首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The kinetics and thermodynamics of Bi(III) exchange between bismuth mononitrilotriacetate (BiL) and human serum transferrin as well as those of the interaction between bismuth-loaded transferrin and transferrin receptor 1 (TFR) were investigated at pH 7.4-8.9. Bismuth is rapidly exchanged between BiL and the C-site of human serum apotransferrin in interaction with bicarbonate to yield an intermediate complex with an effective equilibrium constant K(1) of 6 +/- 4, a direct second-order rate constant k(1) of (2.45 +/- 0.20) x 10(5) M(-1) s(-1), and a reverse second-order rate constant k(-1) of (1.5 +/- 0.5) x 10(6) M(-1) s(-1). The intermediate complex loses a single proton with a proton dissociation constant K(1a) of 2.4 +/- 1 nM to yield a first kinetic product. This product then undergoes a modification in its conformation followed by two proton losses with a first-order rate constant k(2) = 25 +/- 1.5 s(-1) to produce a second kinetic intermediate, which in turn undergoes a last modification in the conformation to yield the bismuth-saturated transferrin in its final state. This last process rate-controls Bi(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau(3)(-1) of (3 +/- 1) x 10(-2) s(-1). The mechanism of bismuth uptake differs from that of iron and probably does not involve the same transition in conformation from open to closed upon iron uptake. The interaction of bismuth-loaded transferrin with TFR occurs in a single very fast kinetic step with a dissociation constant K(d) of 4 +/- 0.4 microM, a second-order rate constant k(d) of (2.2 +/- 1.5) x 10(8) M(-1) s(-1), and a first-order rate constant k(-d) of 900 +/- 400 s(-1). This mechanism is different from that observed with the ferric holotransferrin and implies that the interaction between TFR and bismuth-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of bismuth incorporation by the transferrin receptor-mediated iron acquisition pathway is discussed.  相似文献   

2.
The bis(terpyridine)cobalt(II), Co(terpy)2(2+), reduction of cytochrome c peroxidase compound I, CcP-I, has been investigated using stopped-flow techniques as a function of ionic strength in pH 7.5 buffers at 25 degrees C. Co(terpy)2(2+) initially reduces the Trp191 radical site in CcP-I with an apparent second-order rate constant, k2, equal to 6.0+/-0.4x10(6) M(-1)s(-1) at 0.01 M ionic strength. A pseudo-first-order rate constant of 480 s(-1) was observed for the reduction of CcP-I by 79 microM Co(terpy)2(2+) at 0.01 M ionic strength. The one-electron reduction of CcP-I produces a second enzyme intermediate, CcP compound II (CcP-II), which contains an oxyferryl, Fe(IV), heme. Reduction of the Fe(IV) heme in CcP-II by Co(terpy)2(2+) shows saturation kinetics with a maximum observed rate constant, k3max, of 24+/-2 s(-1) at 0.01 M ionic strength. At low reductant concentrations, the apparent second-order rate constant for Co(terpy)2(2+) reduction of CcP-II, k3, is 1.2+/-0.5x10(6) M(-1) s-1. All three rate constants decrease with increasing ionic strength. At 0.10 M ionic strength, values of k2, k3, and k3max decrease to 6.0+/-0.8x10(5) M(-1) s(-1), 1.2+/-0.5x10(5) M(-1) s(-1), and 11+/-3 s(-1), respectively. Both the product, Co(terpy)2(3+), and ferricytochrome c inhibit the rate of Co(terpy)2(2+) reduction of CcP-I and CcP-II. Gel-filtration studies show that a minimum of two Co(terpy)2(3+) molecules bind to the native enzyme in low ionic strength buffers.  相似文献   

