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.
Lad L  Mewies M  Raven EL 《Biochemistry》2002,41(46):13774-13781
The catalytic mechanism of recombinant soybean cytosolic ascorbate peroxidase (rsAPX) and a derivative of rsAPX in which a cysteine residue (Cys32) located close to the substrate (L-ascorbic acid) binding site has been modified to preclude binding of ascorbate [Mandelman, D., Jamal, J., and Poulos, T. L. (1998) Biochemistry 37, 17610-17617] has been examined using pre-steady-state and steady-state kinetic techniques. Formation (k1 = 3.3 +/- 0.1 x 10(7) M(-1) s(-1)) of Compound I and reduction (k(2) = 5.2 +/- 0.3 x 10(6) M(-1) s(-1)) of Compound I by substrate are fast. Wavelength maxima for Compound I of rsAPX (lambda(max) (nm) = 409, 530, 569, 655) are consistent with a porphyrin pi-cation radical. Reduction of Compound II by L-ascorbate is rate-limiting: at low substrate concentration (0-500 microM), kinetic traces were monophasic but above approximately 500 microM were biphasic. Observed rate constants for the fast phase overlaid with observed rate constants extracted from the (monophasic) dependence observed below 500 microM and showed saturation kinetics; rate constants for the slow phase were linearly dependent on substrate concentration (k(3-slow)) = 3.1 +/- 0.1 x 10(3) M(-1) s(-1)). Kinetic transients for reduction of Compound II by L-ascorbic acid for Cys32-modified rsAPX are monophasic at all substrate concentrations, and the second-order rate constant (k(3) = 0.9 +/- 0.1 x 10(3) M(-1) s(-1)) is similar to that obtained from the slow phase of Compound II reduction for unmodified rsAPX. Steady-state oxidation of L-ascorbate by rsAPX showed a sigmoidal dependence on substrate concentration and data were satisfactorily rationalized using the Hill equation; oxidation of L-ascorbic acid by Cys32-modified rsAPX showed no evidence of sigmoidal behavior. The data are consistent with the presence of two kinetically competent binding sites for ascorbate in APX.  相似文献   

2.
Catalytic oxidation of p-cresol by ascorbate peroxidase   总被引:2,自引:0,他引:2  
Transient and steady state kinetics, together with a range of chromatographic and spectroscopic techniques, have been used to establish the mechanism and the products of the H(2)O(2)-dependent oxidation of p-cresol by ascorbate peroxidase (APX). HPLC, GC-MS, and NMR analyses are consistent with the formation of 2, 2'-dihydroxy-5,5'-dimethylbiphenyl (II) and 4alpha,9beta-dihydro-8, 9beta-dimethyl-3(4H)-dibenzofuranone (Pummerer's ketone, III) as the major products of the reaction. In the presence of cumene hydroperoxide, two additional products were observed which, from GC and MS analyses, were shown to be 1,1-dimethylbenzylalcohol (IV) and bis-(1-methyl-1-phenyl-ethyl)-peroxide (V). The product ratio II:III was dependent on enzyme concentration: at low concentrations Pummerer's ketone (III) predominates and at high concentrations formation of the biphenyl compound (II) is favored. Steady-state data showed a sigmoidal dependence on [p-cresol] that was consistent with the presence of 2.01 +/- 0.15 binding sites for the substrate (25.0 degrees C, sodium phosphate, pH 7.0, mu = 2.2 mM) and independent of ionic strength in the range 2.2-500 mM. Single turnover kinetic experiments (pH 7.0, 5.0 degrees C, mu = 0.10 M) yielded second-order rate constants for Compound I reduction by p-cresol, k(2), of 5.42 +/- 0.10 x 10(5) M(-1) s(-1), respectively. Rate-limiting reduction of Compound II by p-cresol, k(3), showed saturation kinetics, giving values for K(d) = 1.54 +/- 0.12 x 10(-3) M and k(3) = 18.5 +/- 0.7 s(-1). The results are discussed in the more general context of APX-catalyzed aromatic oxidations.  相似文献   

