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.
D Y Cai  M Tien 《Biochemistry》1990,29(8):2085-2091
The oxycomplexes (compound III, oxyperoxidase) of two lignin peroxidase isozymes, H1 (pI = 4.7) and H8 (pI = 3.5), were characterized in the present study. After generation of the ferroperoxidase by photochemical reduction with deazoflavin in the presence of EDTA, the oxycomplex is formed by mixing ferroperoxidase with O2. The oxycomplex of isozyme H8 is very stable, with an autoxidation rate at 25 degrees C too slow to measure at pH 3.5 or 7.0. In contrast, the oxycomplex of isozyme H1 has a half-life of 52 min at pH 4.5 and 29 min at pH 7.5 at 25 degrees C. The decay of isozyme H1 oxycomplex follows a single exponential. The half-lives of lignin peroxidase oxycomplexes are much longer than those observed with other peroxidases. The binding of O2 to ferroperoxidase to form the oxycomplex was studied by stopped-flow methods. At 20 degrees C, the second-order rate constants for O2 binding are 2.3 X 10(5) and 8.9 X 10(5) M-1 s-1 for isozyme H1 and 6.2 X 10(4) and 3.5 X 10(5) M-1 s-1 for isozyme H8 at pH 3.6 and pH 6.8, respectively. The dissociation rate constants for the oxycomplex of isozyme H1 (3.8 Z 10(-3) s-1) and isozyme H8 (1.0 X 10(-3) s-1) were measured at pH 3.6 by CO trapping. Thus, the equilibrium constants (K, calculated from kon/koff) for both isozymes H1 (7.0 X 10(7) M-1) and H8 (6.2 X 10(7) M-1) are higher than that of myoglobin (1.9 Z 10(6) M-1).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The kinetic parameters for transport of the nonmetabolizable glucose analogue 3-O-methyl-D-glucose and the relationship between transport and metabolism of D-glucose and D-fructose were determined in isolated rat hepatocytes at 37 degrees C and pH 7.4. 3-O-Methylglucose at a very low concentration (0.1 mM) equilibrated with the intracellular water with a rate constant of 0.41 s-1. Km for equilibrium exchange entry was 5.5 mM and Vmax was 2.2 mM X s-1 and similar results were obtained when using the zero-trans entry protocol. The rate constant for entry of tracer D-glucose was 0.15 s-1 and Km for glucose was about 20 mM. The phosphorylation rate for D-glucose was much slower than the transport rate. The rate constant for D-fructose entry was about 0.04 s-1, the apparent Km was about 100 mM and Vmax about 5 mM X s-1. The concentration dependence of 3-O-methylglucose inhibition of labelled fructose transport revealed biphasic kinetics indicating that fructose was transferred by both the glucose transporter and a fructose transporter. At concentrations lower than 1 mM, fructose metabolism appeared to be limited by the transport step.  相似文献   

3.
Urea permeability of human red cells   总被引:5,自引:1,他引:4       下载免费PDF全文
The rate of unidirectional [14C]urea efflux from human red cells was determined in the self-exchange and net efflux modes with the continuous flow tube method. Self-exchange flux was saturable and followed simple Michaelis-Menten kinetics. At 38 degrees C the maximal self-exchange flux was 1.3 X 10(-7) mol cm-2 s-1, and the urea concentration for half-maximal flux, K1/2, was 396 mM. At 25 degrees C the maximal self-exchange flux decreased to 8.2 X 10(-8) mol cm-2 s-1, and K1/2 to 334 mM. The concentration-dependent urea permeability coefficient was 3 X 10(-4) cm s-1 at 1 mM and 8 X 10(-5) cm s-1 at 800 mM (25 degrees C). The latter value is consonant with previous volumetric determinations of urea permeability. Urea transport was inhibited competitively by thiourea; the half-inhibition constant, Ki, was 17 mM at 38 degrees C and 13 mM at 25 degrees C. Treatment with 1 mM p-chloromercuribenzosulfonate inhibited urea permeability by 92%. Phloretin reduced urea permeability further (greater than 97%) to a "ground" permeability of approximately 10(-6) cm s-1 (25 degrees C). This residual permeability is probably due to urea permeating the hydrophobic core of the membrane by simple diffusion. The apparent activation energy, EA, of urea transport after maximal inhibition was 59 kJ mol-1, whereas in control cells EA was 34 kJ mol-1 at 1 M and 12 kJ mol-1 at 1 mM urea. In net efflux experiments with no extracellular urea, the permeability coefficient remained constantly high, independent of a variation of intracellular urea between 1 and 500 mM, which indicates that the urea transport system is asymmetric. It is concluded that urea permeability above the ground permeability is due to facilitate diffusion and not to diffusion through nonspecific leak pathways as suggested previously.  相似文献   