3.
The reductive decomposition of both SNAP and SNOCap by ascorbate in aqueous solution (in the presence of EDTA) was thoroughly investigated. Nitric oxide (NO) release from the reaction occurs in an ascorbate concentration and pH dependent manner. Rates and hence NO release increased drastically with increasing pH, signifying that the most highly ionized form of ascorbate is the more reactive species. The experiments were monitored spectrophotometrically, and second-order rate constants calculated at 37 degrees C for the reduction of SNAP are k(b)=9.81+/-1.39 x 10(-3) M(-1) s(-1) and k(c)=662+/-38 M(-1) s(-1) and for SNOCap are k(b)=2.57+/-1.29 x 10(-2) M(-1) s(-1) and k(c)=49.7+/-1.3 M(-1) s(-1). k(b) and k(c) are the second-order rate constants via the ascorbate monoanion (HA-) and dianion (A2-) pathways, respectively. Activation parameters were also calculated and are DeltaHb++ =93+/-7 kJ mol(-1), DeltaSb++ =15+/-2 J K(-1) mol(-1) and DeltaHc++ =51+/-5 kJ mol(-1), DeltaSc++ =-28+/-3 J K(-1) mol(-1) with respect to the reactions involving SNAP. Those for the reaction between SNOCap and ascorbate were calculated to be DeltaHb++ =63+/-11 kJ mol(-1), DeltaSb++ =-71+/-20 J K(-1) mol(-1) and DeltaHc++ =103+/-7 kJ mol(-1), DeltaSc++ =118+/-8 J K(-1) mol(-1). The effect of Cu2+/Cu+ ions on the reductive decompositions of these S-nitrosothiols was also investigated in absence of EDTA. SNOCap exhibits relatively high stability at near physiological conditions (37 degrees C and pH 7.55) even in the presence of micromolar concentrations of Cu2+, with decomposition rate constant being 0.011 M(-1) s(-1) in comparison to SNAP which is known to be more susceptible to catalytic decomposition by Cu2+ (second-order rate constant of 20 M(-1) s(-1) at pH 7.4 and 25 degrees C). It was also observed that the reductive decomposition of SNAP is not catalyzed by alkali metal ions, however, there was an increase in rate as the ionic strength increases from 0.2 to 0.5 mol dm(-3) NaCl.  相似文献   

4.
An important aspect of the catalytic mechanism of microsomal glutathione transferase (MGST1) is the activation of the thiol of bound glutathione (GSH). GSH binding to MGST1 as measured by thiolate anion formation, proton release, and Meisenheimer complex formation is a slow process that can be described by a rapid binding step (K(GSH)d = 47 +/- 7 mM) of the peptide followed by slow deprotonation (k2 = 0.42 +/- 0.03 s(-1). Release of the GSH thiolate anion is very slow (apparent first-order rate k(-2) = 0.0006 +/- 0.00002 s(-)(1)) and thus explains the overall tight binding of GSH. It has been known for some time that the turnover (kcat) of MGST1 does not correlate well with the chemical reactivity of the electrophilic substrate. The steady-state kinetic parameters determined for GSH and 1-chloro-2,4-dinitrobenzene (CDNB) are consistent with thiolate anion formation (k2) being largely rate-determining in enzyme turnover (kcat = 0.26 +/- 0.07 s(-1). Thus, the chemical step of thiolate addition is not rate-limiting and can be studied as a burst of product formation on reaction of halo-nitroarene electrophiles with the E.GS- complex. The saturation behavior of the concentration dependence of the product burst with CDNB indicates that the reaction occurs in a two-step process that is characterized by rapid equilibrium binding ( = 0.53 +/- 0.08 mM) to the E.GS- complex and a relatively fast chemical reaction with the thiolate (k3 = 500 +/- 40 s(-1). In a series of substrate analogues, it is observed that log k3 is linearly related (rho value 3.5 +/- 0.3) to second substrate reactivity as described by Hammett sigma- values demonstrating a strong dependence on chemical reactivity that is similar to the nonenzymatic reaction (rho = 3.4). Microsomal glutathione transferase 1 displays the unusual property of being activated by sulfhydryl reagents. When the enzyme is activated by N-ethylmaleimide, the rate of thiolate anion formation is greatly enhanced, demonstrating for the first time the specific step that is activated. This result explains earlier observations that the enzyme is activated only with more reactive substrates. Taken together, the observations show that the kinetic mechanism of MGST1 can be described by slow GSH binding/thiolate formation followed by a chemical step that depends on the reactivity of the electrophilic substrate. As the chemical reactivity of the electrophile becomes lower the rate-determining step shifts from thiolate formation to the chemical reaction.  相似文献   

