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 effects of neomycin, fluoride and the non-hydrolysable guanine nucleotide analogue GTP gamma S on the kinetics of cell-free activation of NADPH oxidase in membranes of resting human neutrophils were investigated. Arachidonate-mediated activation of the oxidase followed a first-order reaction course (kobs. = 0.39 min-1 at 26 degrees C). In the presence of NaF during the activation process, activity was enhanced while the activation rate was slightly reduced (kobs. = 0.25 min-1 at 26 degrees C). Neomycin blocked activation (half-maximal effect at 25 microM) without affecting rates of superoxide release by preactivated enzyme in vitro or in vivo. In spite of reduced specific activity neither the first-order rate constant of the activation nor the Km of the oxidase were altered by neomycin. Oxidase activated in the presence of GTP gamma S exhibited increased specific activity and unchanged Km; the course of the reaction deviated from first-order kinetics. Kinetic evidence is presented for two separate activation reactions: a GTP gamma S-independent, basal, first-order process and a GTP gamma S-dependent sigmoid activation process. The results are compatible with the existence in neutrophil membranes of two separate pools of dormant oxidase. An alternative scheme of the formation of two active forms of NADPH oxidase is also presented.  相似文献   

2.
Conformational changes between E1 and E2 enzyme forms of a dog kidney Na+/K(+)-ATPase preparation labeled with 5-iodoacetamidofluorescein were followed with a stopped-flow fluorimeter, in terms of the rate constant, kobs, and the steady-state magnitude, % delta F of fluorescence change. On rapid mixing of enzyme plus Mg2+ plus Na+ with saturating (0.5 mM) ATP in the absence of K+, kobs varied with Na+ concentration in the range 0-155 mM, with a K1/2 of 10 mM, while % delta F was relatively insensitive to Na+, with a K1/2 of 0.5 mM. Oligomycin reduced kobs by 98-99% for Na+ greater than or equal to 10 mM, but only by 50% for Na+ = 1 mM; % delta F was reduced at most by 20%. At 155 mM Na+, both kobs and % delta F changed if K+ was present with the enzyme. kobs decreased by 50% when K+ was increased from 0 to 0.2 mM, but increased when K+ was varied in the range 0.2-5 mM. K+ increased % delta F by a factor of 3 with a K1/2 of 0.3-0.5 mM as measured in both stopped-flow and steady-state experiments. These data are considered in terms of the derived presteady-state equations for two alternate schemes for the enzyme, with the E1P to E2P conformational change either preceding (Albers-Post) or following (N?rby-Yoda-Skou) Na+ transport and release. The analysis indicates that: (i) Na+ must be released before the conformational transition, from an E1 form; (ii) the step in which the second and/or third Na+ is released is rate-limiting, but this release is accelerated by Na+; and (iii) the release is also accelerated by K+ acting with low affinity (possibly at extracellular sites).  相似文献   

