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.
Flash photolysis of alkaline peroxynitrite solutions results in the formation of nitrogen monoxide and superoxide. From the rate of recombination it is concluded that the rate constant of the reaction of nitrogen monoxide with superoxide is (1.9 +/- 0.2) x 10(10) M-1 s-1. The pKa of hydrogen oxoperoxonitrate is dependent on the medium. With the stopped-flow technique a value of 6.5 is found at millimolar phosphate concentrations, while at 0.5 M phosphate the value is 7.5. The kinetics of decay do not follow first-order kinetics when the pH is larger than the pKa, combined with a total peroxynitrite and peroxynitrous acid concentration that exceeds 0.1 mM. An adduct between ONOO- and ONOOH is formed with a stability constant of (1.0 +/- 0.1) x 10(4) M. The kinetics of the decay of hydrogen oxoperoxonitrate are not very pressure-dependent: from stopped-flow experiments up to 152 MPa, an activation volume of 1.7 +/- 1.0 cm3 mol-1 was calculated. This small value is not compatible with homolysis of the O-O bond to yield free nitrogen dioxide and the hydroxyl radical. Pulse radiolysis of alkaline peroxynitrite solutions indicates that the hydroxyl radical reacts with ONOO- to form [(HO)ONOO].- with a rate constant of 5.8 x 10(9) M-1 s-1. This radical absorbs with a maximum at 420 nm (epsilon = 1.8 x 10(3) M-1 cm-1) and decays by second-order kinetics, k = 3.4 x 10(6) M-1 s-1. Improvements to the biomimetic synthesis of peroxynitrite with solid potassium superoxide and gaseous nitrogen monoxide result in higher peroxynitrite to nitrite yields than in most other syntheses.  相似文献   

2.
The kinetic and equilibrium CO binding properties of hemoglobin (Hb) M Iwate (alpha2 87 His leads to Tyr beta2) have been investigated. The results show that the alpha(Met) beta2 (CO) tetramer of this protein has a low affinity for CO, as indicated by the stopped-flow and flash-photolysis kinetic, as well as the CO binding equilibrium, measurements. However, it has been found that the phosphate-free alpha2(Met)beta2(CO) tetramer does tend to dimerize extensively (K4.2 = 55 muM). The high-affinity forms seen in earlier kinetic measurements may be explained by this fact. When dimers are accounted for in the functional studies, the results show that the tetramer binds CO noncooperatively either with or without the allosteric cofactor, inositol hexaphosphate (IHP). IHP appears to influence the functional properties of a solution of Hb M Iwate by stabilizing the tetrameric state of aggregation, thereby greatly reducing the population of high-affinity dimers. When the CO "off" rate with IHP present (0.23 s-1) and the CO "on" rate to the tetramer either with or without IHP (1.9 X 10(5) M-1 s-1) at 25 degrees C are sued to calculate the equilibrium constant, the value obtained (8.3 X 10(5) M-1 s-1) is similar to that in equilibrium binding measurements on the phosphate-free tetramer (9.5 X 10(5) M-1) estimated from the observed P 1/2 value at 0.48 mM total heme concentration. By showing that dimers account for the high-affinity component seen in earlier kinetic experiments with Hg M Iwate, we can now more strongly suggest that cooperative CO binding to this tetramer is minimal or absent, with both of the active beta-hemes presenting a very low affinity.  相似文献   

3.
The binding kinetics of methyl alpha- and methyl beta-D-galactopyranoside to the anti-T lectin from peanuts were studied by 13C NMR, employing methyl galactopyranosides specifically enriched in 13C at C-1. Association and dissociation rate constants, as well as their activation parameters, are reported. The association rate constants, 4.6 X 10(4) M-1 s-1 for the alpha-galactopyranoside and 3.6 X 10(4) M-1 s-1 for the beta-galactopyranoside, are several orders of magnitude below those expected for a diffusion-controlled process. For both anomers, the association rate constant was temperature independent, implying that the association process occurs without a significant activation enthalpy. However, a considerable association activation entropy was found for both ligands. The dissociation rate constants were in the range of 9-46 s-1 within a temperature range of 5-35 degrees C for the alpha-galactopyranoside, and in the range of 9-39 s-1 within a temperature range of 5-25 degrees C for the beta-galactopyranoside. A considerable dissociation activation enthalpy of ca. 10 kcal mol-1 was found for both anomers. A two-step binding model, consistent with the present NMR data and with previous UV and CD spectroscopic data, is presented.  相似文献   