5.
The Reaction Rates of NO with Horseradish Peroxidase Compounds I and II   总被引:1,自引:0,他引:1  
In this study the reactions between nitric oxide (NO) and horseradish peroxidase (HRP) compounds I and II were investigated. The reaction between compound I and NO has biphasic kinetics with a clearly dominant initial fast phase and an apparent second-order rate constant of (7.0 +/- 0.3) x 10(5) M(-1) s(-1) for the fast phase. The reaction of compound II and NO was found to have an apparent second-order rate constant of k(app) = (1.3 +/- 0.1) x 10(6) M(-1) s(-1) or (7.4 +/- 0.7) x 10(5) M(-1) s(-1) when measured at 409 nm (the isosbestic point between HRP and HRP-NO) and 419 nm (lambda(max) of compound II and HRP-NO), respectively. Interestingly, the reaction of compound II with NO is unusually high relative to that of compound I, which is usually the much faster reaction. Since horseradish peroxidase is prototypical of mammalian peroxidases with respect to the oxidation of small substrates, these results may have important implications regarding the lifetime and biochemistry of NO in vivo after inflammation where both NO and H(2)O(2) generation are increased several fold.  相似文献   

6.
The intermolecular electron transfer kinetics between nitrite reductase (NiR, cytochrome cd1) isolated from Pseudomonas nautica and three cytochromes c isolated from the same strain, as well as the intramolecular electron transfer between NiR heme c and NiR heme d1, were investigated by cyclic voltammetry. All cytochromes (cytochrome c552, cytochrome c553 and cytochrome C553(548)) exhibited well-behaved electrochemistry. The individual diffusion coefficients and mid-point redox potentials were determined. Under the experimental conditions, only cytochrome c552 established a rapid electron transfer with NiR. At acidic pH, the intermolecular electron transfer (cytochrome c(552red)-->NiR heme cox) is a second-order reaction with a rate constant (k2) of 4.1+/-0.1x10(5) M(-1) s(-1) (pH=6.3 and 100 mM NaCl). Under these conditions, the intermolecular reaction represents the rate-limiting step. A minimum estimate of 33 s(-1) could be determined for the first-order rate constant (k1) of the intramolecular electron transfer reaction NiR heme c(red)-->NiR heme d1ox. The pH dependence of k2 values was investigated at pH values ranging from 5.8 to 8.0. When the pH is progressively shifted towards basic values, the rate constant of the intramolecular electron transfer reaction NiR heme c(red)-->NiR heme d1ox decreases gradually to a point where it becomes rate limiting. At pH 8.0 we determined a value of 1.4+/-0.7 s(-1), corresponding to a k2 value of 2.2+/-1.1x10(4) M(-1) s(-1) for the intermolecular step. The physiological relevance of these results is discussed with a particular emphasis on the proposed mechanism of "dead-end product" formation.  相似文献   