3.
Calcium dissociation from the C-terminal and N-terminal halves of calmodulin, intact bovine brain calmodulin and the respective phenoxybenzamine complexes or melittin complexes was measured directly by stopped-flow fluorescence with the calcium chelator Quin 2 and, when possible, also by protein fluorescence using endogenous tyrosine fluorescence by mixing with EGTA. Calcium dissociation from the C-terminal half of calmodulin, which contains only the two high-affinity calcium-binding sites, and from intact calmodulin was monophasic, with good correlation of the rates of calcium dissociation obtained by the two methods. The apparent rates with Quin 2 and endogenous tyrosine fluorescence were 13.4 s-1 and 12.8 s-1, respectively, in the C-terminal half and 10.5 s-1 and 10.8 s-1, respectively, in intact calmodulin (pH 7.0, 25 degrees C, 100 mM KCl). Alkylation of the C-terminal half resulted in a biphasic calcium dissociation (Quin 2: kobs 1.90 s-1 and 0.73 s-1 respectively; tyrosine: kobs 1.65 s-1 and 0.61 s-1 respectively). Alkylation of intact calmodulin resulted in a four-phase calcium dissociation measured with Quin 2 (kobs 85.3 s-1, 11.1 s-1, 1.92 s-1 and 0.59 s-1); the latter two phases are assumed to represent calcium release from high-affinity sites since they correspond to the biphasic tyrosine fluorescence change in intact alkylated calmodulin (kobs 2.04 s-1 and 0.53 s-1 respectively) and the rate parameters determined in the C-terminal half. Evidently perturbation of the calcium-binding sites by alkylation reduces the rate of calcium dissociation and allows a distinction to be made between dissociation from each of the two high-affinity sites as well as the distinct conformational change on dissociation of each calcium. Alkylation of the N-terminal half resulted in biphasic calcium release with rates (kobs 153 s-1 and 10.9 s-1 respectively) similar to those observed in intact alkylated calmodulin. The rates of calcium dissociation from calmodulin-melittin or fragment-melittin complexes, measured with Quin 2, were slower and monophasic in the C-terminal half (kobs 1.12 s-1), biphasic in the N-terminal half (kobs 140 s-1 and 26.8 s-1 respectively) and triphasic in intact calmodulin (kobs 126 s-1, 12.1 s-1 and 1.38 s-1). Calmodulin antagonists thus increase the apparent calcium affinity of high and low-affinity sites mainly due to a reduced calcium 'off rate', presumably because of conformation restrictions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Ruggiero CE  Dooley DM 《Biochemistry》1999,38(10):2892-2898
The stoichiometry of the topa quinone biogenesis reaction in phenylethylamine oxidase from Arthrobacter globiformis (AGAO) has been determined. We have shown that the 6e- oxidation of tyrosine to topa quinone (TPQ) consumes 2 mol of O2 and produces 1 mol of H2O2/mol of TPQ formed. The rate of H2O2 production is first-order (kobs = 1.0 +/- 0.2 min-1), a rate only slightly lower than the rate of TPQ formation directly determined previously (kobs = 1.5 +/- 0.2 min-1). This gives the following net reaction stoichiometry for TPQ biogenesis: E-Tyr + 2O2 --> E-TPQ + H2O2. This stoichiometry is in agreement with recently proposed mechanisms for TPQ biogenesis, and rules out several possible alternatives.  相似文献   

5.
The kinetics of MgATP-induced electron transfer from the Fe protein (Ac2V) to the VFe protein (AclV) of the vanadium-containing nitrogenase from Azotobacter chroococcum were studied by stopped-flow spectrophotometry at 23 degrees C at pH 7.2. They are very similar to those of the molybdenum nitrogenase of Klebsiella pneumoniae [Thorneley (1975) Biochem. J. 145, 391-396]. Extrapolation of the dependence of kobs. on [MgATP] to infinite MgATP concentration gave k = 46 s-1 for the first-order electron-transfer reaction that occurs with the Ac2V MgATPAclV complex. MgATP binds with an apparent KD = 230 +/- 10 microM and MgADP acts as a competitive inhibitor with Ki = 30 +/- 5 microM. The Fe protein and VFe protein associate with k greater than or equal to 3 x 10(7) M-1.s-1. A comparison of the dependences of kobs. for electron transfer on protein concentrations for the vanadium nitrogenase from A. chroococcum with those for the molybdenum nitrogenase from K. pneumoniae [Lowe & Thorneley (1984) Biochem. J. 224, 895-901] indicates that the proteins of the vanadium nitrogenase system form a weaker electron-transfer complex.  相似文献   