4.
Site-directed mutagenesis has been used to produce variants of a tryptic fragment of bovine liver cytochrome b5 in which Glu44 and Glu56 are mutated to alanine. The reduction potentials measured by spectroelectrochemical titration (in the presence of 1 mM (Ru(NH3)6)3+, pH 7.0 and I=0.1 M) are 4.5, 6.0, 6.0 and 7.5 mV versus the standard hydrogen electrode (SHE) for the wild-type and E44A, E56A and E44/56A mutants of cytochrome b5, respectively. A comparative two-dimensional NMR study of cytochrome b5 and its E44/56A mutant in water solution has been achieved. Resonance assignments of side-chains have been completed successfully. The NMR results suggest that the secondary structures and global folding of the E44/56A mutant remain unchanged, but the mutation of both Glu44 and Glu56 to hydrophobic alanine may lead to the two helices containing mutated residues contracting towards the heme center. The inner mobility of the Gly42 approximately Glu44 segment in cytochrome b5 may be responsible for the difference of the binding mode between Glu44 and Glu56 with cytochrome c. The binding between cytochrome c and cytochrome b5 was studied by optical difference spectra of cytochrome c and variants of cytochrome b5. The association constants (KA) for the wild-type, E44A, E56A, and E44/56A mutants of cytochrome b5 with cytochrome c, are 4.70(+/-0. 10)x10(6) M-1, 1.88(+/-0.03)x10(6) M-1, 2.70(+/-0.13)x10(6) M-1, and 1.14(+/-0.05)x10(6) M-1, respectively. This is indicative that both Glu44 and Glu56 are involved in the complex formation between cytochrome b5 and cytochrome c. The reduction of horse heart ferricytochrome c by recombinant ferrocytochrome b5 and its mutants has been studied. The rate constant of the electron transfer reaction between ferricytochrome c and wild-type ferrocytochrome b5 (1.074(+/-0.49)x10(7) M-1 s-1) is higher than those of the mutant protein E44A (8.98(+/-0.20)x10(6) M-1 s-1), E56A (8.76(+/-0. 39)x10(6) M-1 s-1), and E44/56A (8.02(+/-0.38)x10(6) M-1 s-1) at 15 degreesC, pH 7.0, I=0.35 M. The rate constants are strongly dependent on ionic strength and temperature. These studies, by means of a series of techniques, provide conclusive results that the interaction between cytochrome b5 and cytochrome c is electrostatically guided, and, more importantly, that both Glu44 and Glu56 participate in the electron transfer reaction.  相似文献   

5.
We have prepared a series of biotinylated analogs of omega-conotoxin (omega CgTx) as potent, selective markers for N-type calcium channels. At pH 9.5, reaction of omega CgTx with amidocaproylbiotin succinimidyl ester gives three biotinylated conjugates, labeled at lysines 2 or 24, or at both positions. Kinetic competition assays of 125I-omega CgTx binding to rat brain synaptic membranes show that each conjugate has a similar rate constant for association (1-1.3 x 10(6) M-1 s-1) but not dissociation (1-4 x 10(-4) s-1). Comparison with rate constants obtained for the association (1.2 x 10(7) M-1 s-1) and dissociation (5 x 10(-5) s-1) of native omega CgTx indicates that while biotinylation reduces omega CgTx potency (Kdkin = k-2/k2 = 4 pM for omega CgTx), binding of these labels to membranes is nevertheless of very high affinity (Kdkin 0.1-0.3 nM).  相似文献   