7.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

8.
The reactions of lactoperoxidase (LPO) intermediates compound I, compound II and compound III, with nitrite (NO2(-)) were investigated. Reduction of compound I by NO2(-) was rapid (k2 = 2.3 x 10(7) M(-1) x s(-1); pH = 7.2) and compound II was not an intermediate, indicating that NO2* radicals are not produced when NO2(-) reacts with compound I. The second-order rate constant for the reaction of compound II with NO2(-) at pH = 7.2 was 3.5 x 10(5) M(-1) x s(-1). The reaction of compound III with NO2(-) exhibited saturation behaviour when the observed pseudo first-order rate constants were plotted against NO2(-) concentrations and could be quantitatively explained by the formation of a 1 : 1 ratio compound III/NO2(-) complex. The Km of compound III for NO2(-) was 1.7 x 10(-4) M and the first-order decay constant of the compound III/ NO2(-) complex was 12.5 +/- 0.6 s(-1). The second-order rate constant for the reaction of the complex with NO2(-) was 3.3 x 10(3) M(-1) x s(-1). Rate enhancement by NO2(-) does not require NO2* as a redox intermediate. NO2(-) accelerates the overall rate of catalysis by reducing compound II to the ferric state. With increasing levels of H2O2, there is an increased tendency for the catalytically dead-end intermediate compound III to form. Under these conditions, the 'rescue' reaction of NO2(-) with compound III to form compound II will maintain the peroxidatic cycle of the enzyme.  相似文献   

9.
Rapid kinetic, spectroscopic, and potentiometric studies have been performed on human Theta class glutathione transferase T2-2 to dissect the mechanism of interaction of this enzyme with its natural substrate GSH. Theta class glutathione transferases are considered to be older than Alpha, Pi, and Mu classes in the evolutionary pathway. As in the more recently evolved GSTs, the activation of GSH in the human Theta enzyme proceeds by a forced deprotonation of the sulfhydryl group (pK(a) = 6.1). The thiol proton is released quantitatively in solution, but above pH 6.5, a protein residue acts as an internal base. Unlike Alpha, Mu, and Pi class isoenzymes, the GSH-binding mechanism occurs via a simple bimolecular reaction with k(on) and k(off) values at least hundred times lower (k(on) = (2.7 +/- 0.8) x 10(4) M(-1) s(-1), k(off) = 36 +/- 9 s(-1), at 37 degrees C). Replacement of Arg-107 by alanine, using site-directed mutagenesis, remarkably increases the pK(a) value of the bound GSH and modifies the substrate binding modality. Y107A mutant enzyme displays a mechanism and rate constants for GSH binding approaching those of Alpha, Mu, and Pi isoenzymes. Comparison of available crystallographic data for all these GSTs reveals an unexpected evolutionary trend in terms of flexibility, which provides a basis for understanding our experimental results.  相似文献   

10.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The ability of glutathione to scavenge the superoxide radical is a matter of serious contention in the literature: reported values for the second-order rate constant range from 10(2) to greater than 10(5) M(-1) s(-1). The physiological implications of this discrepancy will determine, for example, whether or not glutathione can compete with Mn-superoxide dismutase for reaction with the radical in the mitochondrial matrix, leading to formation of the potentially harmful glutathionyl radical. Several authors have investigated the kinetics of glutathione oxidation by superoxide using spectrophotometric assays, based on competition between either ferricytochrome c or epinephrine for reaction with the radical. However, these approaches have received criticism because the contributions of various secondary reactions to the overall kinetics have been largely overlooked (e.g., the reduction of ferricytochrome c by glutathione). In the present investigation, we have used electron paramagnetic resonance spectroscopy to monitor competition between GSH and the spin trap 5,5-dimethyl-1-pyrroline N-oxide for reaction with superoxide. This method has been used previously and a rate constant of 1.8 x 10(5) M(-1) s(-1) obtained (Dikalov, S.; Khramtsov, V.; Zimmer, G. Arch. Biochem. Biophys. 326:207-218; 1996). However, we demonstrate that this value is a gross overestimation because the spectrum of the hydroxyl radical adduct of the spin trap was incorrectly assigned to the glutathionyl radical adduct. The relatively high yield of the DMPO hydroxyl radical adduct is shown to be due to the two-electron reduction of the corresponding superoxide radical adduct by glutathione. Taking these factors into consideration, we estimate the second order rate constant for the oxidation of glutathione by superoxide to be approximately 200 M(-1) s(-1).  相似文献   