6.
The kinetics of the oxidation of ferrocyanide by lactoperoxidase compound II has been studied over the pH range 5.2-9.9 at 25 degrees C and an ionic strength of 0.11 M. For all pH values, exponential decay curves are obtained for the reaction of compound II in the presence of ferrocyanide which yielded pseudo-first-order rate constants kobs. The spontaneous decay of compound II in the absence of ferrocyanide occurs at an appreciable rate which was measured independently and used in the data analysis. At all pH values two striking effects were observed when the rate of the decay reaction in the presence of ferrocyanide, kobs, was plotted against ferrocyanide concentration: a saturation effect and positive intercepts which are attributable to the spontaneous decay. The plots of kobs versus ferrocyanide concentration were analyzed in terms of the following parameters: a first-order rate constant k3,obs, a Michaelis constant Km,obs and a spontaneous-decay rate constant k4. The parameters k3,obs and Km,obs describe the reaction of compound II with ferrocyanide, independently of the spontaneous decay. The parameter k4 has only a small pH dependence, whereas plots of the logs of k3,obs and Km,obs versus pH have slopes of -1 at high pH. The major part of the pH dependence can be explained by the influence of a single heme-linked acid group in the LPO-compound-II-ferrocyanide complex.  相似文献   

7.
The disintegration behaviour of liposomes in polyoxyethylene lauryl ether (PLE) and sodium cholate solutions was studied by the turbidity disappearance method. In maximally solubilized systems of liposomes, the molar ratios (phosphatidylcholine/surfactant) were 0.43 and 1.8 for PLE and sodium cholate, respectively. The disintegration process of either unilamellar or multilamellar liposomes followed first-order kinetics. Based on a physical model in which liposomes heterogeneous in size were assumed to disintegrate from the outermost shell one by one, a mathematical expression of the turbidity disappearance rate was introduced and applied to explain the data thus obtained. Model calculations suggested that the number of disintegrated shells would not be so large, even if up to 50% reduction of the initial turbidity was observed. From the dependence of the pseudo-first-order rate constant (kobs) on the surfactant concentration for unilamellar liposomes, it was assumed in general that kobs consists of the contributions of the monomer and micellar fractions: for PLE, both fractions shared in the disintegration, but only the micellar fraction with sodium cholate. Furthermore, in the latter case, kobs depended on the initial liposome concentration. These results are likely to be consistent with the proposed modes of surfactant action classified as type A and type B (Helenius, A. and Simons, K. (1975) Biochim. Biophys. Acta 415, 29-79).  相似文献   

8.
A mathematical reaction-diffusion model is defined to describe the gradual decomposition of polymer microspheres composed of poly(D,L-lactic-co-glycolic acid) (PLGA) that are used for pharmaceutical drug delivery over extended periods of time. The partial differential equation (PDE) model treats simultaneous first-order generation due to chemical reaction and diffusion of reaction products in spherical geometry to capture the microsphere-size-dependent effects of autocatalysis on PLGA erosion that occurs when the microspheres are exposed to aqueous media such as biological fluids. The model is solved analytically for the concentration of the autocatalytic carboxylic acid end groups of the polymer chains that comprise the microspheres as a function of radial position and time. The analytical solution for the reaction and transport of the autocatalytic chemical species is useful for predicting the conditions under which drug release from PLGA microspheres transitions from diffusion-controlled to erosion-controlled release, for understanding the dynamic coupling between the PLGA degradation and erosion mechanisms, and for designing drug release particles. The model is the first to provide an analytical prediction for the dynamics and spatial heterogeneities of PLGA degradation and erosion within a spherical particle. The analytical solution is applicable to other spherical systems with simultaneous diffusive transport and first-order generation by reaction.  相似文献   

