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.
Apocynin has been reported to require dimerization by myeloperoxidase (MPO) to inhibit leukocyte NADPH oxidase. (-)-Epicatechin, a dietary flavan-3-ol, has been identified as a ‘prodrug’ of apocynin-like metabolites that inhibit endothelial NADPH oxidase activity and elevate the cellular level of nitric oxide. Since (-)-epicatechin has tentatively been identified as substrate of MPO, we studied the one-electron oxidation of (-)-epicatechin by MPO. By using multi-mixing stopped-flow technique, we demonstrate that (-)-epicatechin is one of the most efficient electron donors for heme peroxidases investigated so far. Second order rate constants for the (-)-epicatechin-mediated conversion of MPO-compound I to compound II and compound II to resting enzyme were estimated to be 1.9 × 107 and 4.5 × 106 M−1 s−1, respectively (pH 7, 25 °C). The data indicate that (-)-epicatechin is capable of undergoing fast MPO-mediated one-electron oxidation.  相似文献   

2.
The standard reduction potential of the redox couple compound I/native enzyme has been determined for human myeloperoxidase (MPO) and eosinophil peroxidase (EPO) at pH 7.0 and 25 degrees C. This was achieved by rapid mixing of peroxidases with either hydrogen peroxide or hypochlorous acid and measuring spectrophotometrically concentrations of the reacting species and products at equilibrium. By using hydrogen peroxide, the standard reduction potential at pH 7.0 and 25 degrees C was 1.16 +/- 0.01 V for MPO and 1.10 +/- 0.01 V for EPO, independently of the concentration of hydrogen peroxide and peroxidases. In the case of hypochlorous acid, standard reduction potentials were dependent on the hypochlorous acid concentration used. They ranged from 1.16 V at low hypochlorous acid to 1.09 V at higher hypochlorous acid for MPO and from 1.10 V to 1.03 V for EPO. Thus, consistent results for the standard reduction potentials of redox couple compound I/native enzyme of both peroxidases were obtained with all hydrogen peroxide and at low hypochlorous acid concentrations: possible reasons for the deviation at higher concentrations of hypochlorous acid are discussed. They include instability of hypochlorous acid, reactions of hypochlorous acid with different amino-acid side chains in peroxidases as well as the appearance of a compound I-chloride complex.  相似文献   

3.
Myeloperoxidase (MPO), which is involved in host defence and inflammation, is a unique peroxidase in having a globin-like standard reduction potential of the ferric/ferrous couple. Intravacuolar and exogenous MPO released from stimulated neutrophils has been shown to exist in the oxyferrous form, called compound III. To investigate the reactivity of ferrous MPO with molecular oxygen, a stopped-flow kinetic analysis was performed. In the absence of dioxygen, ferrous MPO decays to ferric MPO (0.04 s(-1) at pH 8 versus 1.4 s(-1) at pH 5). At pH 7.0 and 25 degrees C, compound III formation (i.e., binding of dioxygen to ferrous MPO) occurs with a rate constant of (1.1+/-0.1) x 10(4)M(-1)s(-1). The rate doubles at pH 5.0 and oxygen binding is reversible. At pH 7.0, the dissociation equilibrium constant of the oxyferrous form is (173+/-12)microM. The rate constant of dioxygen dissociation from compound III is much higher than conversion of compound III to ferric MPO (which is not affected by the oxygen concentration). This allows an efficient transition of compound III to redox intermediates which actually participate in the peroxidase or halogenation cycle of MPO.  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.
Oxidized low-density lipoproteins (LDL) accumulate in the vascular wall and promote a local inflammatory process contributing to the progression of atheromatous plaque. The key role of myeloperoxidase (MPO) in this process has been documented and the enzyme has been involved in the oxidative modification of apolipoprotein B-100 in the intima and at the surface of endothelial cells. As the inhibition of this last phenomenon could be of relevance in pharmacological interventions, thiol-containing molecules such as glutathione, captopril, and N-acetylcysteine (NAC) and its lysinate salt (NAL) were tested in this system and their properties were compared with those of flufenamic acid (control). This last compound already demonstrated an inhibition of the production of HOCl by MPO and a more intense inhibition of MPO activity than glutathione, NAC, NAL, and captopril. However, NAC and NAL inhibited the oxidative modification of LDL more intensively than captopril and glutathione whereas flufenamic acid had no comparable inhibiting effect. This could be related to the presence of LDL close to the catalytic site of the enzyme. NAC and NAL therefore appeared as the most efficient inhibitors probably as a consequence of their relatively small size. The relevance of such effects has to be documented by in vivo studies.  相似文献   