4.
Exchange rates were calculated as a function of pH from line widths of methylamine resonances in 13C-NMR spectra of N epsilon,N epsilon,N alpha,N alpha-[13C]tetramethyllysine (TML) and N epsilon,N epsilon,N alpha,N alpha-tetramethyllysine methyl ester (TMLME). The pH dependence of the dimethyl alpha-amine exchange rate could be adequately described by assuming base-catalyzed chemical exchange between two diastereotopic methyl populations related by nitrogen inversion. Deprotonation of the alpha-amine was assumed to occur by proton transfer to (1) OH-, (2) water, (3) a deprotonated amine or (4) RCO2-. Microscopic rate constants characterizing each of these transfer processes (k1, k2, k3 and k4, respectively) were determined by fitting the rates calculated from line width analysis to a steady-state kinetic model. Using this procedure it was determined that for both TML and TMLME k2 approximately equal to 1-10 M-1 s-1, k3 approximately equal to 10(6) M-1 s-1 and ki, the rate constant for nitrogen inversion was about 10(8)-10(9) s-1. Upper limits of 10(12) and 10(3) M-1 s-1 could be determined for k1 and k4, respectively. A similar kinetic analysis was used to explain pH-dependent line-broadening effects observed for the N-terminal dimethylalanyl resonance in 13C-NMR spectra of concanavalin A, reductively methylated using 90% [13C]formaldehyde. From exchange data below pH 4 it could be determined that amine inversion was limited by the proton transfer rate to the solvent, with a rate constant estimated at 20 M-1 s-1. Above pH 4, exchange was limited by proton transfer to other titrating groups in the protein structure. Based upon their proximity, the carboxylate side chains of Asp-2 and Asp-218 appear to be likely candidates. The apparent first-order microscopic rate constant characterizing proton transfer to these groups was estimated to be about 1 X 10(4) s-1. Rate constants characterizing nitrogen inversion (ki), proton transfer to OH- (k1) and proton transfer to the solvent (k2) were estimated to be of the same order of magnitude as those determined for the model compounds. On the basis of our results, it is proposed that chemical exchange processes associated with base-catalyzed nitrogen inversion may contribute to 15N or 13C spin-lattice relaxation times in reductively methylated peptides or proteins.  相似文献   

5.
The putative intermediate dienol (2) in the steroid isomerase (KSI) catalyzed conversion of 5-androstene-3,17-dione (1) to 4-androstene-3,17-dione (3) has been independently generated and tested as a substrate for KSI. At pH 7, dienol 2 is converted by KSI to a mixture of 1 (46%) and 3 (54%). The apparent second-order rate constant for reaction of 2 with KSI to produce 3 (kappa cat/Km = 2.3 x 10(8) M-1 s-1) is similar to that for reaction of 1 with KSI (kappa cat/Km = 2.1 x 10(8) M-1 s-1), demonstrating that 2 is kinetically competent. Isomerization of 1 by KSI in D2O gives only 5% of solvent deuterium incorporated into the product 3. When 2 reacts with KSI in D2O, and the product 3 is isolated (from direct reaction of 2 and from subsequent conversion of the 1 initially formed), ca. 80 atom % deuterium is located at C-6 beta, confirming that protonation of the dienol by KSI occurs at the same face as the proton transfer in the KSI catalyzed reaction of 1 to 3.  相似文献   

