dependent kpH 5.8--8.1 = 2 - 10(6)--5 - 10(5) M-1 - s-1. At higher pH values this reaction becomes much slower. The CO2- radical reduces Fe3+-EDTA with kpH 3.8--1- = 5 +/- 1 - 10(7) M-1 - s-1 independent of pH. At pH 9--11.8, Fe2+-EDTA forms a complex with O2- with kFe2+-EDTA + O2 = 2 - 10(6)--4 - 10(6) M-1 - s-1 which is pH dependent. We measured the spectrum of Fe2+-EDTA-O2- and calculated epsilon 290 over max = 6400 +/- 800 M-1 - cm-1 in air-saturated solutions. In O2-saturated solutions another species is formed with a rate constant of 7 +/- 2 s-1. This intermediate absorbs around 300 nm but we were not able to identify it.  相似文献   

14.
W P Lu  Y Sun  M D Bauer  S Paule  P M Koenigs  W G Kraft 《Biochemistry》1999,38(20):6537-6546
Penicillin-binding protein 2a (PBP2a) is the primary beta-lactam resistance determinant of methicillin-resistant Staphylococcus aureus (MRSA). MecA, the gene coding for PBP2a, was cloned with the membrane-anchoring region at the N-terminus deleted. The truncated protein (PBP2a) was overexpressed in Escherichia coli mostly in the soluble form accounting for approximately 25% of soluble cell protein and was purified to homogeneity. The purified protein was shown to covalently bind beta-lactams in an 1:1 ratio as determined by electrospray mass spectrometry. A novel method based on HPLC-elctrospray mass spectrometry has been developed to quantitatively determine the formation of the covalent adducts or acyl-PBP2a complexes. By using this method, combined with kinetic techniques including quench flow, we have extensively characterized the interactions between PBP2a and three beta-lactams and determined related kinetic parameters for the first time. The apparent first-order rate constants (ka) of PBP2a acylation by benzylpenicillin showed a hyperbolic dependence on the concentration of benzylpenicillin. This is consistent with the mechanism that the binding of the penicillin to PBP2a consists of reversible formation of a Michaelis complex followed by formation of the penicilloyl-PBP2a adduct, and allowed the determination of the individual kinetic parameters for these two steps, the dissociation constant Kd of 13.3 mM and the first-order rate constant k2 of 0.22 s-1. From these values, the second-order rate constant k2/Kd, the value reflecting the overall binding efficiency of a beta-lactam, of 16.5 M-1 s-1 was obtained. The fairly high Kd value indicates that benzylpenicillin fits rather poorly into the protein active site. Similar studies on the interaction between PBP2a and methicillin revealed k2 of 0.0083 s-1 and Kd of 16.9 mM, resulting in an even smaller k2/Kd value of 0.49 M-1 s-1. The rate constants k3 for deacylation of the acyl-PBP2a complexes, the third step in the interactions, were measured to be <1.5 x 10(-)5 s-1. These results indicate that the resistance of PBP2a to penicillin inactivation is mainly due to the extremely low penicillin acylating rate in addition to the low association affinity, but not to a fast rate of deacylation. Acylation of PBP2a by a high-affinity cephalosporin, Compound 1, also followed a saturation curve of ka versus the compound concentration, from which k2 = 0.39 s-1, Kd = 0.22 mM, and k2/Kd = 1750 M-1 s-1 were obtained. The 100-fold increase in the k2/Kd value as compared with that of benzylpenicillin is mostly attributable to the decreased (60-fold) Kd, indicating that the cephalosporin fits much better to the binding pocket of the protein.  相似文献   

15.
Human Nedd8-activating enzyme AppBp1-Uba3 was purified to apparent homogeneity from erythrocytes. In the presence of [2,8-3H]ATP and 125I-Nedd8, heterodimer rapidly forms a stable stoichiometric ternary complex composed of tightly bound Nedd8 [3H]adenylate and Uba3-125I-Nedd8 thiol ester. Isotope exchange kinetics show that the heterodimer follows a pseudo-ordered mechanism with ATP the leading and Nedd8 the trailing substrate. Human AppBp1-Uba3 follows hyperbolic kinetics for HsUbc12 transthiolation with 125I-Nedd8 (kcat = 3.5 +/- 0.2 s-1), yielding Km values for ATP (103 +/- 12 microm), 125I-Nedd8 (0.95 +/- 0.18 microm), and HsUbc12 (43 +/- 13 nm) similar to those for ubiquitin activation by Uba1. Wild type 125I-ubiquitin fails to support AppBp1-Uba3 catalyzed activation or HsUbc12 transthiolation. However, modest inhibition of 125I-Nedd8 ternary complex formation by unlabeled ubiquitin suggests a Kd > 300 microm for ubiquitin. Alanine 72 of Nedd8 is a critical specificity determinant for AppBp1-Uba3 binding because 125I-UbR72L undergoes heterodimer-catalyzed hyperbolic HsUbc12 transthiolation and yields Km = 20 +/- 9 microm and kcat = 0.9 +/- 0.3 s-1. These observations demonstrate remarkable conservation in the mechanism of AppBp1-Uba3 that mirrors its sequence conservation with the Uba1 ubiquitin-activating enzyme.  相似文献   