6.
Myeloperoxidase (MPO) catalyzes the formation of oxidants that have been implicated in the pathogenesis of various diseases, including cardiovascular and pulmonary diseases and cancer. Inhibition of MPO oxidant-generating activity represents an attractive target for preventing the progression of inflammatory conditions. Peroxidation and chlorination catalytic activity were utilized to screen for the most effective tryptophan analog that inhibits MPO. Rapid kinetic measurements were performed to determine the mechanisms through which these compounds inhibit the catalytic activity of MPO. Substituents on the amino and carboxyl termini of tryptophan enhance its affinity toward MPO compared to a substituent on the indole ring. Hydrogen-bond donor capabilities and a positive charge on the amino group are not required for MPO inhibition. Hydroxyl-containing indoles did not inhibit MPO H2O2-consumption activity. Elimination of the negative charge from the carboxyl terminus by introducing a hydrophobic character significantly enhanced tryptophan analog affinity for MPO and improved its inhibitory properties. Further mechanistic studies indicated that indole compounds inhibited MPO activity through the accumulation of compound II, an inactive MPO intermediate. Our results show that specific structural features of tryptophan analogs are involved in increasing the affinity for MPO and providing a significantly greater inhibition of MPO's catalytic activities.  相似文献   

7.
Myeloperoxidase (MPO), eosinophil peroxidase, lactoperoxidase, and thyroid peroxidase are heme-containing oxidoreductases (EC 1.7.1.11), which bind ligands and/or undergo a series of redox reactions. Though sharing functional and structural homology, reflecting their phylogenetic origin, differences are observed regarding their spectral features, substrate specificities, redox properties, and kinetics of interconversion of the relevant redox intermediates ferric and ferrous peroxidase, compound I, compound II, and compound III. Depending on substrate availability, these heme enzymes path through the halogenation cycle and/or the peroxidase cycle and/or act as poor (pseudo-)catalases. Based on the published crystal structures of free MPO and its complexes with cyanide, bromide and thiocyanate as well as on sequence analysis and modeling, we critically discuss structure-function relationships. This analysis highlights similarities and distinguishing features within the mammalian peroxidases and intents to provide the molecular and enzymatic basis to understand the prominent role of these heme enzymes in host defense against infection, hormone biosynthesis, and pathogenesis.  相似文献   

8.
The reversibility of the stepwise reduction of Compound I to the ferric state via Compound II was confirmed in horseradish peroxidases A2 and C. The values of E'o (compound I/Compound II) and E'O (Compound II/ferric) were measured from equilibrium data coupled with the K2IrCl6-K3IrCl6 system in a narrow region of pH near 6.3. The ferric enzymes were also oxidized by ferricyanide to Compound II at alkaline pH and the values of E'O (Compound II/ferric) were measured from the equilibrium data. The pH dependence of E'O (Compound II/ferric) was in accord with the equation: E'O = EO + 0.058 log (Kr[H+] + [H+]2)/(KO + [H+]), where Kr and KO are proton dissociation constants in the ferric enzyme and Compound II, respectively. The pH-E'O (Compound I/Compound II) curves were likewise obtained from the equation, E'O = EO + 0.058 log (Kr + [H+]), where Kr is the proton dissociation constant in Compound II. The forward and backward rate constants were measured in each of one-electron transfer reactions of the peroxidases with the K2IrCl6-K3IrCl6 system at various pH values. The E'O values calculated on the assumption that the ratio of the rate constants equals the equilibrium constant were compared with those obtained from the equilibrium data.  相似文献   