6.
Random mutagenesis and screening for enzymatic activity has been used to engineer horse heart myoglobin to enhance its intrinsic peroxidase activity. A chemically synthesized gene encoding horse heart myoglobin was subjected to successive cycles of PCR random mutagenesis. The mutated myoglobin gene was expressed in Escherichia coli LE392, and the variants were screened for peroxidase activity with a plate assay. Four cycles of mutagenesis and screening produced a series of single, double, triple, and quadruple variants with enhanced peroxidase activity. Steady-state kinetics analysis demonstrated that the quadruple variant T39I/K45D/F46L/I107F exhibits peroxidase activity significantly greater than that of the wild-type protein with k1 (for H2O2 oxidation of metmyoglobin) of 1. 34 x 10(4) M-1 s-1 ( approximately 25-fold that of wild-type myoglobin) and k3 [for reducing the substrate (2, 2'-azino-di-(3-ethyl)benzthiazoline-6-sulfonic acid] of 1.4 x 10(6) M-1 s-1 (1.6-fold that of wild-type myoglobin). Thermal stability of these variants as measured with circular dichroism spectroscopy demonstrated that the Tm of the quadruple variant is decreased only slightly compared with wild-type (74.1 degreesC vs. 76.5 degreesC). The rate constants for binding of dioxygen exhibited by the quadruple variant are identical to the those observed for wild-type myoglobin (kon, 22.2 x 10(-6) M-1 s-1 vs. 22.3 x 10(-6) M-1 s-1; koff, 24.3 s-1 vs. 24.2 s-1; KO2, 0.91 x 10(-6) M-1 vs. 0.92 x 10(-6) M-1). The affinity of the quadruple variant for CO is increased slightly (kon, 0.90 x 10(-6) M-1s-1 vs. 0.51 x 10(-6) M-1s-1; koff, 5.08 s-1 vs. 3.51 s-1; KCO, 1.77 x 10(-7) M-1 vs. 1.45 x 10(-7) M-1). All four substitutions are in the heme pocket and within 5 A of the heme group.  相似文献   

7.
The kinetics of formation and dissociation of the horse metmyoglobin/fluoride complex has been investigated between pH 3.4 and 11. The ionic strength dependence of the reaction has been measured at integral pH values between pH 5 and 10. Hydrofluoric acid, HF, binds to metmyoglobin with a rate constant of (4.7 +/- 0. 7) x 10(4) M-1 s-1. An apparent ionization in metmyoglobin with a pKa of 4.4 +/- 0.5 influences the rate of HF binding and is attributed to the distal histidine, His-64. Protonation of His-64 increases the HF binding rate by a factor of 2.6. The fluoride anion, F-, binds to metmyoglobin with a rate constant of (5.6 +/- 1.4) x 10(-2) M-1 s-1, about 10(6) times slower than HF. Binding of either HF or F- to hydroxymetmyoglobin cannot be detected. Protonation of the distal histidine facilitates HF dissociation from the metmyoglobin/fluoride complex. HF dissociates with a rate constant of 1.9 +/- 0.3 s-1. The fluoride anion dissociates 2000 times more slowly, with a rate constant of (8.7 +/- 1.6) x 10(-4) s-1. The apparent pKa for His-64 ionization in the fluorometmyoglobin complex is 5.7 +/- 0.1. The association and dissociation rate constants are relatively independent of ionic strength with secondary kinetic salt effects sufficient to account for the ionic strength variation of both, consistent with the idea that association and dissociation of neutral HF dominate the kinetics of fluoride binding to metmyoglobin.  相似文献   