6.
The time dependence of the release of fibrinopeptides from fibrinogen was studied as a function of the concentration of fibrinogen, thrombin, and Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization. The release of fibrinopeptides during fibrin assembly was shown to be a highly ordered process. Rate constants for individual steps in the formation of fibrin were evaluated at pH 7.4, 37 degrees C, gamma/2 = 0.15. The initial event, thrombin-catalyzed proteolysis at Arg-A alpha 16 to release fibrinopeptide A (kcat/Km = 1.09 X 10(7) M-1s-1) was followed by association of the resulting fibrin I monomers. Association of fibrin I was found to be a reversible process with rate constants of 1 X 10(6) M-1s-1 and 0.064 s-1 for association and dissociation, respectively. Assuming random polymerization of fibrin I monomer, the equilibrium constant for fibrin I association (1.56 X 10(7) M-1) indicates that greater than 80% of the fibrin I protofibrils should contain more than 10 monomeric units at 37 degrees C, pH 7.4, when the fibrin I concentration is 1.0 mg/ml. Association of fibrin I monomers was shown to result in a 6.5-fold increase in the susceptibility of Arg-B beta 14 to thrombin-mediated proteolysis. The 6.5-fold increase in the observed specificity constant from 6.5 X 10(5) M-1s-1 to 4.2 X 10(6) M-1s-1 upon association of fibrin I monomers and the rate constant for fibrin association indicates that most of the fibrinopeptide B is released after association of fibrin I monomers. The interaction between a pair of polymerization sites in fibrin I dimer was found to be weaker than the interaction of fibrin I with Gly-Pro-Arg-Pro and weaker than the interaction of fibrin I with fibrinogen.  相似文献   

7.
K Yusa  K Shikama 《Biochemistry》1987,26(21):6684-6688
Hydrogen peroxide, one of the potent oxidants in muscle tissues, can induce very rapid oxidation of oxymyoglobin (MbO2) to metmyoglobin (metMb) with an apparent rate constant of 7.5 X 10(4) h-1 M-1 (i.e., 20.8 s-1 M-1) over the wide pH range of 5.5-10.2 in 0.1 M buffer at 25 degrees C. Its molecular mechanism, however, is quite different from that of the autoxidation of MbO2 to metMb. Kinetic analysis has revealed that the hydrogen peroxide oxidation proceeds through the formation of ferryl-Mb(IV) from deoxy-Mb(II), which is in equilibrium with MbO2, by a two-equivalent oxidation with H2O2. Once the ferryl species is formed, it reacts rapidly with another deoxy-Mb(II) in a bimolecular fashion so as to yield 2 mol of metMb(III). Under physiological conditions, the rate-determining step was the oxidation of the deoxy species by H2O2, its rate constant being estimated to be on the order of 3.6 X 10(3) s-1 M-1 at 25 degrees C. These findings leads us to the view that a good supply of dioxygen provides rather an important defense against the oxidation of myoglobin with hydrogen peroxide in cardiac and skeletal muscle tissues.  相似文献   

8.
31P NMR measurement of ATP synthesis rate in perfused intact rat hearts   总被引:1,自引:0,他引:1  
Using 31P NMR and the saturation-transfer method, the unidirectional rate of ATP synthesis was measured in isolated, Langendorff-perfused, isovolumic rat hearts operating at a rate pressure product of 25.6 +/- 2.5 (SE) X 10(3) mmHg X min-1 and consuming O2 at a rate of 35 +/- 2 mumol O2 X min-1 X (g dry wt)-1, at 37 degrees C. This rate was 7.2 +/- 0.9 mumol X s-1 X (g dry wt)-1 and was related to the rate of oxygen atom consumption by a ratio of 6.3 +/- 0.9. These data show that in the intact heart the unidirectional rate of ATP synthesis exceeds the net rate of ATP synthesis and consumption by approximately a factor of 2.  相似文献   

9.
Bilirubin oxidase has been used to probe the surface topography of phycocyanins (C-phycocyanin and phycocyanin-645), peridinin-chlorophyll a-protein and phytochrome. The enzyme catalyzes oxidation of the tetrapyrrolic chromophores in these proteins. Relative rates of oxidation were 78.0 X 10(-6) s-1 (monitored at 617 nm) and 58.0 X 10(-6) s-1 (592 nm) for C-phycocyanin, 43.0 X 10(-6) s-1 for phycocyanin-645, 0.3 X 10(-6) s-1 (at 671 nm) and 1.3 X 10(-6) s-1 (at 480 nm) for peridinin-chlorophyll a-protein. The relative rate of free chlorophyllin a was 2.8 X 10(4) s-1 whereas upon binding to human serum albumin its rate of oxidation was reduced to 3.3 X 10(-3) s-1. Relative rates for the oxidation of Pr and Pfr forms of phytochrome were 2.9 and 19.5 s-1, respectively, which are consistent with earlier finding [( 1984) Plant Physiol. 74, 755-758] that indicated a preferential exposure of tetrapyrrolic chromophore in the Pfr form. In general, kcat/Km values derived from the Lineweaver-Burk plots followed the same trend as the relative rates of oxidation. For example, the kcat/Km for the free chlorophyllin a was 2.8 X 10(6) M-1 s-1 but it was only 1.1 M-1s-1 for the chlorophyll a in peridinin-chlorophyll a-protein where the chlorophyll is shielded by protein. These results reflect varying degrees of protection of the tetrapyrrolic chromophores from the enzymatic oxidation and prove that bilirubin oxidase can be generally used as a probe for deducing the topography of tetrapyrrolic chromophores.  相似文献   