16.
Sphingosine inhibited [3H]methylhistidine-thyrotropin-releasing hormone (MeTRH) binding to intact GH3 cells and to GH3 membranes. This inhibition was dependent on the concentration of sphingosine and on the ratio of sphingosine to cell number (or membrane protein) and was partly reversed by washing. In intact cells, the IC50 was 63 microM (1.8 X 10(6) cells/ml; 2 nM MeTRH), and 100 microM sphingosine was found, by Scatchard analysis, to increase the apparent dissociation constant (Kd) from 1.1 +/- 0.3 to 6.5 +/- 2.3 nM and to decrease the maximal binding capacity (Bmax) to 41 +/- 9.5% of control. Kinetic analysis showed that the major effect of sphingosine on Kd was due to a marked decrease in the apparent association rate constant for MeTRH from 2.5 +/- 0.4 X 10(5) M-1 s-1 to 0.10 +/- 0.015 X 10(5) M-1 s-1. At 100 microM, sterylamine was as effective as sphingosine in inhibiting MeTRH binding, whereas sphinganine was less effective, and psychosine and steroylsphingosine were without effect. The following observations show that sphingosine inhibition of MeTRH binding did not involve protein kinase C. The IC50 for sphingosine inhibition of MeTRH binding was the same in GH3 cells that had been incubated with 1 microM phorbol 12-myristate 13-acetate for 16 h, to "down-regulate" protein kinase C, as in control cells. Sphingosine inhibited MeTRH binding to membranes isolated from GH3 cells that contain very little protein kinase C activity. In GH3 membranes, 100 microM sphingosine increased the Kd for MeTRH from 3.4 +/- 0.1 to 13 +/- 3.1 nM but did not significantly decrease Bmax (12 +/- 5.0% of control, p greater than 0.05). And, 1-(5-isoquinolinesulfonyl)-2-methylpiperazine dihydrochloride, an inhibitor of protein kinase C, failed to decrease MeTRH binding to intact GH3 cells or to membranes, and did not interfere with the effects of sphingosine. These data show that sphingosine and its analogs have complex actions to inhibit MeTRH binding to GH3 cells, at least some of which are independent of protein kinase C, and thereby demonstrate that sphingolipids cannot be used as specific inhibitors of protein kinase C.  相似文献   

17.
Leupeptin (acyl peptidyl-L-argininal) is a potent inhibitor of trypsin and related proteases. We analyzed the association of leupeptim with bovine trypsin kinetically, assuming that it proceeds by a pathway which involves two steps: E + I in equilibrium K1 Complex I k-2 in equilibrium k+2 Complex II. The observed dissociation constant (K1) for the first step was 1.24 X 10(-3) M (at pH 8.2 15 degrees C) and the two first-order rate constants (k+2 and k-2) were 166 s-1 and 1.75 X 10(-3.s-1, respectively (at pH 8.2, 15 degrees C). The dissociation constant (Kd) for the whole process was calculated from these parameters to be 1.34 X 10(-8) M. This value is compatible with that determined directly by an independent static method (2.36 X 10(-8) M). We also measured Kd for the leupeptine complex of anhydrotrypsin, a trypsin derivative in which the active-site hydroxyl group is missing. The observed value was about 5 orders of magnitude larger than Kd and was rather similar to K1 in native trypsin. A elupeptin isomer which contains a D-argininal residue did not show strong affinity towards trypsin. These findings suggest that complex II consists of a covalent hemiacetal adduct formed between the serine hydroxyl group in the enzyme active site and the aldehyde group in the inhibitor. The pH dependencies of the dissociation constant and other parameters show that deprotonation of the charge-relay sustem in the active site is important for the formation and stabilization of complex II.  相似文献   

