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.
Myeloperoxidase plays a fundamental role in oxidant production by neutrophils. The enzyme uses hydrogen peroxide to oxidize chloride (Cl-), bromide (Br-), iodide (I-), and the pseudohalide thiocyanate (SCN-) to their respective hypohalous acids. This study for the first time presents transient kinetic measurements of the oxidation of these halides and thiocyanate by the myeloperoxidase intermediate compound I, using the sequential mixing stopped-flow technique. At pH 7 and 15 degrees C, the two-electron reduction of compound I to the native enzyme by Cl- has a second-order rate constant of (2.5 +/- 0.3) x 10(4) M(-1) s(-1), whereas reduction of compound I by SCN- has a second-order rate constant of (9.6 +/- 0.5) x 10(6) M(-1) s(-1). Iodide [(7.2 +/- 0.7) x 10(6) M(-1) s(-1)] is shown to be a better electron donor for compound I than Br- [(1.1 +/- 0.1) x 10(6) M(-1) s(-1)]. The pH dependence studies suggest that compound I reduction by (pseudo-)halides is controlled by a residue with a pKa of about 4.6. The protonation of this group is necessary for optimum (pseudo-)halide anion oxidation. These transient kinetic results are underlined by steady-state spectral and kinetic investigations. SCN- is shown to be most effective in shifting the system myeloperoxidase/hydrogen peroxide from the peroxidatic cycle to the halogenation cycle, whereas iodide is shown to be more effective than bromide which in turn is much more effective than chloride. Decreasing pH increases the rate of this transition. Our results show that thiocyanate is an important substrate of myeloperoxidase in most environments and that hypothiocyanate is likely to contribute to leukocyte antimicrobial activity.  相似文献   

2.
A new ruthenium-labeled cytochrome c derivative was designed to measure the actual rate of electron transfer to the Trp-191 radical cation and the oxyferryl heme in cytochrome c peroxidase compound I {CMPI(FeIV = O,R.+)}. The H39C,C102T variant of yeast iso-1-cytochrome c was labeled at the single cysteine residue with a tris (bipyridyl)ruthenium(II) reagent to form Ru-39-Cc. This derivative has the same reactivity with CMPI as native yCc measured by stopped-flow spectroscopy, indicating that the ruthenium group does not interfere with the interaction between the two proteins. Laser excitation of the 1:1 Ru-39-Cc-CMPI complex in low ionic strength buffer (2 mM phosphate, pH 7) resulted in electron transfer from RuII* to heme c FeIII with a rate constant of 5 x 10(5) s-1, followed by electron transfer from heme c Fe II to the Trp-191 indolyl radical cation in CMPI(FeIV = O,R*+) with a rate constant of k(eta) = 2 x 10(6) s-1. A subsequent laser flash led to electron transfer from heme c to the oxyferryl heme in CMPII-(FeIV = O,R) with a rate constant of k(etb) = 5000 s-1. The location of the binding domain was determined using a series of surface charge mutants of CcP. The mutations D34N, E290N, and A193F each decreased the values of k(eta) and k(etb) by 2-4-fold, consistent with the use of the binding domain identified in the crystal structure of the yCc-CcP complex for reduction of both redox centers [Pelletier, H., & Kraut, J. (1992) Science 258, 1748-1755]. A mechanism is proposed for reduction of the oxyferryl heme in which internal electron transfer in CMPII(FeIV = O,R) leads to the regeneration of the radical cation in CMPII-(FeIII,R*+), which is then reduced by yCcII. Thus, both steps in the complete reduction of CMPI involve electron transfer from yCcII to the Trp-191 radical cation using the same binding site and pathway. Comparison of the rate constant k(eta) with theoretical predictions indicate that the electron transfer pathway identified in the crystalline yCc-CcP complex is very efficient. Stopped-flow studies indicate that native yCcII initially reduces the Trp-191 radical cation in CMPI with a second-order rate constant ka, which increases from 1.8 x 10(8) M-1 s-1 at 310 mM ionic strength to > 3 x 10(9) M-1 s-1 at ionic strengths below 100 mM. A second molecule of yCcII then reduces the oxyferryl heme in CMPII with a second-order rate constant kb which increases from 2.7 x 10(7) M-1 s-1 at 310 mM ionic strength to 2.5 x 10(8) M-1 s-1 at 160 mM ionic strength. As the ionic strength is decreased below 100 mM the rate constant for reduction of the oxyferryl heme becomes progressively slower as the reaction is limited by release of the product yCcIII from the yCcIII-CMPII complex. Both ruthenium photoreduction studies and stopped-flow studies demonstrate that the Trp-191 radical cation is the initial site of reduction in CMPI under all conditions of ionic strength.  相似文献   