9.
Irreversible disassembly of the 4Fe-4S cluster in Chromatium vinosum high-potential iron protein (HiPIP) has been investigated in the presence of a low concentration of guanidinium hydrochloride. From the dependence of degradation rate on [H+], it is deduced that at least three protons are required to trigger efficient cluster degradation. Under these conditions the protonated cluster shows broadened M?ssbauer signals, but delta EQ (1.1 mm/s) and delta (0.44 mm/s) are similar to the native form. Collapse of the protonated transition state complex, revealed by rapid-quench M?ssbauer experiments, occurs with a measured rate constant kobs approximately 0.72 +/- 0.35 s-1 that is consistent with results from time-resolved electronic absorption and fluorescence (kobs approximately 0.4 +/- 0.1 s-1) and EPR (kobs approximately 0.62 +/- 0.18 s-1) measurements. Apparently, guanidinium hydrochloride serves to perturb the tertiary structure of the protein, facilitating protonation of the cluster, but not degradation per se. Release of iron ions occurs even more slowly with kobs approximately 0.07 +/- 0.02 s-1, as determined by the appearance of the g = 4.3 EPR signal. Proton-mediated cluster degradation is sensitive to the oxidation state of the cluster, with the oxidized state showing a two-fold slower rate in acidic solutions as a result of increased electrostatic repulsion with the cluster. Consistent results are obtained from absorption, fluorescence, M?ssbauer and EPR measurements.  相似文献   

10.
The uptake and degradation of bovine serum albumin (BSA), bovine liver catalase, and rabbit muscle enolase have been studied in cultured mouse peritoneal macrophages (MPM) and baby hamster kidney fibroblasts (BHK cells). Rates constant for the uptake of the three proteins by MPM were similar. In addition, BSA accumulation was independent of BSA concentration in the uptake medium and was not inhibited by a large excess of serum, suggesting that protein accumulation was by fluid phase pinocytosis. Following an overnight uptake, 20-30% of the accumulated protein was subsequently regurgitated into the medium in a trichloroacetic acid/phosphotungstic acid-precipitable form. This material co-migrated with the authentic protein during molecular sieve chromatography on Sephadex G-50. The rates of appearance of trichloroacetic acid/phosphotungstic acid-insoluble products were greater than expected for cell death and leakage. The observed first order rate constants, kobs, for the appearance of trichloroacetic acid/phosphotungstic acid-soluble and trichloroacetic acid/phosphotungstic acid-insoluble products in the culture medium were identical, indicating that both products were released in parallel from MPM and BHK cells. The kobs for intracellular BSA degradation and regurgitation were independent of the initial BSA concentration in the uptake medium, but were decreased about 35% when degradation was allowed to proceed in the presence of high concentrations of serum. Degradation was also inhibited by chloroquine and pepstatin. Inhibition of degradation was accompanied by an increase in the total amount of regurgitated protein appearing in the medium. Remarkably, however, these inhibitors also decreased kobs for regurgitation, thereby preserving the similarity in the observed rate constants for the appearance of trichloroacetic acid/phosphotungstic acid-soluble and trichloroacetic acid/phosphotungstic acid-insoluble products. These and other results were inconsistent with desorption of proteins from the surface of the culture dish or the surface of cells as the source of the trichloroacetic acid/phosphotungstic acid-insoluble label appearing in the medium.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
Guo WX  Huang KX 《Biopolymers》2004,74(3):248-255
Poly(dimer acid-brassylic acid) [P(DA-BA)] copolymers and poly(dimer acid-pentadecandioic acid) [P(DA-PA)] copolymers were prepared by melt polycondensation of the corresponding mixed anhydride prepolymers. The copolymers were characterized by Fourier transform infrared (FTIR), gel permeation chromatography (GPC), differential scanning calorimetry (DSC), wide angle x-ray powder-diffraction, and thermal gravimetric analysis (TGA). In vitro studies show that all the copolymers are degradable in phosphate buffer at 37 degrees C, and leaving an oily dimer acid residue after hydrolysis for the copolymer with high content of dimer acid. The release profiles of hydrophilic model drug, ciprofloxcin hydrochloride, from the copolymers, follow first-order release kinetics. All the preliminary results suggested that the copolymer might be potentially used as drug delivery devices.  相似文献   