10.
S W King  V R Lum  T H Fife 《Biochemistry》1987,26(8):2294-2300
The carbamate ester N-(phenoxycarbonyl)-L-phenylalanine binds well to carboxypeptidase A in the manner of peptide substrates. The ester exhibits linear competitive inhibition toward carboxypeptidase A catalyzed hydrolysis of the amide hippuryl-L-phenylalanine (Ki = 1.0 X 10(-3) M at pH 7.5) and linear noncompetitive inhibition toward hydrolysis of the specific ester substrate O-hippuryl-L-beta-phenyllactate (Ki = 1.4 X 10(-3) M at pH 7.5). Linear inhibition shows that only one molecule of inhibitor is bound per active site at pH 7.5. The hydrolysis of the carbamate ester is not affected by the presence of 10(-8)-10(-9) M enzyme (the concentrations employed in inhibition experiments), but at an enzyme concentration of 3 X 10(-6) M catalysis can be detected. The value of kcat at 30 degrees C, mu = 0.5 M, and pH 7.45 is 0.25 s-1, and Km is 1.5 X 10(-3) M. The near identity of Km and Ki shows that Km is a dissociation constant. Substrate inhibition can be detected at pH less than 7 but not at pH values above 7, which suggests that a conformational change is occurring near that pH. The analogous carbonate ester O-(phenoxycarbonyl)-L-beta-phenyllactic acid is also a substrate for the enzyme. The Km is pH independent from pH 6.5 to 9 and has the value of 7.6 X 10(-5) M in that pH region. The rate constant kcat is pH independent from pH 8 to 10 at 30 degrees C (mu = 0.5 M) with a limiting value of 1.60 s-1. Modification of the carboxyl group of glutamic acid-270 to the methoxyamide strongly inhibits the hydrolysis of O-(phenoxycarbonyl)-L-beta-phenyllactic acid. Binding of beta-phenyllactate esters and phenylalanine amides must occur in different subsites, but the ratios of kcat and kcat/Km for the structural change from hippuryl to phenoxy in each series are closely similar, which suggests that the rate-determining steps are mechanistically similar.  相似文献   

11.
M W Washabaugh  W P Jencks 《Biochemistry》1988,27(14):5044-5053
Rate constants for C(2)-proton exchange from thiamin, N(1')-methylthiamin, and several 3-substituted-4-methylthiazolium ions catalyzed by D2O and deuterioxide ion were determined by 1H NMR at 30 degrees C and ionic strength 2.0 M. Values of pKa for the thiazolium ions, including thiamin itself, were found to be in the range pKa = 17-19; the pKa values for N(1')-protonated thiamin and free thiamin C(2)-H in H2O are 17.7 and 18.0, respectively. The pKa value for N(1')-protonated thiamin was calculated from the observed rate constant for the pD-independent reaction with D2O after correction for a secondary solvent deuterium isotope effect of kH2O/kD2O = 2.6. The pKa value for free thiamin was calculated from the rate constant for catalysis by OD- after correction by a factor of 3.3 = 8/2.4 for an 8-fold negative deviation of kOD from the Br?nsted plot of slope 1.0 for general base catalysis and a secondary solvent isotope effect of kOD/kOH = 2.4. Values of k-a = 2 X 10(10) and 3 X 10(9) M-1 s-1 were assumed for diffusion-controlled protonation of the C(2) ylide in the reverse direction by H3O+ and H2O, respectively. The Hammett rho I value for the exchange reaction catalyzed by deuterioxide ion or D2O is 8.4 +/- 0.2. There is no positive deviation of the rate constants for free or N(1')-substituted thiamin analogues in either Hammett correlation. This shows that the aminopyrimidinyl group does not provide significant intramolecular catalysis of nonenzymic C(2)-proton removal in the coenzyme.  相似文献   