3.
Cellular defense against excessive peroxynitrite generation is required to protect against DNA strand-breaks and mutations and against interference with protein tyrosine-based signaling and other protein functions due to formation of 3-nitrotyrosine. We recently demonstrated a role of selenium-containing enzymes catalyzing peroxynitrite reduction. Glutathione peroxidase (GPx) protected against the oxidation of dihydrorhodamine 123 (DHR) by peroxynitrite more effectively than ebselen (2-phenyl-1,2-benzisoselenazol-3(2H)-one), a selenoorganic compound exhibiting a high second-order rate constant for the reaction with peroxynitrite, 2 x 10(6) M-1s-1. The maintenance of protection by GPx against peroxynitrite requires GSH as reductant. Similarly, selenomethionine but not selenomethionine oxide exhibited inhibition of rhodamine 123 formation from DHR caused by peroxynitrite. In steady-state experiments, in which peroxynitrite was infused to maintain a 0.2 microM concentration, GPx in the presence of GSH, but neither GPx nor GSH alone, effectively inhibited the hydroxylation of benzoate by peroxynitrite. Under these steady-state conditions peroxynitrite did not cause loss of 'classical' GPx activity. GPx, like selenomethionine, protected against protein 3-nitrotyrosine formation in human fibroblast lysates, shown in Western blots. The formation of nitrite rather than nitrate from peroxynitrite was enhanced by GPx, ebselen or selenomethionine. The selenoxides can be effectively reduced by glutathione, establishing a biological line of defense against peroxynitrite. The novel function of GPx as a peroxynitrite reductase may extend to other selenoproteins containing selenocysteine or selenomethionine. Recent work on organotellurium compounds revealed peroxynitrite reductase activity as well. Inhibition of dihydrorhodamine 123 oxidation correlated well with the GPx-like activity of a variety of diaryl tellurides.  相似文献   

4.
Regulated assembly of myosin II in Dictyostelium discoideum amoebae partially controls the orderly formation of contractile structures during cytokinesis and cell migration. Kinetic and structural analyses show that Dictyostelium myosin II assembles by a sequential process of slow nucleation and controlled growth that differs in rate and mechanism from other conventional myosins. Nuclei form by an ordered progression from myosin monomers to parallel dimers to 0.43 microns long antiparallel tetramers. Lateral addition of dimers to bipolar tetramers completes the assembly of short (0.45 microns) blunt-ended thick filaments. Myosin heads are not staggered along the length of tapered thick filaments as in skeletal muscle, nor are bipolar minifilaments formed as in Acanthamoeba. The overall assembly reaction incorporating both nucleation and growth could be kinetically characterized by a second-order rate constant (kobs,N+G) of 1.85 x 10(4) M-1 s-1. Individual rate constants obtained for nucleation, kobs,N = 4.5 x 10(3) M-1 s-1, and growth, kobs,G = 2.5 x 10(4) M-1 s-1, showed Dictyostelium myosin II to be the slowest assembling myosin analyzed to date. Nucleation and growth stages were independently regulated by Mg2+, K+, and actin filaments. Increasing concentrations of K+ from 50 to 150 mM specifically inhibited lateral growth of dimers off nuclei. Intracellular concentrations of Mg2+ (1 mM) accelerated nucleation but maintained distinct nucleation and growth phase kinetics. Networks of actin filaments also accelerated the nucleation stage of assembly, mechanistically accounting for spontaneous formation of actomyosin contractile fibers via myosin assembly (Mahajan et al., 1989). The distinct assembly mechanism and regulation utilized by Dictyostelium myosin II demonstrates that myosins from smooth muscle, striated muscle, and two types of amoebae form unique thick filaments by different pathways.  相似文献   