12.
1. Reactions of enzymes with site-specific reagents may involve intermediate adsorptive complexes formed by parallel reactions in several protonic states. Accordingly, a profile of the apparent second-order rate constant for the modification reaction (Kobs., the observed rate constant under conditions where the reagent concentration is low enough for the reaction to be first-order in reagent) against pH can, in general, reflect free-reactant-state molecular pKa values only if a quasi-equilibrium condition exists around the reactive protonic state (EHR) of the adsorptive complex. 2. Usually the condition for quasi-equilibrium is expressed in terms of the rate constants around EHR: (formula: see text) i.e. k mod. less than k-2. This often cannot be assessed directly, particularly if it is not possible to determine kmod. 3. It is shown that kmod. must be much less than k-2, however, if kobs. (the pH-independent value of kobs.) less than k+2. 4. Since probable values of k+2 greater than 10(6)M-1.S-1 and since values of kobs. for many modification reactions less than 10(6)M-1.S-1, the equilibrium assumption should be valid, and kinetic study of such reactions should provide reactant-state pKa values. 5. This may not apply to catalyses, because for them the value of kcat./Km may exceed 5 X 10(5)M-1.S-1. 6. The conditions under which the formation of an intermediate complex by parallel pathways may come to quasi-equilibrium are discussed in the Appendix.  相似文献   

13.
Chemical modification of ferredoxin--NADP+ reductase from the cyanobacteria Anabaena has been performed using the alpha-dicarbonyl reagent phenylglyoxal. Inactivation of both the diaphorase and cytochrome-c reductase activities, characteristic of the enzyme, indicates the involvement of one or more arginyl residues in the catalytic process of the enzyme. The determination of the rate constants for the inactivation process under different conditions, including those in which substrates, NADP+ and ferredoxin, as well as other NADP+ analogs were present, indicates the involvement of two different groups in the inactivation process, one that reacts very rapidly with the reagent (kobs = 8.3 M-1 min-1) and is responsible for the binding of NADP+, and a second less reactive group (kobs = 0.9 M-1 min-1), that is involved in the binding of ferredoxin. Radioactive labeling of the enzyme with [14C]phenylglyoxal confirms that two groups are modified while amino acid analysis of the modified protein indicates that the modified groups are arginine residues. The identification of the amino acid residues involved in binding and catalysis of the substrates of ferredoxin--NADP+ reductase will help to elucidate the mechanism of the reaction catalyzed by this important enzyme.  相似文献   

14.
Treatment of the Cu(II)-Fe(III) derivative of pig allantoic fluid acid phosphatase with hydrogen peroxide caused irreversible inactivation of the enzyme and loss of half of the intensity of the visible absorption spectrum. Phosphate, a competitive inhibitor, protected against this inactivation, suggesting that it occurred as a result of a reaction at the active site. The native Fe(II)-Fe(III) enzyme was irreversibly inactivated by H2O2 to a much smaller extent than the Cu(II)-Fe(III) derivative, whereas the Zn(II)-Fe(III) derivative was stable to H2O2 treatment. The rates of inactivation of the Cu(II)-Fe(III) and Fe(II)-Fe(III) enzymes in the presence of H2O2 were increased by addition of ascorbate. These results suggest involvement of a Fenton-type reaction, generating hydroxyl radicals which react with essential active site groups. Experiments carried out on the Fe(II)-Fe(III) enzyme showed that irreversible inactivation by H2O2 in the presence of ascorbate obeyed pseudo first-order kinetics. A plot of kobs for this reaction against H2O2 concentration (at saturating ascorbate) was hyperbolic, giving kobs(max) = 0.41 +/- 0.025 min-1 and S0.5(H2O2) = 1.16 +/- 0.18 mM. A kinetic scheme is presented to describe the irreversible inactivation, involving hydroxyl radical generation by reaction of H2O2 with Fe(II)-Fe(III) enzyme, reduction of the product Fe(III)-Fe(III) enzyme by ascorbate and reaction of hydroxyl radical with an essential group in the enzyme.  相似文献   