12.
Bifunctional catalase-peroxidases are the least understood type of peroxidases. A high-level expression in Escherichia coli of a fully active recombinant form of a catalase-peroxidase (KatG) from the cyanobacterium Anacystis nidulans (Synechococcus PCC 6301) is reported. Since both physical and kinetic characterization revealed its identity with the wild-type protein, the large quantities of recombinant KatG allowed the examination of both the spectral characteristics and the reactivity of its redox intermediates by using the multi-mixing stopped-flow technique. The homodimeric acidic protein (pI = 4.6) contained high catalase activity (apparent K(m) = 4.8 mM and apparent k(cat) = 8850 s(-1)). Cyanide is shown to be an effective inhibitor of the catalase reaction. The second-order rate constant for cyanide binding to the ferric protein is (6.9 +/- 0.2) x 10(5) M(-1 )s(-1) at pH 7.0 and 15 degrees C and the dissociation constant of the cyanide complex is 17 microM. Because of the overwhelming catalase activity, peroxoacetic acid has been used for compound I formation. The apparent second-order rate constant for formation of compound I from the ferric enzyme and peroxoacetic acid is (1.3 +/- 0.3) x 10(4 )M(-1 )s(-1) at pH 7.0 and 15 degrees C. The spectrum of compound I is characterized by about 40% hypochromicity, a Soret region at 406 nm, and isosbestic points between the native enzyme and compound I at 355 and 428 nm. Rate constants for reduction of KatG compound I by o-dianisidine, pyrogallol, aniline and isoniazid are shown to be (7.3 +/- 0.4) x 10(6) M(-1 )s(-1), (5.4 +/- 0.3) x 10(5) M(-1 )s(-1), (1.6 +/- 0.3) x 10(5) M(-1 )s(-1) and (4.3 +/- 0.2) x 10(4) M(-1 )s(-1), respectively. The redox intermediate formed upon reduction of compound I did not exhibit the classical red-shifted peroxidase compound II spectrum which characterizes the presence of a ferryl oxygen species. Its spectral features indicate that the single oxidizing equivalent in KatG compound II is contained on an amino acid which is not electronically coupled to the heme.  相似文献   

13.
The interaction of dimethylsulfoxide (Me2SO) with glutathione was investigated under non-equilibrium conditions in solution using 1H NMR and in intact erythrocytes using 1H spin-echo NMR. In solution the reaction was observed to follow second-order kinetics (Rate = k1[glutathione][Me2SO]) at 300 K pH 7.4, k(sol) = 4.7 x 10(-5)mol(-1)L(1)s(-1). In intact erythrocytes the rate constant for the cellular environment, k(cell), was found to be slightly larger at 8.1 x 10(-5)mol(-1)L(1)s(-1). Furthermore, the reaction of Me2SO with erythrocyte glutathione showed a biphasic dependence on the Me2SO concentration, with little oxidation of glutathione occurring until the Me2SO concentration exceeded 0.5 molL(-1). The results suggest that at lower concentrations, Me2SO can be effectively removed, most probably by reaction with glutathione, which is regenerated by glutathione reductase, although preferential reaction with other cellular components (e.g., membrane or cellular thiols) cannot be ruled out. Thus the concentrations of Me2SO that are commonly used in cryopreservation of mammalian cells ( approximately 1.4 molL(-1)) can cause oxidation of intracellular glutathione.  相似文献   