12.
Product formation during the oxidation of xanthine oxidase has been examined directly by using cytochrome c peroxidase as a trapping agent for hydrogen peroxide and the reduction of cytochrome c as a measure of superoxide formation. When fully reduced enzyme is mixed with high concentrations of oxygen, 2 molecules of H2O2/flavin are produced rapidly, while 1 molecule of O2-/flavin is produced rapidly and another produced much more slowly. Time courses for superoxide formation and those for the absorbance changes due to enzyme oxidation were fitted successfully to the mechanism proposed earlier (Olson, J. S., Ballou, D. P., Palmer, G., and Massey, V. (1974) J. Biol. Chem. 249, 4363-4382). In this scheme, each oxidative step is initiated by the very rapid and reversible formation of an oxygen.FADH2 complex (the apparent KD = 2.2 X 10(-4) M at 20 degrees C, pH 8.3). In the cases of 6- and 4-electron-reduced enzyme, 2 electrons are transferred rapidly (ke = 60 s-1) to generate hydrogen peroxide and partially oxidized xanthine oxidase. In the case of the 2-electron-reduced enzyme, only 1 electron is transferred rapidly and superoxide is produced. The remaining electron remains in the iron-sulfur centers and is removed slowly by a second order process (ks = 1 X 10(4) M-1 s-1). When the pH is decreased from 9.9 to 6.2, both the apparent KD for oxygen binding and the rapid rate of electron transfer are decreased about 20-fold. This result is suggestive of uncompetitive inhibition and implies that proton binding to the enzyme-flavin active site affects primarily the rate of electron transfer, not the formation of the initial oxygen complex.  相似文献   

13.
Acetyldithio-CoA has been shown to be a competent nucleophilic substrate but not an electrophilic substrate for the Claisen condensation catalyzed by thiolase, which normally dimerizes acetyl (Ac)-CoA to acetoacetyl-CoA. Acting as the nucleophile, the kcat/Km for dithioacetyl-CoA is comparable to that of Ac-CoA, the normal substrate. With acetoacetyl-pantetheine acetylating the thiolase to provide the electrophile, the kcat and kcat/Km for the Claisen condensation are 2.1 s-1 and 8.3 X 10(4) M-1 s-1, respectively. The product of the reaction is 3-ketobutyryldithio-CoA. The 3-ketobutyryldithio-CoA has a spectrally determined pKa of 6.55 and the enolate has a lambda max of 357 nm, epsilon 357 = 21,000 cm-1 M-1. Product analysis indicates that acetyldithio-CoA does not serve as the electrophilic partner in the enzymic condensation. This failure is attributed to the inability demonstrated in this study of acetyldithio-CoA to thioacetylate the active site Cys89 of the Zoogloea ramigera thiolase. 1H NMR studies in D2O indicate that thiolase catalyzes the exchange of the alpha-hydrogens, without Cys89 being acetylated, with a rate of 0.63 +/- 0.25 s-1. In the presence of a large excess of acetoacetyl-pantetheine, present to acetylate Cys89 and prevent the thiolytic back reaction, solvent exchange of the alpha-hydrogens can still be detected by observing the isotope-shifted 13C NMR spectrum of [2-13C]acetyldithio-CoA. The exchange of the acetyldithio-CoA alpha-hydrogens with solvent promoted by the acetylated enzyme, must proceed at a rate comparable to that of the condensation reaction.  相似文献   

14.
The H(+)-ATPase from chloroplasts (CF0F1) was isolated, purified and reconstituted into liposomes from phosphatidylcholine/phosphatidic acid. A transmembrane pH difference, delta pH, and a transmembrane electric potential difference, delta psi, were generated by an acid/base transition. The rate of ATP synthesis was measured at constant delta pH and constant delta psi as a function of temperature between 5 degrees C and 45 degrees C. The activation energy was 55 kJ mol-1. CF0F1 was coreconstituted with bacteriorhodopsin at a molar ratio of approximately 1:170 in the same type of liposomes. Illumination of the proteoliposomes leads to proton transport into the vesicles generating a constant delta pH = 1.8. The dependence of the rate of ATP synthesis on ADP concentration was measured with CF0F1 in the oxidized state, E(ox), and in the reduced state, E(red). The results can be described by Michaelis-Menten kinetics with the following parameters: Vmax = 0.5 s-1, Km = 8 microM for E(ox) and Vmax = 2.0 s-1, Km = 8 microM for E(red).  相似文献   