15.
Calcium release from high and low-affinity calcium-binding sites of intact bovine brain calmodulin (CaM) and from the tryptic fragment 78-148, purified by high-pressure liquid chromatography, containing only the high-affinity calcium-binding sites, was determined by fluorescence stopped-flow with 2-p-toluidinylnaphthalene sulfonate (TNS). The tryptic fragments 1-77 and 78-148 each contain a calcium-dependent TNS-binding site, as shown by the calcium-dependent increase in TNS fluorescence. The rate of the monophasic fluorescence decrease in endogenous tyrosine on calcium dissociation from intact calcium-saturated calmodulin (kobs 10.8 s-1 and 3.2 s-1 at 25 degrees C and 10 degrees C respectively) as well as the rate of equivalent slow phase of the biphasic decrease in TNS fluorescence (kobsslow 10.6 s-1 and 3.0 s-1 at 25 degrees C and 10 degrees C respectively) and the rate of the solely monophasic decrease in TNS fluorescence, obtained with fragment 78-148 (kobs 10.7 s-1 and 3.5 s-1 at 25 degrees C and 10 degrees C respectively), were identical, indicating that the rate of the conformational change associated with calcium release from the high-affinity calcium-binding sites on the C-terminal half of calmodulin is not influenced by the N-terminal half of the molecule. The fast phase of the biphasic decrease of TNS fluorescence, observed by the N-terminal half of the molecule. The fast phase of the biphasic decrease of TNS fluorescence, observed with intact calmodulin only (kobsfast 280 s-1 at 10 degrees C) but not with fragment 78-148, is most probably due to the conformational change associated with calcium release from low-affinity sites on the N-terminal half. The calmodulin fragments 1-77 and 78-148 neither activated calcium/calmodulin-dependent protein kinase of cardiac sarcoplasmic reticulum nor inhibited calmodulin-dependent activation at a concentration approximately 1000-fold greater (5 microM) than that of the calmodulin required for half-maximum activation (5.9 nM at 0.8 mM Ca2+ and 5 mM Mg2+) of calmodulin-dependent phosphoester formation.  相似文献   

16.
The pre-steady-state kinetics of MgATP hydrolysis by nitrogenase from Klebsiella pneumoniae were studied by stopped-flow calorimetry at 6 degrees C and at pH 7.0. An endothermic reaction (delta Hobs. = +36 kJ.mol of ATP-1; kobs. = 9.4 s-1) in which 0.5 proton.mol of ATP-1 was released, has been assigned to the on-enzyme cleavage of MgATP to yield bound MgADP + Pi. The assignment is based on the similarity of these parameters to those of the corresponding reaction that occurs with rabbit muscle myosin subfragment-1 (delta Hobs. = +32 kJ.mol of ATP-1; kobs. = 7.1 s-1; 0.2 proton released.mol of ATP-1) [Millar, Howarth & Gutfreund (1987) Biochem. J. 248, 683-690]. MgATP-dependent electron transfer from the nitrogenase Fe-protein to the MoFe-protein was monitored by stopped-flow spectrophotometry at 430 nm and occurred with kobs. value of 3.0 s-1 at 6 degrees C. Thus, under these conditions, hydrolysis of MgATP precedes electron transfer within the protein complex. Evidence is presented that suggests that MgATP cleavage and subsequent electron transfer are reversible at 6 degrees C with an overall equilibrium constant close to unity, but that, at 23 degrees C, the reactions are essentially irreversible, with an overall equilibrium constant greater than or equal to 10.  相似文献   