3.
It is demonstrated that horseradish peroxidase (HRP) mixed with chlorite follows the whole peroxidase cycle. Chlorite mediates the two-electron oxidation of ferric HRP to compound I (k(1)) thereby releasing hypochlorous acid. Furthermore, chlorite acts as one-electron reductant of both compound I (k(2)) and compound II (k(3)) forming chlorine dioxide. The strong pH-dependence of all three reactions clearly suggests that chlorous acid is the reactive species. Typical apparent bimolecular rate constants at pH 5.6 are 1.4 x 10(5)M(-1)s(-1) (k(1)), 2.25 x 10(5)M(-1)s(-1) (k(2)), and 2.4 x 10(4)M(-1)s(-1) (k(3)), respectively. Moreover, the reaction products hypochlorous acid and chlorine dioxide, which are known to induce heme bleaching and amino acid modification upon longer incubation times, also mediate the oxidation of ferric HRP to compound I (2.4 x 10(7)M(-1)s(-1) and 2.7 x 10(4)M(-1)s(-1), respectively, pH 5.6) but do not react with compounds I and II. A reaction scheme is presented and discussed from both a mechanistic and thermodynamic point of view. It helps to explain the origin of contradictory data so far found in the literature on this topic.  相似文献   

4.
Nitric oxide-derived oxidants (e.g., peroxynitrite) are believed to participate in antimicrobial activities as part of normal host defenses but also in oxidative tissue injury in inflammatory disorders. A similar role is ascribed to the heme enzyme myeloperoxidase (MPO), the most abundant protein of polymorphonuclear leukocytes, which are the terminal phagocytosing effector cells of the innate immune system. Concomitant production of peroxynitrite and release of millimolar MPO are characteristic events during phagocytosis. In order to understand the mode of interaction between MPO and peroxynitrite, we have performed a comprehensive stopped-flow investigation of the reaction between all physiological relevant redox intermediates of MPO and peroxynitrite. Both iron(III) MPO and iron(II) MPO are rapidly converted to compound II by peroxynitrite in monophasic reactions with calculated rate constants of (6.8+/-0.1) x 10(6) M(-1)s(-1) and (1.3+/-0.2) x 10(6) M(-1)s(-1), respectively (pH 7.0 and 25 degrees C). Besides these one- and two-electron reduction reactions of peroxynitrite, which produce nitrogen dioxide and nitrite, a one-electron oxidation to the oxoperoxonitrogen radical must occur in the fast monophasic transition of compound I to compound II mediated by peroxynitrite at pH 7.0 [(7.6+/-0.1) x 10(6) M(-1)s(-1)]. In addition, peroxynitrite induced a steady-state transition from compound III to compound II with a rate of (1.0+/-0.3) x 10(4) M(-1)s(-1). Thus, the interconversion among the various oxidation states of MPO that is prompted by peroxynitrite is remarkable. Reaction mechanisms are proposed and the physiological relevance is discussed.  相似文献   

5.
The reaction of native myeloperoxidase (MPO) and its redox intermediate compound I with hydrogen peroxide, ethyl hydroperoxide, peroxyacetic acid, t-butyl hydroperoxide, 3-chloroperoxybenzoic acid and cumene hydroperoxide was studied by multi-mixing stopped-flow techniques. Hydroperoxides are decomposed by MPO by two mechanisms. Firstly, the hydroperoxide undergoes a two-electron reduction to its corresponding alcohol and heme iron is oxidized to compound I. At pH 7 and 15 degrees C, the rate constant of the reaction between 3-chloroperoxybenzoic acid and ferric MPO was similar to that with hydrogen peroxide (1.8x10(7) M(-1) s(-1) and 1.4x10(7) M(-1) s(-1), respectively). With the exception of t-butyl hydroperoxide, the rates of compound I formation varied between 5.2x10(5) M(-1) s(-1) and 2.7x10(6) M(-1) s(-1). Secondly, compound I can abstract hydrogen from these peroxides, producing peroxyl radicals and compound II. Compound I reduction is shown to be more than two orders of magnitude slower than compound I formation. Again, with 3-chloroperoxybenzoic acid this reaction is most effective (6. 6x10(4) M(-1) s(-1) at pH 7 and 15 degrees C). Both reactions are controlled by the same ionizable group (average pK(a) of about 4.0) which has to be in its conjugated base form for reaction.  相似文献   