5.
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   

6.
7.
Peroxidation of substrates such as ascorbic acid, pyrogallol, or ferulic acid, as well as indole acetic acid oxidation catalyzed by wheat germ peroxidase (WGP)2 C2, were found to be activated by Ca2+. This activation is independent of the stabilizing effect of structural Ca2+ reported for peroxidases. Steady state kinetics of ferulic acid oxidation catalyzed by WGP C2 showed an increase in the rate of compound I formation and of compound II decomposition in the presence of the ion, evidenced as an increase in rate constants k1, from 8.9 x 10(5) to 4.5 x 10(5) M-1 cm-1, and k3, from 4.4 x 10(5) to 1.1 x 10(6) M-1 cm-1. The dissociation constant Kd, for the cyanide derivative of the enzyme showed a marked decrease from 220 to 34 microM in the presence of Ca2+, thus implying an effect of the ion in the H2O2 binding step. In the presence of Ca2+, a conformational change in the protein was revealed by tryptophan fluorescence, providing a basis for the activation mechanism. Other peroxidases such as horseradish peroxidase and WGP C3 were not activated by Ca2+. The results suggest the existence of a physiological mechanism of control of peroxidase isozymes activity mediated by Ca2+.  相似文献   

8.
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   

9.
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   

10.
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   

11.
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   

12.
The antioxidant properties of the antithrombotic drug dipyridamole have been studied using lipid oxidation assays based on the generation of peroxy radicals by azo compounds. Dipyridamole was observed to prevent both peroxidation of arachidonic acid micelles in aqueous solution and peroxidation of methyl linoleate in organic solvents; in contrast to vitamin E, dipyridamole was found to scavenge both hydrophilic and hydrophobic radicals. The rate constant for the reaction of dipyridamole with methyl linoleate peroxyl radicals at 37 degrees C was calculated as 2 x 10(6) M-1s-1, in comparison to 1 x 10(6) M-1s-1 of vitamin E under the same conditions. The antioxidant efficiency of the drug was confirmed in experiments with radiolysis-induced oxidation and through measurements of malondialdehyde production and diene formation. As a result of radical scavenging, a relatively stable dipyridamole radical was formed that could be detected by electron spin resonance spectroscopy. The particular antioxidant properties of dipyridamole may explain the vasodilating and antiplatelet effects of this cardiovascular drug.  相似文献   

13.
The oxidized intermediates generated upon exposure of bovine liver catalase to hydrogen peroxide (H2O2) and superoxide radical (O2-) fluxes were examined with UV-visible spectrophotometry. H2O2 and O2- were generated by means of glucose/glucose oxidase and xanthine/xanthine oxidase systems. Serial overlay of absorption spectra in the Soret (350-450 nm) and visible (450-700 nm) regions showed that three oxidized intermediates, namely Compounds I, II and III, can be observed upon exposure of catalase to enzymatically generated H2O2 and O2-. Compound I is formed during the reaction of native enzyme with H2O2 and disappears in two ways: (i) via the catalytic reaction with H2O2 to restore native catalase and (ii) via the reaction with O2- to form Compound II. At low H2O2 concentrations (< 4.8 x 10(-9) M H2O2), Compound II reverts towards the native state mainly in a direct one-step reaction, whereas at higher H2O2 concentrations the pathway of Compound II back to the native enzyme involves Compound III. Formation of the latter from Compound II and H2O2 is irreversible and the rate constant of this reaction is 6.1 +/- 0.2 x 10(4) M-1 s-1. The formation of Compound III through the direct reaction of O2- with native enzyme has also been observed. Depending on the experimental conditions, the inactivation of catalase by O2- can be due to accumulation of Compound II ("slow" inhibition) or to the formation of Compound III ("rapid" inhibition) part of which leads to a dead end product. Formation of Compound III and of this dead end product are responsible for the irreversible inactivation in presence of an excess of H2O2.  相似文献   