9.
A comparative study on the reactivity of five indole derivatives (tryptamine, N-acetyltryptamine, tryptophan, melatonin, and serotonin), with the redox intermediates compound I (k2) and compound II (k3) of the plant enzyme horseradish peroxidase (HRP) and the two mammalian enzymes lactoperoxidase (LPO) and myeloperoxidase (MPO), was performed using the sequential-mixing stopped-flow technique. The calculated bimolecular rate constants (k2, k3) revealed substantial differences regarding the oxidazibility of the substrates by redox intermediates at pH 7.0 and 25 degrees C. With HRP it was shown that k2 and k3 are mainly determined by the reduction potential (Eo') of the substrate with k2 being 7-45 times higher than k3. Compound I of mammalian peroxidases was a much better oxidant than HRP compound I with the consequence that the influence of the indole structure on k2 of LPO and MPO was small varying by a factor of only 88 and 38, respectively, which is in strong contrast to a factor of 160,000 determined for k2 of HRP. Interestingly, the k3 values for all three enzymes were very similar. Oxidation of substrates by mammalian peroxidase compound II is strongly constrained by the nature of the substrate. The k3 values for the five indoles varied by a factor of 3,570 (LPO) and 200,000 (MPO), suggesting that the reduction potential of compound II of mammalian peroxidase is less positive than that of compound I, which is in contrast to the plant enzyme.  相似文献   

10.
The membrane-bound component of the phosphotransferase system of Escherichia coli, responsible for the phosphorylative uptake of methyl-α-d-glucoside has an essential thiol group which becomes available to inactivation by thiol reagents in the presents of the phosphate-accepting sugar or when phosphoenolpyruvate synthesis is inhibited. The form resistant to the thiol reagent requires not only the absence of sugar and an intact phosphoenol-pyruvate generating system, but also an intact system generating phosphorylated Hpr which is impaired by heating of a thermosensitive enzyme I mutant.  相似文献   

11.
12.
Transketolase (TK), a thiamin diphosphate (ThDP) dependent enzyme, catalyzes the reversible transfer of a two-carbon unit from keto- to aldo-substrates. Dihydroxyethylthiamin diphosphate (DHEThDP), formed as a result of cleavage of the donor substrate, serves as an intermediate of the TK reaction. TK from the yeast Saccharomyces cerevisiae is unique among thiamin enzymes displaying enzymatic activity after reconstitution with a methylated analogue of the native cofactor, 4′-methylamino-ThDP. The reconstitution of the apoenzyme with both ThDP and the methylated analogue can be analyzed by near UV circular dichroism. It was demonstrated that in the native holoenzyme and in the complex of TK with 4′-methylamino-ThDP the formation of the dihydroxyethyl-based carbanion/enamine took place with comparable rate constants, whereas the protonation of the reactive species was much faster in the complex with the analogue. The enzymatic activity of the enzyme reconstituted with 4′-methylamino-ThDP was 10fold higher in the ferricyanide assay. We suggest that a methylation of the 4′-amino group of ThDP impairs the resonance stabilization of the carbanion/enamine intermediate both sterically and electronically, thus allowing either a faster protonation or oxidation reaction by ferricyanide. The formation of the optically active DHE-4′-methylamino-ThDP was monitored by near UV circular dichroism spectra and corroborated by 1H NMR analysis. The protonated form of the intermediate DHE-4′-methylamino-ThDP was released from the active sites of TK and accumulated in the medium on preparative scale.  相似文献   