8.
The effect of free fatty acids (FFA) and non-enzymatic glycation on the binding kinetics of dansylsarcosine (DS) to human serum albumin (HSA) was studied using the stopped-flow technique. The influence of FFA on the binding parameters of 25% glycated HSA depended on the type of fatty acid. The addition of stearic, oleic and linoleic acids in a concentration of 0.3 mmol/l showed no inhibitory effects on the association rate constant (k2) value for DS binding to 25% glycated HSA (k2 without FFA: 385 +/- 10 s-1, k2 with FFA > or = 385 +/- 10 s-1). In contrast, shorter chain fatty acids (hexanoic, octanoic, decanoic, lauric and myristic acids) showed marked inhibitory effects for 0.3 mmol/l FFA (k2 range: 233 +/- 32 to 69 +/- 5 s-1) and for 0.6 mmol/l FFA (k2 range: 125 +/- 3 to 20 +/- 4 s-1). The association rate constant (k2) as well as the affinity constant (KA) of DS were markedly affected by glycation: k2 was 686 +/- 61 s-1 for 7% glycated HSA, 385 +/- 10 s-1 for 25% glycated HSA and 209 +/- 12 s-1 for 50% glycated HSA. KA decreased from 6.1 +/- 2.9 x 10(5) M-1 for 7% glycated HSA, to 5.1 +/- 0.1 x 10(5) M-1 for 25% glycated HSA and to 1.3 +/- 0.6 x 10(5) M-1 for 50% glycated HSA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Earlier investigations of the oxidation of 3,5,3',5'-tetramethylbenzidine (TMB) using horseradish peroxidase and prostaglandin H-synthase have shown the formation of a cation free radical of TMB in equilibrium with a charge-transfer complex, consistent with either a two- or a one-electron initial oxidation. In this work, we exploited the distinct spectroscopic properties of myeloperoxidase and its oxidized intermediates, compounds I and II, to establish two successive one-electron oxidations of TMB. By employing stopped-flow techniques under transient-state and steady-state conditions, we also determined the rate constants for the elementary steps of the myeloperoxidase-catalyzed oxidation of TMB at pH 5.4 and 20 degrees C. The second-order rate constant for compound I formation from the reaction of native enzyme with H2O2 is 2.6 x 10(7) M-1 s-1. Compound I undergoes a one-electron reduction to compound II in the presence of TMB, and the rate constant for this reaction was determined to be (3.6 +/- 0.1) x 10(6) M-1 s-1. The spectral scans show that compound II accumulates in the steady state. The rate constant for compound II reduction to native enzyme by TMB obtained under steady-state conditions is (9.4 +/- 0.6) x 10(5) M-1 s-1. The results are applied to a new, more accurate assay for myeloperoxidase based upon the formation of the charge-transfer complex between TMB and its diimine final product.  相似文献   

10.
The disordered N- (1-73) and C- (74-108) fragments of oxidized Escherichia colithioredoxin (Trx) reconstitute the native structure upon association [Tasayco, M. L., & Chao, K. (1995) Proteins: Struct., Funct., Genet. 22, 41-44]. Kinetic measurements of the formation of the complex (1-73/74-108) at 20 degrees C under apparent pseudo-first-order conditions using stopped-flow far-UV CD and fluorescence spectroscopies indicate association coupled to folding, an apparent rate constant of association [kon = (1330 +/- 54) M-1 s-1], and two apparent unimolecular rate constants [k1 = (0. 037 +/- 0.007) s-1 and k2 = (0.0020 +/- 0.0005) s-1]. The refolding kinetics of the GuHCl denatured Trx shows the same two slowest rate constants. An excess of N- over C-fragment decreases the kon, and the slowest phase disappears when a P76A variant is used. Stopped-flow fluorescence measurements at 20 degrees C indicate a GuHCl-dependent biphasic dissociation/unfolding process of the complex, where the slowest phase corresponds to 90% of the total. Their rate constants, extrapolated to zero denaturant, k-1 = (9 +/- 3) x 10(-5) s-1 and k-2 = (3.4 +/- 1.2) x 10(-5) s-1, show m# values of (4.0 +/- 0.4) kcal mol-1 M-1 and (3.5 +/- 0.1) kcal mol-1 M-1, respectively. Our results indicate that: (i) a compact intermediate with trans P76 and defined tertiary structure seems to participate in both the folding and unfolding processes; (ii) not all the N-fragment is competent to associate with the C-fragment; (iii) conversion to an association competent form occurs apparently on the time scale of P76 isomerization; and (iv) the P76A variation does not alter the association competency of the C-fragment, but it permits its association with "noncompetent" forms of the N-fragment.  相似文献   