14.
The velocity of the oxidative renaturation of reduced ribonuclease A catalyzed by protein disulfide isomerase (PDI) is strongly dependent on the composition of a glutathione/glutathione disulfide redox buffer. As with the uncatalyzed, glutathione-mediated oxidative folding of ribonuclease, the steady-state velocity of the PDI-catalyzed reaction displays a distinct optimum with respect to both the glutathione (GSH) and glutathione disulfide (GSSG) concentrations. Optimum activity is observed at [GSH] = 1.0 mM and [GSSG] = 0.2 mM. The apparent kcat at saturating RNase concentration is 0.46 +/- 0.05 mumol of RNase renatured min-1 (mumol of PDI)-1 compared to the apparent first-order rate constant for the uncatalyzed reaction of 0.02 +/- 0.01 min-1. Changes in GSH and GSSG concentration have a similar effect on the rate of both the PDI-catalyzed and uncatalyzed reactions except under the more oxidizing conditions employed, where the catalytic effectiveness of PDI is diminished. The ratio of the velocity of the catalyzed reaction to that of the uncatalyzed reaction increases as the quantity [GSH]2/[GSSG] increases and approaches a constant, limiting value at [GSH]2/[GSSG] greater than 1 mM, suggesting that a reduced, dithiol form of PDI is required for optimum activity. As long as the glutathione redox buffer is sufficiently reducing to maintain PDI in an active form [( GSH]2/[GSSG] greater than 1 mM), the rate acceleration provided by PDI is reasonably constant, although the actual rate may vary by more than an order of magnitude. PDI exhibits half of the maximum rate acceleration at a [GSH]2/[GSSG] of 0.06 +/- 0.01 mM.  相似文献   

15.
Rate constants for the hydrolysis of L-proline methyl ester to form proline and methanol in D(2)O buffered at neutral pD and 25 degrees C and the deuterium enrichment of the proline product determined by electrospray ionization mass spectrometry are reported. The data give k(DO) = 5.3 +/- 0.5 M(-1) s(-1) as the second-order rate constant for carbon deprotonation of N-protonated proline methyl ester by deuterioxide ion in D(2)O at 25 degrees C and I = 1.0 (KCl). The data provide good estimates of carbon acidities of pK(a) = 21 for N-protonated proline methyl ester and pK(a) = 29 for proline zwitterion in water and of the second-order rate constant k(HO) = 4.5 x 10(-5) M(-1) s(-1) for carbon deprotonation of proline zwitterion by hydroxide ion at 25 degrees C. There is no detectable acceleration of the deprotonation of N-protonated proline methyl ester by the Br?nsted base 3-quinuclidinone in water, and it is not clear that such Br?nsted catalysis would make a significant contribution to the rate acceleration for deprotonation of bound proline at proline racemase. A comparison of the first-order rate constants k(HO)[HO(-)] = 4.5 x 10(-11) s(-1) for deprotonation of free proline zwitterion in water at pH 8 and k(cat) = 2600 s(-1) for deprotonation of proline bound to the active site of proline racemase at pH 8 shows that the enzymatic rate acceleration for proline racemase is ca. 10(13)-fold. This corresponds to a 19 kcal/mol stabilization of the transition state for deprotonation of the enzyme-bound carbon acid substrate by interaction with the protein catalyst. It is suggested that (1) much of the rate acceleration of the enzymatic over the nonenzymatic reaction in water may result from transfer of the substrate proline zwitterion from the polar solvent water to a nonpolar enzyme active site and (2) the use of thiol anions rather than oxygen anions as Br?nsted bases at this putative nonpolar enzyme active site may be favored, because of the smaller energetic price for desolvation of thiol anions than for desolvation of the more strongly solvated oxygen anions.  相似文献   