6.
Recently, it was suggested that melatonin (N-acetyl-5-methoxytryptamine) is oxidized by activated neutrophils in a reaction most probably involving myeloperoxidase (Biochem. Biophys. Res. Commun. (2000) 279, 657-662). Myeloperoxidase (MPO) is the most abundant protein of neutrophils and is involved in killing invading pathogens. To clarify if melatonin is a substrate of MPO, we investigated the oxidation of melatonin by its redox intermediates compounds I and II using transient-state spectral and kinetic measurements at 25 degrees C. Spectral and kinetic analysis revealed that both compound I and compound II oxidize melatonin via one-electron processes. The second-order rate constant measured for compound I reduction at pH 7 and pH 5 are (6.1 +/- 0.2) x 10(6) M(-1) s(-1) and (1.0 +/- 0.08) x 10(7) M(-1) s(-1), respectively. The rates for the one-electron reduction of compound II back to the ferric enzyme are (9.6 +/- 0.3) x 10(2) M(-1) s(-1) (pH 7) and (2.2 +/- 0.1) x 10(3) M(-1) s(-1) (pH 5). Thus, melatonin is a much better electron donor for compound I than for compound II. Steady-state experiments showed that the rate of oxidation of melatonin is dependent on the H(2)O(2) concentration, is not affected by superoxide dismutase, and is quickly terminated by sodium cyanide. Melatonin can markedly inhibit the chlorinating activity of MPO at both pH 7 and pH 5. The implication of these findings in the activated neutrophil is discussed.  相似文献   

7.
The equilibria and kinetics of the interactions of proflavine (PR) and its platinum-containing derivative [PtCl(tmen)(2)HNC(13)H(7)(NHCH(2)CH(2))(2)](+) (PRPt) with double-stranded poly(A) have been investigated by spectrophotometry and Joule temperature-jump relaxation at ionic strength 0.1 M, 25 degrees C, and pH 5.2. Spectrophotometric measurements indicate that base-dye interactions are prevailing. T-jump experiments with polarized light showed that effects due to field-induced alignment could be neglected. Both of the investigated systems display two relaxation effects. The kinetic features of the reaction are discussed in terms of a two-step series mechanism in which a precursor complex DS(I) is formed in the fast step, which is then converted to a final complex in the slow step. The rate constants of the fast step are k(1) = (2.5 +/- 0.4) x 10(6) M(-1) s(-1), k(-1) = (2.4 +/- 0.1) x 10(3) s(-1) for poly(A)-PR and k(1) = (2.3 +/- 0.1) x 10(6) M(-1) s(-1), k(-1) = (1.6 +/- 0.2) x 10(3) s(-1) for poly(A)-PRPt. The rate constants for the slow step are k(2) = (4.5 +/- 0.5) x 10(2) s(-1), k(-2) = (1.7 +/- 0.1) x 10(2) s(-1) for poly(A)-PR and k(2) = 9.7 +/- 1.2 s(-1), k(-2) = 10.6 +/- 0.2 s(-1) for poly(A)-PRPt. Spectrophotometric measurements yield for the equilibrium constants and site size the values K = (4.5 +/- 0.1) x 10(3) M(-1), n = 1.3 +/- 0.5 for poly(A)-PR and K = (2.9 +/- 0.1) x 10(3) M(-1), n = 2.3 +/- 0.6 for poly(A)-PRPt. The values of k(1) are similar and lower than expected for diffusion-limited reactions. The values of k(-1) are similar as well. It is suggested that the formation of DS(I) involves only the proflavine residues in both systems. In contrast, the values of k(2) and k(-2) in poly(A)-PRPt are much lower than in poly(A)-PR. The results suggest that in the complex DS(II) of poly(A)-PRPt both proflavine and platinum residues are intercalated. In addition, a very slow process was detected and ascribed to the covalent binding of Pt(II) to the adenine.  相似文献   