11.
It is thought that only a limited number of residues in the extended binding pocket of coagulation proteases are critical for substrate and inhibitor specificity. A candidate residue from the crystal structures of thrombin and factor Xa (FXa) that may be critical for specificity at the S2 subsite is residue 99. Residue 99 is Tyr in FXa and Thr in activated protein C (APC). To determine the role of residue 99 in S2 specificity, a Gla-domainless mutant of protein C (GDPC) was prepared in which Thr99 was replaced with Tyr of FXa. GDPC T99Y bound Ca2+ and was activated by the thrombin-thrombomodulin complex normally. The T99Y mutant, similar to FXa, hydrolyzed the chromogenic substrates with a Gly at the P2 positions. This mutant was also inhibited by antithrombin (AT) (k2 = 4.2 +/- 0.2 x 10(1) M-1 s-1), and heparin accelerated the reaction >350-fold (k2 = 1.5 +/- 0.1 x 10(4) M-1 s-1). The T99Y mutant, however, did not activate prothrombin but inactivated factor Va approximately 2-fold better than wild type. To try to switch the specificity of FXa, both Tyr99 and Gln192 of FXa were replaced with those of APC in the Gla-domainless factor X (GDFX Y99T/Q192E). This mutant was folded correctly as it bound Ca2+ with a similar affinity as GDFX and was also activated by the Russell's viper venom at similar rate, but it cleaved the chromogenic substrates with a Gly at the P2 positions poorly. The mutant, instead, cleaved the APC-specific chromogenic substrates efficiently. The Y99T/Q192E mutant became resistant to inhibition by AT in the absence of heparin but was inhibited by AT almost normally in the presence of heparin (k2 = 3.4 +/- 0.5 x 10(5) M-1 s-1). The Y99T/Q192E mutant did not inactivate factor Va, and prothrombin activation by this mutant was impaired. These results indicate that 1) residue 99 is critical for enzyme specificity at the S2 subsite, 2) a role for heparin in acceleration of FXa inhibition by AT may involve the S2-P2 modulation, and 3) the exchange of residues 99 and 192 in FXa and APC may switch the enzyme specificity with the chromogenic substrates and inhibitors but not with the natural substrates.  相似文献   

12.
The electrogenic Na/K pump current (Ip) was studied in the dissociated neostriatal neurons of the rat by using the nystatin-perforated patch recording mode. The Ip was activated by external K+ in a concentration-dependent manner with an EC50 of 0.7 mM at a holding potential (VH) of -40 mV. Other monovalent cations also caused Ip and the order of potency was Tl+>K+, Rb+>NH4+, Cs+>Li+. The Ip decreased with membrane hyperpolarization in an external solution containing 150 mM Na+, while the Ip did not show such voltage dependency without external Na+. Ouabain showed a steady-state inhibition of Ip in a concentration- and temperature-dependent manner at a VH of -40 mV. The IC50 values at 20 and 30 degrees C were 7.1 x 10(-6) and 1.3 x 10(-6) M, respectively. The decay of Ip after adding ouabain well fitted with a single exponential function. At a VH of -40 Mv, the association (k+1) and dissociation (k-1) rate constants estimated from the time constant of the current decay at 20 degrees C were 4.0 x10(2) s-1 M-1 and 6.3 x 10(-3) s-1, respectively. At 30 degrees C, k+1 increased to 2.8 x 10(3) s-1 M-1 while k-1 showed no such change with a value of 1.8 x 10(-3) s-1. A continuous Na+ influx was demonstrated by both the Na+-dependent leakage current and tetrodotoxin-sensitive Na+ current, which resulted in the continuous activation of the Na/K pump. It was thus concluded that the Na/K pump activity was well-maintained in the dissociated rat neostriatal neurons with distinct functional properties and that the activity of the pump was tightly connected with Na+ influxes.  相似文献   

13.
Glutathione peroxidases and their mimics, e.g., ebselen or diaryl tellurides, efficiently reduce peroxynitrite/peroxynitrous acid (ONOO-/ONOOH) to nitrite and protect against oxidation and nitration reactions. Here, we report the second-order rate constant for the reaction of the reduced form of glutathione peroxidase (GPx) with peroxynitrite as (8.0 +/- 0.8) x 10(6) M-1 s-1 (per GPx tetramer) at pH 7.4 and 25 degreesC. The rate constant for oxidized GPx is about 10 times lower, (0.7 +/- 0.2) x 10(6) M-1 s-1. On a selenium basis, the rate constant for reduced GPx is similar to that obtained previously for ebselen. The data support the conclusion that GPx can exhibit a biological function by acting as a peroxynitrite reductase.  相似文献   