18.
The rate-limiting step of ethanol oxidation by alcohol dehydrogenase (E) at substrate inhibitory conditions (greater than 500 mM ethanol) is shown to be the dissociation rate of NADH from the abortive E-ethanol-NADH complex. The dissociation rate constant of NADH decreased hyperbolically from 5.2 to 1.4 s-1 in the presence of ethanol causing a decrease in the Kd of NADH binding from 0.3 microM for the binary complex to 0.1 microM for the abortive complex. Correspondingly, ethanol binding to E-NADH (Kd = 37 mM) was tighter than to enzyme (Kd = 109 mM). The binding rate of NAD+ (7 X 10(5) M-1s-1) to enzyme was not affected by the presence of ethanol, further substantiating that substrate inhibition is totally due to a decrease in the dissociation rate constant of NADH from the abortive complex. Substrate inhibition was also observed with the coenzyme analog, APAD+, but a single transient was not found to be rate limiting. Nevertheless, the presence of substrate inhibition with APAD+ is ascribed to a decrease in the dissociation rate of APADH from 120 to 22 s-1 for the abortive complex. Studies to discern the additional limiting transient(s) in turnover with APAD+ and NAD+ were unsuccessful but showed that any isomerization of the enzyme-reduced coenzyme-aldehyde complex is not rate limiting. Chloride increases the rate of ethanol oxidation by hyperbolically increasing the dissociation rate constant of NADH from enzyme and the abortive complex to 12 and 2.8 s-1, respectively. The chloride effect is attributed to the binding of chloride to these complexes, destabilizing the binding of NADH while not affecting the binding of ethanol.  相似文献   

19.
Kim S  Cullis DN  Feig LA  Baleja JD 《Biochemistry》2001,40(23):6776-6785
The recently described EH domain recognizes proteins containing Asn-Pro-Phe (NPF) sequences. Using nuclear magnetic resonance (NMR) data, we determined the solution structure of the EH domain from the Reps1 protein and characterized its binding to linear and cyclic peptides derived from a novel targeting protein. The structure calculation included 1143 distance restraints and 122 angle restraints and resulted in structures with a root-mean-square deviation of 0.40 +/- 0.05 A for backbone atoms of superimposed secondary structural elements. The structure comprises two helix-loop-helix motifs characteristic of EF-hand domains. Titration data with NPF-containing peptides showed evidence of intermediate exchange on the NMR chemical shift time scale, which required an analysis that includes curve fitting to obtain accurate equilibrium constants and dissociation rate constants. The cyclic and linear peptides bound with similar affinities (Kd = 65 +/- 17 and 46 +/- 14 microM, respectively) and to the same hydrophobic pocket formed between helices B and C. The cyclic peptide formed a complex that dissociated more slowly (k(off) = 440 +/- 110 s(-1)) than the linear peptide (k(off) = 1800 +/- 250 s(-1)), but had little change in affinity because of the slower rate of association of the cyclic peptide. In addition, we characterized binding to a peptide containing a DPF sequence (Kd = 0.5 +/- 0.2 mM). The characterization of binding between the Reps1 EH domain and its target proteins provides information about their role in endocytosis.  相似文献   

20.
N R Woodruff  K E Neet 《Biochemistry》1986,25(24):7956-7966
The association kinetics of 125I beta nerve growth factor (NGF) binding to the PC12 clonal cell line have been examined in detail at 0.5 and 37 degrees C. These data were examined by utilizing a reversible second-order integrated rate equation, and the results were not consistent with a simple bimolecular process. Two association rates were required to explain the results adequately. At 37 degrees C, the faster component was estimated to have a second-order association rate constant of 1.4 X 10(7) M-1 s-1, while the rate constant for the slower component (3.8 X 10(6) M-1 s-1) was about 4-fold lower. As shown by others, the temperature dependence of the dissociation kinetics indicated that while the rapidly dissociating component was only slightly slowed by lowering the chase temperature to 0.5 degrees C, the second component was slowed by about 270-fold, from 8 X 10(-4) s-1 to 3 X 10(-6) s-1. The binding data that describe the slowly dissociating component were obtained by utilizing this differential temperature dependence and revealed a concave downward Scatchard plot. The binding parameters determined from computer analysis using a nonlinear fitting program (LIGAND) suggest that this component consists of (a) an interacting class of about 4000 sites/cell that have a first stoichiometric steady-state dissociation constant of 65 pM and a second stoichiometric interaction constant of 16 pM, indicative of positively cooperative interactions, and (b) a class of sites consistent with a ratio of sites/Kd of about 11.1 sites/(cell X pM). The steady-state binding results at 37 degrees C indicated only one class of binding sites (155,000 +/- 18,000 sites/cell) that had an apparent Kd of 0.52 +/- 0.03 nM. One class of sites was also observed at 0.5 degrees C, and the receptor concentration was found to be reduced (99,000 +/- 7600 sites/cell) while the Kd was increased (1.7 +/- 0.14 nM). A significant level of positively cooperative interactions was observed frequently at 37 degrees C that was not due to a failure to reach steady-state conditions, internalization, or degradation. Since cooperativity of binding was never observed at 0.5 degrees C, a membrane event may be involved. Determination of the contribution of the different classes of NGF receptors found on PC12 cells to the biological actions of NGF requires a clear understanding of their kinetic properties and their relationship to each other. The studies presented here indicate that their interactions are more complex than previously described.  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