16.
The rate constant for the reduction of the tyrosyl radical with selenocysteine has been measured to investigate whether selenocysteine is capable of repair of protein radicals. Tyrosyl radicals, both free in solution and in insulin, were generated by means of pulse radiolysis and laser flash photolysis in aqueous solution. The rate constant for the reaction of free N-acetyl-tyrosyl-amine radicals with selenocysteine is (8 +/- 2) x 10 (8) M (-1) s (-1), and that for tyrosyl radicals in insulin is (1.6 +/- 0.4) x 10 (8) M (-1) s (-1). The rate constant for the reaction of selenoglutathione with the N-acetyl-tyrosyl-amine radical is (5 +/- 2) x 10 (8) M (-1) s (-1). In contrast, cysteine and glutathione react more slowly than their selenium analogues with the tyrosyl radical: the reactions of N-acetyl-tyrosyl-amine radicals with cysteine and glutathione are 3 and 5 orders of magnitude slower, respectively, than those with selenocysteine and selenoglutathione, while those of tyrosyl radicals in insulin are 3 and 2 orders of magnitude slower, respectively.  相似文献   

17.
The reaction between antiplasmin (A) and a low-molecular-weight form of plasmin (P) proceeds in at least two steps: a fast reversible second-order reaction followed by a slower irreversible first-order transition, and may be represented by: P +A k1 in equilibrium k-1 PA k2 leads to PA'. The low-Mr plasmin, which is obtained by limited elastase digestion, is composed of an intact B chain and a small A chain lacking the lysine-binding sites. The k1 of the reaction is (6.5 +/- 0.5) x 10(5) M-1 s-1 which is 30--60 times smaller than that for normal plasmin and antiplasmin. The dissociation constant of the first step is 1.9 x 10(-9) M which is 10 times higher than for normal plasmin and antiplasmin. The rate constant of the second step is (4.2 +/- 0.2) x 10(-3) s-1 for both normal and low-Mr plasmin. Low Mr plasmin which has substrate bound to its active site does not react or reacts only very slowly with antiplasmin. The reaction rate, however, is only slightly influenced by 6-aminohexanoic acid in concentrations up to 1 mM which decrease the reaction rate of normal plasmin approximately 50-fold. The findings further indicate that the lysine-binding site(s) of plasmin are of great importance for the rate of its reaction with antiplasmin.  相似文献   

18.
Myeloperoxidase (MPO) is the most abundant protein in neutrophils and plays a central role in microbial killing and inflammatory tissue damage. Because most of the non-steroidal anti-inflammatory drugs and other drugs contain a thiol group, it is necessary to understand how these substrates are oxidized by MPO. We have performed transient kinetic measurements to study the oxidation of 14 aliphatic and aromatic mono- and dithiols by the MPO intermediates, Compound I (k3) and Compound II (k4), using sequential mixing stopped-flow techniques. The one-electron reduction of Compound I by aromatic thiols (e.g. methimidazole, 2-mercaptopurine and 6-mercaptopurine) varied by less than a factor of seven (between 1.39 +/- 0.12 x 10(5) M(-1) s(-1) and 9.16 +/- 1.63 x 10(5) M(-1) s(-1)), whereas reduction by aliphatic thiols was demonstrated to depend on their overall net charge and hydrophobic character and not on the percentage of thiol deprotonation or redox potential. Cysteamine, cysteine methyl ester, cysteine ethyl ester and alpha-lipoic acid showed k3 values comparable to aromatic thiols, whereas a free carboxy group (e.g. cysteine, N-acetylcysteine, glutathione) diminished k3 dramatically. The one-electron reduction of Compound II was far more constrained by the nature of the substrate. Reduction by methimidazole, 2-mercaptopurine and 6-mercaptopurine showed second-order rate constants (k4) of 1.33 +/- 0.08 x 10(5) M(-1) s(-1), 5.25 +/- 0.07 x 10(5) M(-1) s(-1) and 3.03 +/- 0.07 x 10(3) M(-1) s(-1). Even at high concentrations cysteine, penicillamine and glutathione could not reduce Compound II, whereas cysteamine (4.27 +/- 0.05 x 10(3) M(-1) s(-1)), cysteine methyl ester (8.14 +/- 0.08 x 10(3) M(-1) s(-1)), cysteine ethyl ester (3.76 +/- 0.17 x 10(3) M(-1) s(-1)) and alpha-lipoic acid (4.78 +/- 0.07 x 10(4) M(-1) s(-1)) were demonstrated to reduce Compound II and thus could be expected to be oxidized by MPO without co-substrates.  相似文献   