14.
Enantiomers of 4-nitrophenyl 4-X-phenacyl methylphosphonate esters (X = H, PMN; CH3; and CH3O) inactivate human alpha-thrombin with rate constants 4-235 M-1 s-1 in pH 6.5, 0.025 M citrate buffer, and 0.15 M NaCl at 7.0 +/- 0.1 degrees C. Stereoselectivity of the inactivation of thrombin is 2-39 and favors the levorotatory enantiomers. The pH-dependence of inactivation of thrombin by (-)-PMN is sigmoidal and consistent with the participation of a catalytic residue with a pKa of 8.0 +/- 0.1 in 0.15 M NaCl and a pKa of 7.4 +/- 0.2 in 0.15 M choline chloride in the nucleophilic attack of the catalytic Ser at phosphorus. The solvent isotope effect on ki/Ki in the pH-independent region of the reaction is 2.26 +/- 0.17. Thrombin activity returns from the adducts on the 2-7 h time scale at 25.0 +/- 0.1 degrees C via a self-catalyzed intramolecular reaction. The pH dependence of reactivation is significant from the adduct formed with (-)-CH3O-PMN and (-)-CH3-PMN and less so from the adducts formed with the other enantiomers of the inhibitors. Kinetic pKs approximately 7.2, with the exception of the adducts with (-)-PMN and (-)-CH3O-PMN, indicate that a pH-dependent conformational change affects the rate of dephosphonylation. A structural interpretation of the stereoselectivity and other mechanistic features is provided based on the energy-optimized structures of the adducts. Pharmaco-medical use of human alpha-thrombin covalently modified by the PMNs is suggested.  相似文献   

15.
Fe(III) uptake by the iron-delivery and iron-scavenging protein, hen ovotransferrin has been investigated in vitro between pH 6.5 and 9. In the absence of any ferric chelate, apo-ovotransferrin loses two protons with K1a = 50 +/- 1 nM and K2a = 4.0 +/- 0.1 nM. These acid-base equilibria are independent of the interaction of the protein with bicarbonate. The interaction with bicarbonate occurs with two different affinity constants, KC = 9.95 +/- 0.15 mM and KN = 110 +/- 10 mM. FeNAc3 exchanges its Fe(III) with the C-site of the protein in interaction with bicarbonate, direct rate constants k1 = 650 +/- 25 M-1 s-1, reverse rate constant k-1 = (6.0 +/- 0.1) x 10(3) M-1 s-1 and equilibrium constant K1 = 0.11 +/- 0.01. This iron-protein intermediate loses then a single proton, K3a = 3.50 +/- 0.35 nM, and undergoes a first change in conformation followed by a two or three proton loss, first order rate constant k2 = 0.30 +/- 0.01 s-1. This induces a new modification in conformation followed by the loss of one or two protons, first order rate constant k3 = (1.50 +/- 0.05) x 10(-2) s-1. These modifications in the monoferric protein conformation are essential for iron uptake by the N-site of the protein. In the last step, the monoferric and diferric proteins attain their final state of equilibrium in about 15,000 s. The overall mechanism of iron uptake by ovotransferrin is similar but not identical to those of serum transferrin and lactoferrin. The rates involved are, however, closer to lactoferrin than serum transferrin, whereas the affinities for Fe(III) are lower than those of serum transferrin and lactoferrin. Does this imply that the metabolic function transferrins is more related to kinetics than to thermodynamics?  相似文献   