14.
Kinetics of photomodification of 26-meric deoxyribonucleotide pTTGCCTTGAATGGGAA-GAGGGTCATT with derivatives of the complementary oligonucleotides pTCTTCCCATTC, pTCTTCCCA, and pTTCCCA bearing a residue of (p-azidotetrafluorobenzoyl)aminopropylamine(-ArN3) attached to the terminal phosphate (reagents I, II, and III, respectively) was studied at 37 degrees C. It was established that during irradiation the reagents are inactivated, loosing their affinity to the target. A kinetic equation describing the modification was suggested. From the dependence of the time-limited modification level on the reagent concentration, the association constants of the reagents with the target were determined: [Kx = (9.9 +/- 0.4) x 10(4), (1.1 +/- 0.1) x 10(5), and (8.4 +/- 2.1) x 10(6) M-1 for reagents I, II, and III, respectively] and the efficiency of the modification in the complex gamma ef (ca. 0.3 for all the reagents) were determined. From the dependence of the modification level [PZ]/p0 on time for reagent II, the rate constant was determined for the rate-determining step of the photomodification k0 = (7.9 +/- 0.9) x 10(-3) s-1, which is close to the rate constant for the photolysis of p-azidotetrafluorobenzoic acid kp = (5.5 +/- 0.3) x 10(-3) s-1.  相似文献   

15.
The final step of the catalytic cycle of cytochrome oxidase, the reduction of oxyferryl heme a3 in compound F, was investigated using a binuclear polypyridine ruthenium complex (Ru2C) as a photoactive reducing agent. The net charge of +4 on Ru2C allows it to bind electrostatically near CuA in subunit II of cytochrome oxidase. Photoexcitation of Ru2C with a laser flash results in formation of a metal-to-ligand charge-transfer excited state, Ru2C, which rapidly transfers an electron to CuA of cytochrome oxidase from either beef heart or Rhodobacter sphaeroides. This is followed by reversible electron transfer from CuA to heme a with forward and reverse rate constants of k1 = 9.3 x 10(4) s-1 and k-1 = 1.7 x 10(4) s-1 for R. sphaeroides cytochrome oxidase in the resting state. Compound F was prepared by treating the resting enzyme with excess hydrogen peroxide. The value of the rate constant k1 is the same in compound F where heme a3 is in the oxyferryl form as in the resting enzyme where heme a3 is ferric. Reduction of heme a in compound F is followed by electron transfer from heme a to oxyferryl heme a3 with a rate constant of 700 s-1, as indicated by transients at 605 and 580 nm. No delay between heme a reoxidation and oxyferryl heme a3 reduction is observed, showing that no electron-transfer intermediates, such as reduced CuB, accumulate in this process. The rate constant for electron transfer from heme a to oxyferryl heme a3 was measured in beef cytochrome oxidase from pH 7.0 to pH 9.5, and found to decrease upon titration of a group with a pKa of 9.0. The rate constant is slower in D2O than in H2O by a factor of 4.3, indicating that the electron-transfer reaction is rate-limited by a proton-transfer step. The pH dependence and deuterium isotope effect for reduction of isolated compound F are comparable to that observed during reaction of the reduced, CO-inhibited CcO with oxygen by the flow-flash technique. This result indicates that electron transfer from heme a to oxyferryl heme a3 is not controlled by conformational effects imposed by the initial redox state of the enzyme. The rate constant for electron transfer from heme a to oxyferryl heme a3 is the same in the R. sphaeroides K362M CcO mutant as in wild-type CcO, indicating that the K-channel is not involved in proton uptake during reduction of compound F.  相似文献   