19.
Ferrimyoglobin at pH 7.4 binds nitric oxide to yield nitric oxide adducts. In the presence of glutathione (GSH), nitrosoadducts of Mb(III) react with it to give nitrosoglutathione, whose concentration has been determined with an apparatus based on a specific and sensitive solid-state amperometric gas sensor. The reaction constant between the adduct and glutathione, kGSH = (47 +/- 1) M(-1) x s(-1), obtained by UV-Vis spectroscopy kinetic measurements, is about one-eighth of the constant with OH- determined by other authors. We can explain this fact with the higher nucleophilicity of OH- compared to GSH, due to the bulkiness and charge of the species. It is known that the formation of nitrosothiols starting from nitrite or NO (nitrogen monoxide) and glutathione, in the absence of oxygen, is impossible. Thus, from a biological point of view, it is important to point out that GSH reacts with NO in the presence of ferrimyoglobin, even at physiological pH, to form nitrosoglutathione.  相似文献   

20.
Degradation of the reduced pyridine nucleotides NMNH and NADH by HOCl involves two distinct stages: a fast reaction, k = 4.2 x 10(5) M(-1) s(-1), leads to generation of stable pyridine products (Py/Cl) with a strong absorption band at 275 nm (epsilon = 12.4 x 10(3) M(-1) cm(-1) in the case of NMNH); secondarily, a subsequent reaction of HOCl, k = 3.9 x 10(3) M(-1) s(-1), leads to a complete loss of the aromatic absorption band of the pyridine ring. HOBr and HOI(I(2)) react similarly. Apparent rate constants of the primary reactions of HOX species with NMNH at pH 7.2 increase in the order HOCl (3 x 10(5) M(-1) s(-1)) < HOBr( approximately 4 x 10(6) M(-1) s(-1)) < HOI(I(2))( approximately 6.5 x 10(7) M(-1) s(-1)). HOBr reacts fast also with the primary product Py/Br, k approximately 9 x 10(5) M(-1) s(-1), while the reactions of HOI and I(2) with Py/I are slower, approximately 1.4 x 10(3) M(-1) s(-1) and >6 x 10(3) M(-1) s(-1), respectively. Halogenation of the amide group of NMN(+) by HOX species is many orders of magnitude slower than oxidation of NMNH. Taurine inhibits HOCl-induced oxidation of NADH, but HOBr-induced oxidation is not inhibited because the taurine monobromamine rapidly oxidizes NADH, and oxidation by HOI(I(2)) is not inhibited because taurine is inert toward HOI(I(2)). Also sulfur compounds (GSH, GSSG, and methionine) are less efficient in protecting NADH against oxidation by HOBr and HOI(I(2)) than against oxidation by HOCl. The results suggest that reactions of HOBr and HOI(I(2)) in a cellular environment are much more selectively directed toward irreversible oxidation of NADH than reactions of HOCl. It is noteworthy that the rather inert N-chloramines react with iodide to generate HOI(I(2)), i.e., the most reactive and selective oxidant of reduced pyridine nucleotides. NMR investigations show that the primary stable products of the reaction between NMNH and HOCl are various isomeric chlorohydrins originating from a nonstereospecific electrophilic addition of HOCl to the C5&dbond;C6 double bond of the pyridine ring. The primary products (Py/X) of NMNH all exhibit similar absorption bands around 275 nm and are hence likely to result from analogous addition of HOX to the C5&dbond;C6 bond of the pyridine ring. Since the Py/X species are stable and inert toward endogeneous reductants like ascorbate and GSH, they may generally be useful markers for assessing the contribution of hypohalous acids to inflammatory injury.  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