8.
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   

9.
The binding of proflavine (D) to single stranded poly(A) (P) was investigated at pH 7.0 and 25 degrees C using T-jump, stopped-flow and spectrophotometric methods. Equilibrium measurements show that an external complex PD(I) and an internal complex PD(II) form upon reaction between P and D and that their concentrations depend on the polymer/dye concentration ratio (C(P)/C(D)). For C(P)/C(D)<2.5, cooperative formation of stacks external to polymer strands prevails (PD(I)). Equilibria and T-jump experiments, performed at I=0.1M and analyzed according to the Schwarz theory for cooperative binding, provide the values of site size (g=1), equilibrium constant for the nucleation step (K( *)=(1.4+/-0.6)x10(3)M(-1)), equilibrium constant for the growth step (K=(1.2+/-0.6)x10(5)M(-1)), cooperativity parameter (q=85) and rate constants for the growth step (k(r)=1.2x10(7)M(-1)s(-1), k(d)=1.1 x 10(2)s(-1)). Stopped-flow experiments, performed at low ionic strength (I=0.01 M), indicate that aggregation of stacked poly(A) strands do occur provided that C(P)/C(D)<2.5.  相似文献   

10.
The oxidation of D-lactobionic acid by Cr(VI) yields the 2-ketoaldobionic acid and Cr(3+) as final products when a 20-times or higher excess of the aldobionic acid over Cr(VI) is used. The redox reaction takes place through a complex multistep mechanism, which involves the formation of intermediate Cr(IV) and Cr(V) species. Cr(IV) reacts with lactobionic acid much faster than Cr(V) and Cr(VI) do, and cannot be directly detected. However, the formation of CrO(2)(2+), observed by the first time for an acid saccharide/Cr(VI) system, provides indirect evidence for the intermediacy of Cr(IV) in the reaction path. Cr(VI) and the intermediate Cr(V) react with lactobionic acid at comparable rates, being the complete rate laws for the Cr(VI) and Cr(V) consumption expressed by: -d[Cr(VI)]/dt=[k(I)+k(II)[H(+)]][lactobionicacid][Cr(VI)], where k(I)=(4.1+/-0.1) x 10(-3) M(-1) s(-1) and k(II)=(2.1+/-0.1) x 10(-2) M(-2) s(-1); and -d[Cr(V)]/dt=[k(III)[H(+)]+(k(IV)+k(V)[H(+)])[lactobionicacid]] [Cr(V)], where k(III)=(1.8+/-0.1) x 10(-3) M(-1) s(-1), k(IV)=(1.1+/-0.1) x 10(-2) M(-1) s(-1) and k(V)=(1.0+/-0.1) x 10(-2) M(-2) s(-1), at 33 degrees C. The Electron Paramagnetic Resonance (EPR) spectra show that five-co-ordinate oxo-Cr(V) bischelates are formed at pH 1-5 with the aldobionic acid bound to Cr(V) through the alpha-hydroxyacid group.  相似文献   

11.
Reaction of lactoperoxidase compound I with halides and thiocyanate   总被引:2,自引:0,他引:2  
Lactoperoxidase (LPO) is found in mucosal surfaces and exocrine secretions, including milk, tears, and saliva, and has physiological significance in antimicrobial defense which involves (pseudo-) halide oxidation. This study for the first time presents transient kinetic measurements of the reactivity of its competent redox intermediate compound I with halides and thiocyanate, using the sequential stopped-flow technique. Compound I was produced with either H(2)O(2) [(1.1 +/- 0.1) x 10(7) M(-1) s(-1)] or hypochlorous acid [(3.2 +/- 0.1) x 10(7) M(-1) (s-1)]. At pH 7 and 15 degrees C, the two-electron reduction of compound I to native LPO by bromide and iodide has a second-order rate constant of (4.1 +/- 0.1) x 10(4) M(-1) s(-1) and (1.2 +/- 0.04) x 10(8) M(-1) s(-1), respectively. With thiocyanate the reaction is extremely fast (2.0 x 10(8) M(-1) s(-1)), whereas chloride cannot function as electron donor. The results are discussed with respect to known kinetic data of homologous mammalian peroxidases and to the physiological role of LPO in antimicrobial defense.  相似文献   