16.
2-Hydroxy-6-oxo-6-phenylhexa-2,4-dienoate (6-phenyl-HODA) hydrolase (BphD), an enzyme of the biphenyl biodegradation pathway encoded by the bphD gene of Burkholderia cepacia LB400, was hyperexpressed and purified to apparent homogeneity. SDS-polyacrylamide gel electrophoresis confirmed that BphD has a subunit molecular mass of 32 kDa, while gel filtration demonstrated that it is a homotetramer of molecular weight 122,000. The enzyme hydrolyzed 6-phenyl-HODA with a kcat of 5.0 (+/- 0.07) s-1 and a kcat/Km of 2.0 (+/- 0.08) x 10(7) M-1 s-1 (100 mM phosphate, pH 7.5, 25 degreesC). The specificity of BphD for other 2-hydroxy-6-oxohexa-2,4-dienoates (HODAs) decreased markedly with the size of the C6 substituent; 6-methyl-HODA, the meta cleavage product of 3-methylcatechol, was hydrolyzed approximately 2300 times less specifically than 6-phenyl-HODA. By comparison, the homologous hydrolase from the toluene degradation pathway, TodF, showed highest specificity for 6-methyl- and 6-ethyl-HODA (kcat/Km of 2.0 (+/- 0.05) x 10(6) M-1 s-1 and 9.0 (+/- 0.5) x 10(6) M-1 s-1, respectively). TodF showed no detectable activity toward 6-phenyl-HODA and 6-tert-butyl-HODA. Neither BphD nor TodF hydrolyzed 5-methyl-HODA efficiently. The kcat of BphD determined by monitoring product formation was about half that determined by monitoring substrate disappearance, suggesting that some uncoupling of substrate utilization and product formation occurs during the enzyme catalyzed reaction. Crystals of BphD were obtained using ammonium sulfate combined with polyethylene glycol 400 as the precipitant. Diffraction was observed to a resolution of at least 1.9 A, and the evaluation of self-rotation functions confirmed 222 (D2) molecular symmetry.  相似文献   

17.
Plasminogen activator inhibitor-1 (PAI-1) is the primary inhibitor of the plasminogen activators (PAs), tissue-type plasminogen activator (tPA), and urokinase-type plasminogen activator (uPA). A library of PAI-1 mutants containing substitutions at the P1 and P1' positions was screened for functional activity against tPA and thrombin. Several PAI-1 variants that were inactive against uPA in a previous study (Sherman, P. M., Lawrence, D. A., Yang, A. Y., Vandenberg, E. T., Paielli, D., Olson, S. T., Shore, J. D., and Ginsburg, D. (1992) J. Biol. Chem. 267, 7588-7595) had significant inhibitory activity toward tPA. This set of tPA-specific PAI-1 mutants contained a wide range of amino acid substitutions at P1 including Asn, Gln, His, Ser, Thr, Leu, Met, and all the aromatic amino acids. This group of mutants also demonstrated a spectrum of substitutions at P1'. Kinetic analyses of selected variants identified P1Tyr and P1His as the most efficient tPA-specific inhibitors, with second-order rate constants (ki) of 4.0 x 10(5) M-1s-1 and 3.6 x 10(5) M-1s-1, respectively. Additional PA-specific PAI-1 variants containing substitutions at P3 through P1' were constructed. P3Tyr-P2Ser-P1Lys-P1'Trp and P3Tyr-P2Ser-P1Tyr-P1'Met had ki values of 1.7 x 10(6) M-1s-1 and 2.5 x 10(6) M-1s-1 against tPA, respectively, but both were inactive against uPA. In contrast, P2Arg-P1Lys-P1'Ala inhibited uPA 74-fold more rapidly than tPA. The mutant PAI-1 library was also screened for inhibitory activity toward thrombin in the presence and absence of the cofactor heparin. While wild-type PAI-1 and several P1Arg variants inhibited thrombin in the absence of heparin, a number of variants were thrombin inhibitors only in the presence of heparin. These results demonstrate the importance of the reactive center residues in determining PAI-1 target specificity and suggest that second sites of interaction between inhibitors and proteases can also contribute to target specificity. Finally, the PA-specific mutants described here should provide novel reagents for dissecting the physiological role of PAI-1 both in vitro and in vivo.  相似文献   