15.
Steady-state kinetic parameters were compared for the action of alpha- and gamma-thrombin on the physiologically important thrombin substrates fibrinogen and factor XIII at 37 degrees C, pH 7.4, and 0.14 M NaCl. gamma-Thrombin, an alpha-thrombin derivative proteolytically cleaved at R-B73 and K-B154, was observed to catalyze the release of fibrinopeptide A (FPA) from fibrinogen with a specificity constant (kcat/Km) of 5 X 10(3) M-1 s-1. This value was approximately 2400-fold lower than the specificity constant for the corresponding alpha-thrombin-catalyzed reaction. The low specificity constant was attributed to an increase in Km and a decrease in kcat for gamma-thrombin-catalyzed release of FPA from fibrinogen. Conversion of alpha-thrombin to gamma-thrombin also resulted in an approximately 800-fold reduction in the specificity constant for thrombin-catalyzed release of fibrinopeptide B (FPB) from fibrin I, as well as a loss in discriminatory power. Whereas alpha-thrombin preferentially released FPA from intact fibrinogen, gamma-thrombin released FPA and FPB from intact fibrinogen at similar rates. In contrast to the large difference in specificity constants observed for alpha- and gamma-thrombin catalysis with fibrin(ogen) as substrate, the specificity constant (2.6 X 10(4) M-1 s-1) observed for gamma-thrombin-catalyzed release of activation peptide from factor XIII was only 5-fold lower than the corresponding value for the alpha-thrombin-catalyzed reaction. Additionally, the promotion of factor XIII activation by fibrin characteristic of the alpha-thrombin-catalyzed reaction did not occur in the gamma-thrombin-catalyzed reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
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   

17.
Glycolate oxidase from spinach has been expressed in Saccharomyces cerevisiae. The active enzyme was purified to near-homogeneity (purification factor approximately 1400-fold) by means of hydroxyapatite and anion-exchange chromatography. The purified glycolate oxidase is nonfluorescent and has absorbance peaks at 448 (epsilon = 9200 M-1 cm-1) and 346 nm in 0.1 M phosphate buffer, pH 8.3. The large bathochromic shift of the near-UV band indicates that the N(3) position is deprotonated at pH 8.3. A pH titration revealed that the pK of the N(3) is shifted from 10.3 in free flavin to 6.4 in glycolate oxidase. Glycolate oxidase is competitively inhibited by oxalate with a Kd of 0.24 mM at 4 degrees C in 0.1 M phosphate buffer, pH 8.3. Three pieces of evidence demonstrate that glycolate oxidase stabilizes a negative charge at the N(1)-C(2 = O) locus: the enzyme forms a tight sulfite complex with a Kd of 2.7 x 10(-7) M and stabilizes the anionic flavosemiquinone and the benzoquinoid form of 8-mercapto-FMN. Steady-state analysis at pH 8.3, 4 degrees C, yielded a Km = 1 x 10(-3) M for glycolate and Km = 2.1 x 10(-4) M for oxygen. The turnover number has been determined to be 20 s-1. Stopped-flow studies of the reductive (k = 25 s-1) and oxidative (k = 8.5 x 10(4) M-1 s-1) half-reactions have identified the reduction of glycolate oxidase to be the rate-limiting step.  相似文献   

18.
The Mg2+-dependent ATPase (adenosine 5'-triphosphatase) mechanism of myosin and subfragment 1 prepared from frog leg muscle was investigated by transient kinetic technique. The results show that in general terms the mechanism is similar to that of the rabbit skeletal-muscle myosin ATPase. During subfragment-1 ATPase activity at 0-5 degrees C pH 7.0 and I0.15, the predominant component of the steady-state intermediate is a subfragment-1-products complex (E.ADP.Pi). Binary subfragment-1-ATP (E.ATP) and subfragment-1-ADP (E.ADP) complexes are the other main components of the steady-state intermediate, the relative concentrations of the three components E.ATP, E.ADP.Pi and E.ADP being 5.5:92.5:2.0 respectively. The frog myosin ATPase mechanism is distinguished from that of the rabbit at 0-5 degrees C by the low steady-state concentrations of E.ATP and E.ADP relative to that of E.ADP.Pi and can be described by: E + ATP k' + 1 in equilibrium k' - 1 E.ATP k' + 2 in equilibrium k' - 2 E.ADP.Pi k' + 3 in equilibrium k' - 3 E.ADP + Pi k' + 4 in equilibrium k' - 4 E + ADP. In the above conditions successive forward rate constants have values: k' + 1, 1.1 X 10(5)M-1.S-1; k' + 2 greater than 5s-1; k' + 3, 0.011 s-1; k' + 4, 0.5 s-1; k'-1 is probably less than 0.006s-1. The observed second-order rate constants of the association of actin to subfragment 1 and of ATP-induced dissociation of the actin-subfragment-1 complex are 5.5 X 10(4) M-1.S-1 and 7.4 X 10(5) M-1.S-1 respectively at 2-5 degrees C and pH 7.0. The physiological implications of these results are discussed.  相似文献   