12.
Eosinophil peroxidase, the major granule protein in eosinophils, is the least studied human peroxidase. Here, we have performed spectral and kinetic measurements to study the nature of eosinophil peroxidase intermediates, compounds I and II, and their reduction by the endogenous one-electron donors ascorbate and tyrosine using the sequential-mixing stopped-flow technique. We demonstrate that the peroxidase cycle of eosinophil peroxidase involves a ferryl/porphyrin radical compound I and a ferryl compound II. In the absence of electron donors, compound I is shown to be transformed to a species with a compound II-like spectrum. In the presence of ascorbate or tyrosine compound I is reduced to compound II with a second-order rate constant of (1.0+/-0.2)x10(6) M(-1) s(-1) and (3.5+/-0.2)x10(5) M(-1) s(-1), respectively (pH 7.0, 15 degrees C). Compound II is then reduced by ascorbate and tyrosine to native enzyme with a second-order rate constant of (6.7+/-0.06)x10(3) M(-1) s(-1) and (2.7+/-0.06)x10(4) M(-1) s(-1), respectively. This study revealed that eosinophil peroxidase compounds I and II are able to react with tyrosine and ascorbate via one-electron oxidations and therefore generate monodehydroascorbate and tyrosyl radicals. The relatively fast rates of the compound I reduction demonstrate that these reactions may take place in vivo and are physiologically relevant.  相似文献   

13.
The unfolding and refolding of creatine kinase (ATP:creatine N-phosphotransferase (CK), EC 2.7.3.2) during denaturation and reactivation by trifluoroethanol (TFE) have been studied. Significant aggregation was observed when CK was denatured at TFE concentrations between 10% and 40% (v/v). 50% TFE (v/v) was used to study the denaturation and unfolding of CK. The activity loss of CK was a very quick process, as was the marked conformational changes during denaturation followed by fluorescence emission spectra and far-ultraviolet CD spectra. DTNB modification and size exclusion chromatography were used to find that CK dissociated and was in its monomer state after denaturation with 50% TFE. Reactivation and refolding were observed after 80-fold dilution of the denatured CK into 0.05 M Tris-HCl buffer, pH 8.0. The denatured CK recovered about 38% activity following a two phase course (k(1)=4.82+/-0.41x10(-3) s(-1), k(2)=0.60+/-0.01x10(-3) s(-1)). Intrinsic fluorescence maximum intensity changes showed that the refolding process also followed biphasic kinetics (k(1)=4.34+/-0.27x10(-3) s(-1), k(2)=0.76+/-0.02x10(-3) s(-1)) after dilution into the proper solutions. The far-ultraviolet CD spectra ellipticity changes at 222 nm during the refolding process also showed a two phase course (k(1)=4.50+/-0.07x10(-3) s(-1), k(2)=1.13+/-0.05x10(-3) s(-1)). Our results suggest that TFE can be used as a reversible denaturant like urea and GuHCl. The 50% TFE induced CK denaturation state, which was referred to as the 'TFE state', and the partially refolded CK are compared with the molten globule state. The aggregation caused by TFE during denaturation is also discussed in this paper.  相似文献   

14.
Oxidation of l-serine and l-threonine by a silver(III) complex anion, [Ag(HIO(6))(2)](5-), has been studied in aqueous alkaline medium. The oxidation products of the amino acids have been identified as ammonia, glyoxylic acid and aldehyde (formaldehyde for serine and acetaldehyde for threonine). Kinetics of the oxidation reactions has been followed by the conventional spectrophotometry in the temperature range of 20.0-35.0 degrees C and the reactions display an overall second-order behavior: first-order with respect to both Ag(III) and the amino acids. Analysis of influences of [OH(-)] and [periodate] on the second-order rate constants k' reveals an empirical rate expression: k(')=(k(a)+k(b)[OH(-)])K(1)/([H(2)IO(6)(3-)](e)+K(1)), where [H(2)IO(6)(3-)](e) is equilibrium concentration of periodate, and where k(a)=6.1+/-0.5M(-1)s(-1), k(b)=264+/-6M(-2)s(-1), and K(1)=(6.5+/-1.3)x10(-4)M for serine and k(a)=12.6+/-1.7M(-1)s(-1), k(b)=(5.5+/-0.2)x10(2)M(-2)s(-1), and K(1)=(6.2+/-1.5)x10(-4)M for threonine at 25.0 degrees C and ionic strength of 0.30M. Activation parameters associated with k(a) and k(b) have also been derived. A reaction mechanism is proposed to involve two pre-equilibria, leading to formation of an Ag(III)-periodato-amino acid ternary complex. The ternary complex undergoes a two-electron transfer from the coordinated amino acid to the metal center via two parallel pathways: one pathway is spontaneous and the other is assisted by a hydroxide ion. Potential applications of the Ag(III) complex as a reagent for modifications of peptides and proteins are implicated.  相似文献   