17.
The influence of chemical modification of arginine residues (using phenylglyoxal) in ferredoxin-NADP+ reductase (FNR), and of carboxyl groups (using glycine ethyl ester) in flavodoxin (Fld), on the kinetics of electron transfer between FNR and Fld, and between ferredoxin (Fd) and FNR, was examined using laser flash photolysis methods. All proteins were obtained from the cyanobacterium Anabaena PCC7119. Reduction by laser-generated 5-deazariboflavin semiquinone of the FAD moiety of phenylglyoxal-modified FNR occurred with a second-order rate constant 2.5-fold smaller than that obtained for reduction of native FNR, indicating either a small degree of steric hindrance of the cofactor, or a decrease in its redox potential, upon chemical modification. In contrast, no changes were found in the kinetics of reduction of the FMN cofactor of Fld modified by glycine ethyl ester as compared with the native protein. The observed rate constants for reoxidation of Fdred (reduced Fd) by FNRox (oxidized FNR) were dramatically decreased (approximately 100-fold) when phenylglyoxal-modified FNR was used. In contrast to the reaction involving the native proteins, no ionic strength effects on kobs values were found. These results, and those obtained upon varying the protein concentration, indicate that the rate constant for complex formation and the attractive electrostatic interaction between the two proteins were greatly diminished by chemical modification of arginine residues of FNR. When phenylglyoxal-modified FNRsq (FNR semiquinone) was used to reduce Fldox (oxidized Fld), similar inhibitory effects were observed. In this case, the limiting first-order rate constant for Fldsq (Fld semiquinone) formation via intracomplex electron transfer from FNRsq was approximately 12-fold smaller than that obtained for the native FNR (600 s-1 vs 7000 s-1). Again, ionic strength effects were diminished. The glycine-ethyl-ester-modified Fld yielded a limiting first-order rate constant for intracomplex electron transfer from FNRsq to Fldox which was approximately 7-fold smaller (1000 s-1) than that obtained with native Fld, and ionic strength effects were again diminished. These results indicate that complex formation can still occur between modified FNR and native Fld, and between native FNR and modified Fld, but that the geometry of these complexes is altered so as to decrease the effectiveness of interprotein electron transfer. The results are discussed in terms of the specific structural features of the proteins involved.  相似文献   

18.
Addition of sodium fluoride in the millimolar concentration range to a solution containing the sarcoplasmic reticulum CaATPase undergoing turnover in its vesicular or nonionic detergent-solubilized forms produced a slow (time range of minutes) complete loss of enzymatic activity. In the presence of magnesium and the absence of calcium, similar results were obtained under nonturnover conditions. Time courses were adequately fit by a function corresponding to a monophasic transformation with a pseudo first order rate constant kobs. In the absence of Mg2+ (EDTA present) no inhibition developed; kobs depended hyperbolically on the Mg2+ concentration with the half maximal effect occurring near 4 mM. The fluoride concentration dependence of kobs showed no evidence of approaching saturation (highest [F-] used was 40 mM) and corresponded to a rate law which was approximately second-order with respect to fluoride. A number of ligands known to bind to the CaATPase were found to decrease kobs. Calcium prevented onset of fluoride inhibition with a midpoint in the micromolar range, implying an effect due to binding at the high affinity transport sites. ATP also protected with a midpoint in the micromolar range, consistent with an effect caused by active site binding of the nucleotide; protection was only partial, suggesting the ATPase can bind fluoride and ATP simultaneously. Prevention of fluoride inhibition by Pi occurred with a [Pi]1/2 of 12 mM at pH 6.5, a concentration similar to that which produces active site phosphorylation. Finally, protection by orthovanadate was found to be competitive and have a midpoint of 5 microM. These results point to an effect exerted at or near the phosphorylation site. The value of kobs increased from essentially zero above pH 8 to a plateau below pH 6; the transition had a midpoint near pH 7.2. Inhibition persisted after removal (with EGTA present) of unbound fluoride by dialysis. Reversal of fluoride inhibition was very slow, with a t1/2 of 16 h at 37 degrees C. These results suggest that fluoride behaves like a slow, tight-binding inhibitor of the ATPase and that the resulting complex is a stable transition (or intermediate) state analog. Plausible molecular bases for our results are that fluoride acts at the phosphorylation site as an analog of Pi or of hydroxide, which may be considered a substrate in the normal hydrolysis of the phosphorylated enzyme. A role for aluminum was ruled out after finding that the addition of EGTA to 10 mM or aluminum sulfate to 0.2 mM or deferoxamine to 0.5 mM produced no significant change in kobs.  相似文献   