18.
Quantitative characterization of the interaction of des-kringle1-5-plasmin (microplasmin) with fibrin(ogen) and plasma protease inhibitors may serve as a tool for further evaluation of the role of kringle domains in the regulation of fibrinolysis. Comparison of fibrin(ogen) degradation products yielded by plasmin, miniplasmin (des-kringle1-4-plasmin), microplasmin, and trypsin on SDS gel electrophoresis indicates that the differences in the enzyme structure result in different rates of product formation, whereas the products of the four proteases are very similar in molecular weight. Kinetic parameters show that plasmin is the most efficient enzyme in fibrinogen degradation, and the kcat/KM ratio decreases in parallel with the loss of the kringle domains. The catalytic sites of the four proteases have similar affinities for fibrin (KM values between 0.12 and 0.21 microM). Trypsin has the highest catalytic constant for fibrin digestion (kcat = 0.47 s-1), and among plasmins with different kringle structures, the loss of kringle5 results in a markedly lower catalytic rate constant (kcat = 0.0076 s-1 for microplasmin vs 0.048 s-1 for miniplasmin and 0.064 s-1 for plasmin). In addition, microplasmin is inactivated by plasmin inhibitor (k" = 3.9 x 10(5) M-1 s-1) and antithrombin (k" = 1.4 x 10(3) M-1 s-1) and the rate of inactivation decreases in the presence of fibrin(ogen). Heparin (250 nM) accelerates the inactivation of microplasmin by antithrombin (k" = 10.5 x 10(3) M-1 s-1 ), whereas that by plasmin inhibitor is not affected (k" = 4.2 x 10(5) M-1 s-1).  相似文献   

19.
The kinetics of saccharide binding to the treatment form of concanavalin A have been studies at pH 7.2 with the temperature-jump method. 4-Methylumbelliferyl alpha-D-mannopyranoside was used as a ligand; its fluorescence is totally quenched upon binding. A single relaxation of ligand fluorescence (tau = 20-400 ms) was observed and was investigated at three different temperatures, using kinetic titration and dilution types of experiments. The concentration dependence of the relaxation time and amplitude was consistent with a single-step bimolecular association and independent binding sites. In the temperature range 13-24 degrees C the association and dissociation rate parameters are in the range (6-10) X 10(4) M-1 s-1 and (1.4-3.2)s-1 respectively, corresponding to activation energies for the forward and reverse reactions equal to approx. 13 and 8 kcal/mol (54 and 33 kJ/mol) respectively. Two additional relaxations of protein fluorescence (3 ms and larger than 1 s at 25 degrees C) were unaffected by carbohydrate binding. Tetrameric concanavalin A shows carbohydrate binding parameters that are almost identical to those of native or derivatized dimeric concanavalin A.  相似文献   

20.
The kinetic properties of the ba3 oxidase from Thermus thermophilus were investigated by stopped-flow spectroscopy in the temperature range of 5-70 degrees C. Peculiar behavior in the reaction with physiological substrates and classical ligands (CO and CN-) was observed. In the O2 reaction, the decay of the F intermediate is significantly slower (k' = 100 s-1 at 5 degrees C) than in the mitochondrial enzyme, with an activation energy E of 10.1 +/- 0.9 kcal mol-1. The cyanide-inhibited ba3 oxidizes cyt c522 quickly (k approximately 5 x 10(6) M-1 s-1 at 25 degrees C) and selectively, with an activation energy E of 10.9 +/- 0.9 kcal mol-1, but slowly oxidizes ruthenium hexamine, a fast electron donor for the mitochondrial enzyme. Cyt c552 oxidase activity is enhanced up to 60 degrees C and is maximal at extremely low ionic strengths, excluding formation of a high-affinity cyt c522-ba3 electrostatic complex. The thermophilic oxidase is less sensitive to cyanide inhibition, although cyanide binding under turnover is much quicker (seconds) than in the fully oxidized state (days). Finally, the affinity of reduced ba3 for CO at 20 degrees C (Keq = 1 x 10(5) M-1) was found to be smaller than that of beef heart aa3 (Keq = 4 x 10(6) M-1), partly because of an unusually fast, strongly temperature-dependent CO dissociation from cyt a32+ of ba3 (k' = 0.8 s-1 vs k' = 0.02 s-1 for beef heart aa3 at 20 degrees C). The relevance of these results to adaptation of respiratory activity to high temperatures and low environmental O2 tensions 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