15.
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   

16.
In human myeloperoxidase the heme is covalently attached to the protein via two ester linkages between the carboxyl groups of Glu242 and Asp94 and modified methyl groups on pyrrole rings A and C of the heme as well as a sulfonium ion linkage between the sulfur atom of Met243 and the beta-carbon of the vinyl group on pyrrole ring A. In the present study, wild-type recombinant myeloperoxidase (recMPO) and the variant Glu242Gln were produced in Chinese hamster ovary cells and investigated in a comparative sequential-mixing stopped-flow study in order to elucidate the role of the Glu242-heme ester linkage in the individual reaction steps of both the halogenation and peroxidase cycle. Disruption of the ester bond increased heme flexibility, blue shifted the UV-vis spectrum, and, compared with recMPO, decelerated cyanide binding (1.25 x 10(4) versus 1.6 x 10(6) M(-)(1) s(-)(1) at pH 7 and 25 degrees C) as well as compound I formation mediated by either hydrogen peroxide (7.8 x 10(5) versus 1.9 x 10(7) M(-)(1) s(-)(1)) or hypochlorous acid (7.5 x 10(5) versus 2.3 x 10(7) M(-)(1) s(-)(1)). The overall chlorination and bromination activity of Glu242Gln was 2.0% and 24% of recMPO. The apparent bimolecular rate constants of compound I reduction by chloride (65 M(-)(1) s(-)(1)), bromide (5.4 x 10(4) M(-)(1) s(-)(1)), iodide (6.4 x 10(5) M(-)(1) s(-)(1)), and thiocyanate (2.2 x10(5) M(-)(1) s(-)(1)) were 500, 25, 21, and 63 times decreased compared with recMPO. By contrast, Glu242Gln compound I reduction by tyrosine was only 5.4 times decreased, whereas tyrosine-mediated compound II reduction was 60 times slower compared with recMPO. The effects of exchange of Glu242 on electron transfer reactions are discussed.  相似文献   

17.
We describe the purification of a cysteine proteinase from latex of Carica candamarcensis, hereby designated CC23. The enzyme has been purified by ion-exchange chromatography and behaves electrophoretically as a monomer of M(r) 23,000 and optimal pH of 8.0. It displays a basic isoelectric point, has one cysteine residue in the active site by titration with E-64, confirmed by DNA sequencing, and responds to proteinase inhibitors as a classic cysteine proteinase. The K(m) and k(cat)/K(m) for CC23 using BAPNA were respectively 14.7 +/- 1.8 x 10(-4) M and 1.3 x 10(3) M(-1) s(-1). Therefore, the catalytic efficiency of CC23 is sixfold higher than that of CC-I, another proteinase from the same plant. DNA primers were designed to amplify by PCR a genomic sequence related to this enzyme. An 895-bp DNA fragment was cloned and sequenced. It shows strong homology with chymopapain isoform IV from C. papaya. The translated sequence is similar to that of chymopapain isoform II (73%) and CC-III (77%) from C. candamarcensis.  相似文献   