19.
Human Hageman factor, a plasma proteinase zymogen, was activated in vitro under a near physiological condition (pH 7.8, ionic strength I = 0.14, 37 degrees C) by Pseudomonas aeruginosa elastase, which is a zinc-dependent tissue destructive neutral proteinase. This activation was completely inhibited by a specific inhibitor of the elastase, HONHCOCH(CH2C6H5)CO-Ala-Gly-NH2, at a concentration as low as 10 microM. In this activation Hagemen factor was cleaved, in a limited fashion, liberating two fragments with apparent molecular masses of 40 and 30 kDa, respectively. The appearance of the latter seemed to correspond chronologically to the generation of activated Hageman factor. Kinetic parameters of the enzymatic activation were kcat = 5.8 x 10(-3) s-1, Km = 4.3 x 10(-7) M and kcat/Km = 1.4 x 10(4) M-1 x s-1. This Km value is close to the plasma concentration of Hageman factor. Another zinc-dependent proteinase, P. aeruginosa alkaline proteinase, showed a negligible Hageman factor activation. In the presence of a negatively charged soluble substance, dextran sulfate (0.3-3 micrograms/ml), the activation rate by the elastase increased several fold, with the kinetic parameters of kcat = 13.9 x 10(-3) s-1, Km = 1.6 x 10(-7) M and kcat/Km = 8.5 x 10(4) M-1 x s-1. These results suggested a participation of the Hageman factor-dependent system in the inflammatory response to pseudomonal infections, due to the initiation of the system by the bacterial elastase.  相似文献   

20.
Q Su  J P Klinman 《Biochemistry》1999,38(26):8572-8581
Glucose oxidase catalyzes the oxidation of glucose by molecular dioxygen, forming gluconolactone and hydrogen peroxide. A series of probes have been applied to investigate the activation of dioxygen in the oxidative half-reaction, including pH dependence, viscosity effects, 18O isotope effects, and solvent isotope effects on the kinetic parameter Vmax/Km(O2). The pH profile of Vmax/Km(O2) exhibits a pKa of 7.9 +/- 0.1, with the protonated enzyme form more reactive by 2 orders of magnitude. The effect of viscosogen on Vmax/Km(O2) reveals the surprising fact that the faster reaction at low pH (1.6 x 10(6) M-1 s-1) is actually less diffusion-controlled than the slow reaction at high pH (1.4 x 10(4) M-1 s-1); dioxygen reduction is almost fully diffusion-controlled at pH 9.8, while the extent of diffusion control decreases to 88% at pH 9.0 and 32% at pH 5.0, suggesting a transition of the first irreversible step from dioxygen binding at high pH to a later step at low pH. The puzzle is resolved by 18O isotope effects. 18(Vmax/Km) has been determined to be 1.028 +/- 0.002 at pH 5.0 and 1.027 +/- 0.001 at pH 9.0, indicating that a significant O-O bond order decrease accompanies the steps from dioxygen binding up to the first irreversible step at either pH. The results at high pH lead to an unequivocal mechanism; the rate-limiting step in Vmax/Km(O2) for the deprotonated enzyme is the first electron transfer from the reduced flavin to dioxygen, and this step accompanies binding of molecular dioxygen to the active site. In combination with the published structural data, a model is presented in which a protonated active site histidine at low pH accelerates the second-order rate constant for one electron transfer to dioxygen through electrostatic stabilization of the superoxide anion intermediate. Consistent with the proposed mechanisms for both high and low pH, solvent isotope effects indicate that proton transfer steps occur after the rate-limiting step(s). Kinetic simulations show that the model that is presented, although apparently in conflict with previous models for glucose oxidase, is in good agreement with previously published kinetic data for glucose oxidase. A role for electrostatic stabilization of the superoxide anion intermediate, as a general catalytic strategy in dioxygen-utilizing enzymes, is discussed.  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