16.
A ruthenium-labeled cytochrome c derivative was prepared to meet two design criteria: the ruthenium group must transfer an electron rapidly to the heme group, but not alter the interaction with cytochrome c oxidase. Site-directed mutagenesis was used to replace His39 on the backside of yeast C102T iso-1-cytochrome c with a cysteine residue, and the single sulfhydryl group was labeled with (4-bromomethyl-4' methylbipyridine) (bis-bipyridine)ruthenium(II) to form Ru-39-cytochrome c (cyt c). There is an efficient pathway for electron transfer from the ruthenium group to the heme group of Ru-39-cyt c comprising 13 covalent bonds and one hydrogen bond. Electron transfer from the excited state Ru(II*) to ferric heme c occurred with a rate constant of (6.0 +/- 2.0) x 10(5) s-1, followed by electron transfer from ferrous heme c to Ru(III) with a rate constant of (1.0 +/- 0.2) x 10(6) s-1. Laser excitation of a complex between Ru-39-cyt c and beef cytochrome c oxidase in low ionic strength buffer (5 mM phosphate, pH7) resulted in electron transfer from photoreduced heme c to CuA with a rate constant of (6 +/- 2) x 10(4) s-1, followed by electron transfer from CuA to heme a with a rate constant of (1.8 +/- 0.3) x 10(4) s-1. Increasing the ionic strength to 100 mM leads to bimolecular kinetics as the complex is dissociated. The second-order rate constant is (2.5 +/- 0.4) x 10(7) M-1s-1 at 230 mM ionic strength, nearly the same as that of wild-type iso-1-cytochrome c.  相似文献   

17.
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   

18.
The reactions of NO. with O2.- and with HO2. were studied using the pulse radiolysis technique under pseudo first order conditions where ([O2.-]o + [HO2.]o) > [NO.]o at pH 3.3-10.0. The rate constant of the reaction of NO. with O2.- was determined both by monitoring the decay of O2.- at 250 nm and the formation of ONOO- at 302 nm to be (4.3 +/- 0.5) x 10(9) M-1s-1, independent of ionic strength and pH in the range of 6.1-10.0. The rate constant of the reaction of NO. with HO2.- was determined by following the decay of HO2. at 250 nm to be (3.2 +/- 0.3) x 10(9) M-1s-1 at pH 3.3.  相似文献   

19.
Kinetics of peroxynitrite reaction with amino acids and human serum albumin   总被引:1,自引:0,他引:1  
An initial rate approach was used to study the reaction of peroxynitrite with human serum albumin (HSA) through stopped-flow spectrophotometry. At pH 7.4 and 37 degreesC, the second order rate constant for peroxynitrite reaction with HSA was 9.7 +/- 1.1 x 10(3) M-1 s-1. The rate constants for sulfhydryl-blocked HSA and for the single sulfhydryl were 5.9 +/- 0.3 and 3.8 +/- 0.8 x 10(3) M-1 s-1, respectively. The corresponding values for bovine serum albumin were also determined. The reactivity of sulfhydryl-blocked HSA increased at acidic pH, whereas plots of the rate constant with the sulfhydryl versus pH were bell-shaped. The kinetics of peroxynitrite reaction with all free L-amino acids were determined under pseudo-first order conditions. The most reactive amino acids were cysteine, methionine, and tryptophan. Histidine, leucine, and phenylalanine (and by extension tyrosine) did not affect peroxynitrite decay rate, whereas for the remaining amino acids plots of kobs versus concentration were hyperbolic. The sum of the contributions of the constituent amino acids of the protein to HSA reactivity was comparable to the experimentally determined rate constant, where cysteine and methionine (seven residues in 585) accounted for an estimated 65% of the reactivity. Nitration of aromatic amino acids occurred in HSA following peroxynitrite reaction, with nitration of sulfhydryl-blocked HSA 2-fold higher than native HSA. Carbon dioxide accelerated peroxynitrite decomposition, enhanced aromatic amino acid nitration, and partially inhibited sulfhydryl oxidation of HSA. Nitration in the presence of carbon dioxide increased when the sulfhydryl was blocked. Thus, cysteine 34 was a preferential target of peroxynitrite both in the presence and in the absence of carbon dioxide.  相似文献   

20.
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   

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