18.
The binding of Ru(phen)(2)dppz(2+) (dppz=dipyrido[3,2-a:2',3'-c]phenazine) to DNA was investigated at pH 7.0 and 25 degrees C using stopped-flow and spectrophotometric methods. Equilibrium measurements show that two modes of binding, whose characteristics depend on the polymer to dye ratio (C(P)/C(D)), are operative. The binding mode occurring for values of C(P)/C(D) higher than 3 exhibits positive cooperativity, which is confirmed by kinetic experiments. The reaction parameters are K=2 x 10(3)M(-1), omega=550, n=1, k(r)=(1.9+/-0.5) x 10(7)M(-1)s(-1) and k(d)=(9.5+/-2.5)x10(3)s(-1) at I=0.012 M. The results are discussed in terms of prevailing surface interaction with DNA grooves accompanied by partial intercalation of the dppz residue. The other binding mode becomes operative for C(P)/C(D)<3 and the equilibria analysis shows this is an ordinary intercalation mode (K=1.3 x 10(6) M(-1), n=1.5 at I=0.012 M and K=2 x 10(5) M(-1), n=1.2 at I=0.21 M). Similar behaviour is displayed by double-stranded poly(A).  相似文献   

19.
Whether age-related differences in blood lactate concentrations (BLC) reflect specific BLC kinetics was analyzed in 15 prepubescent boys (age 12.0 +/- 0.6 yr, height 1.54 +/- 0.06 m, body mass 40.0 +/- 5.2 kg), 12 adolescents (16.3 +/- 0.7 yr, 1.83 +/- 0.07 m, 68.2 +/- 7.5 kg), and 12 adults (27.2 +/- 4.5 yr, 1.83 +/- 0.06 m, 81.6 +/- 6.9 kg) by use of a biexponential four-parameter kinetics model under Wingate Anaerobic Test conditions. The model predicts the lactate generated in the extravasal compartment (A), invasion (k(1)), and evasion (k(2)) of lactate into and out of the blood compartment, the BLC maximum (BLC(max)), and corresponding time (TBLC(max)). BLC(max) and TBLC(max) were lower (P < 0.05) in boys (BLC(max) 10.2 +/- 1.3 mmol/l, TBLC(max) 4.1 +/- 0.4 min) than in adolescents (12.7 +/- 1.0 mmol/l, 5.5 +/- 0.7 min) and adults (13.7 +/- 1.4 mmol/l, 5.7 +/- 1.1 min). No differences were found in A related to the muscle mass (A(MM)) and k(1) between boys (A(MM): 22.8 +/- 2.7 mmol/l, k(1): 0.865 +/- 0.115 min(-1)), adolescents (22.7 +/- 1.3 mmol/l, 0.692 +/- 0.221 min(-1)), and adults (24.7 +/- 2.8 mmol/l, 0.687 +/- 0.287 min(-1)). The k(2) was higher (P < 0.01) in boys (2.87 10(-2) +/- 0.75 10(-2) min(-1)) than in adolescents (2.03 x 10(-2) +/- 0.89 x 10(-2) min(-1)) and adults (1.99 x 10(-2) +/- 0.93 x 10(-2) min(-1)). Age-related differences in the BLC kinetics are unlikely to reflect differences in muscular lactate or lactate invasion but partly faster elimination out of the blood compartment.  相似文献   

20.
A sensitive competitive method was developed for assessing the reactivity of compounds toward glutathione and toward thiols in general. The method employs the reaction of the fluorogenic reagent fluorescein-5-maleimide (FM) with glutathione (GSH) to generate a large increase in fluorescence emission. When the reaction is measured in the presence of a compound that competes with FM toward GSH, the rate constant for fluorescent product formation increases while the total amount of product formed at the end of the reaction decreases. These changes in the presence of a series of competitor concentrations allow one to calculate the rate constant of the reaction of the competitor with GSH. At 23 degrees C, pH 7.40 in PBS buffer the second-order rate constant of the FM-GSH reaction is k2 = (1.67 +/- 0.32) x 10(4) M(-1) x s(-1). Two GSH-reactive compounds were evaluated: the second-order rate constant for the reaction of PNU-27707 with GSH under our experimental conditions is k(i) = 5660 +/- 266 M(-1) x s(-1), while that of PNU-37802 is k(i) = 21,200 +/- 1600 M(-1) x s(-1). The method is easily adaptable to a high-throughput screening format.  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