19.
Stopped flow spectrophotometry was used to investigate the kinetics of the transition of the phosphoglycerate dehydrogenase (3-phosphoglycerate: NAD oxidoreductase, EC 1.1.1.95) reaction from the active to the inhibited rate upon the addition of the physiological inhibitor serine. The transition was characterized by a single first order rate constant (kobs,i) which was independent of enzyme concentration. At pH 8.5, kobs,i increased in a hyperbolic manner with serine concentration from 2 to 8 s-1. The increase in kobs,i occurred at serine concentrations where the steady state inhibition was virtually complete. These results indicate that serine inhibition is an allosteric process involving a conformational change in the enzyme. A model is presented in which serine at low concentrations binds exclusively to the inhibited state of the enzyme and shifts the equilibrium toward that state; at high serine concentrations, serine binds to the active state, facilitating its conversion to the inhibited state. An alternative model, which we favor, proposes two classes of inhibitor binding sites. The kinetics of the fluorescence quenching of enzyme-bound NADH by serine (Sugimoto, E., and Pizer, L.I. (1968) J. Biol. Chem. 243, 2090-2098), measured by stopped flow fluorimetry, was also characterized by a single first order rate constant (kobs,f.q.) which was independent of enzyme concentration. At pH 8.5, kobs,f.q. ranged from 0.4 s-1 at low serine concentrations to 1.1 s-1 at high serine concentrations. These results indicate that the fluorescence quenching induced by serine is a manifestation of a structural change in the enzyme. Enzyme and excess NADH were mixed with substrate and serine in the stopped flow instrument, and enzyme-bound NADH fluorescence was monitored by exciting through the protein at 285 nm. A rapid fluorescence quenching process, which occurred within the mixing time, was followed by a slower fluorescence enhancement process which terminated in a steady state level corresponding to the quenched fluorescence of the enzyme NADH serine complex. The rapid quenching was the result of substrate binding (Dubrow, R., and Pizer, L.I. (1977) J. Biol. Chem. 252, 1539-1551). The fluorescence enhancement was characterized by a single first order rate constant whose value for a given serine concentration corresponded with Kobs,j. This data shows that the quenched state of the enzyme-NADH-complex is the state which is directly responsible for the inhibition of enzyme activity. During catalysis the quenched state is achieved from a different initial conformation, and consequently at a different rate, than in the absence of substrate. kobs,j and kobs,f.q. were also measured using glycine, another inhibitor. The ultraviolet difference spectrum between enzyme and enzyme plus serine was determined and proposed to be the result of the same structural change which is responsible for the fluorescence quenching by serine.  相似文献   

20.
To give the first demonstration of neighboring group-controlled drug delivery rates, a series of novel, polymerizable ester drug conjugates was synthesized and fully characterized. The monomers are suitable for copolymerization in biomaterials where control of drug release rate is critical to prophylaxis or obviation of infection. The incorporation of neighboring group moieties differing in nucleophilicity, geometry, and steric bulk in the conjugates allowed the rate of ester hydrolysis, and hence drug liberation, to be rationally and widely controlled. Solutions (2.5 x 10-5 mol dm-3) of ester conjugates of nalidixic acid incorporating pyridyl, amino, and phenyl neighboring groups hydrolyzed according to first-order kinetics, with rate constants between 3.00 +/- 0.12 x 10-5 s -1 (fastest) and 4.50 +/- 0.31 x 10- 6 s-1 (slowest). The hydrolysis was characterized using UV-visible spectroscopy. When copolymerized with poly(methyl methacrylate), free drug was shown to elute from the resulting materials, with the rate of release being controlled by the nature of the conjugate, as in solution. The controlled molecular architecture demonstrated by this system offers an attractive class of drug conjugate for the delivery of drugs from polymeric biomaterials such as bone cements in terms of both sustained, prolonged drug release and minimization of mechanical compromise as a result of release. We consider these results to be the rationale for the development of "designer" drug release biomaterials, where the rate of required release can be controlled by predetermined molecular architecture.  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