13.
Studies that elucidate the behavior of the hemoglobins (Hbs) and myoglobins upon reaction with hydrogen peroxide are essential to the development of oxygen carrier substitutes. Stopped-flow kinetics and resonance Raman data show that the reaction between hydrogen peroxide and oxygenated and deoxygenated ferric Hb I (oxy- and deoxy-HbI) from Lucina pectinata produce compound I and compound II ferryl species. The rate constants ratio (k23/k41) between the formation of compound II from compound I (k23) and the oxidation of the ferrous HbI (k41, i.e., 25 M(-1) s(-1)) of 12 x 10(-4) M suggests that HbI has a peroxidative capacity for removing H2O2 from solution. Resonance Raman presents the formation of both, met-aquo-HbI and compound II ferryl species in the cyclic reaction of HbI with H2O2. The ferric HbI species is maintained by the presence of H2O2; it can produce HbI compound I, or it can be reduced to a deoxy-HbI derivative to form HbI compound II upon reaction with H2O2. The compound II ferryl vibration frequency appears at 805 and 769 cm(-1) for HbIFe(IV)=(16)O and HbIFe(IV)=(18)O species, respectively. This ferryl mode indicates the absence of hydrogen bonding between the carbonyl group of the distal Q64 and the HbIFe(IV)=O ferryl moiety. The observation suggests that both the trans-ligand effect and the polarizabilty of the HbI heme pocket are responsible for the observed ferryl oxo vibrational energy. The vibrational mode also suggests that the carbonyl group of the distal Q64 is oriented toward the iron of the heme group, increasing the distal pocket electron density.  相似文献   

14.
A. Ried  B. Hessenberg  H. Metzler  R. Ziegler 《BBA》1977,459(2):175-186
Action spectra of light reaction I and light reaction II from red algae (marine members of Florideae and Bangiales) were measured with 550 nm (light 2) or 699 nm (light 1) background light, using a Teflon-covered platinum electrode for O2 measurement. Care was taken to ensure that maximum enhancement was reached by the background light.The action spectra of light reaction I, we found under these conditions, are very similar to the thallus absorption, whilst the action spectra of light reaction II show, besides strong bands of the phycobilins, only minor bands of chlorophyll a, which account for only 10–20% of the total chlorophyll.The spectra are discussed on the basis of two main types of models of energy distribution over both photosynthetic systems. If this distribution is considered to be invariable (models 1a and b), one has to assume that almost exactly half of the total chlorophyll is not involved in the supply of the non-cyclic electron transport with excitation energy. This part, however, has to be thought of as incorporated in the thylakoid membrane in a similar manner to the chlorophyll in photosystem I. However, if one supposes an almost complete equilibration in the energy distribution over both systems as long as the primary absorption in photosystem II prevails (models 2a and b), there is no need for the assumption of such photosynthetically ‘inactive’ or less active chlorophyll. Some evidence is shown that strongly supports model 2.  相似文献   

15.
Glutaredoxins (Grxs) are glutathione-dependent oxidoreductases that belong to the thioredoxin superfamily catalyzing thiol-disulfide exchange reactions via active site cysteine residues. Focusing on the human dithiol glutaredoxins having a C-X-Y-C active site sequence motif, the redox potentials of hGrx1 and hGrx2 were determined to be -232 and -221 mV, respectively, using a combination of redox buffers, protein-protein equilibrium and thermodynamic linkage. In addition, a nonactive site disulfide was identified between Cys28 and Cys113 in hGrx2 using redox buffers and chemical digestion. This disulfide confers nearly five kcal mol(-1) additional stability by linking the C-terminal helix to the bulk of the protein. The redox potential of this nonactive site disulfide was determined to be -317 mV and is thus expected to be present in all but the most reducing conditions in vivo. As all human glutaredoxins contain additional nonactive site cysteine residues, a full phylogenetic analysis was performed to help elucidate their structural and functional roles. Three distinct groups were found: Grx1, Grx2, and Grx5, the latter representing a highly conserved group of monothiol glutaredoxins having a C-G-F-S active site sequence, with clear homologs from bacteria to human. Grx1 and Grx2 diverged from a common ancestor before the origin of vertebrates, possibly even earlier in animal evolution. The highly stabilizing nonactive site disulfide observed in hGrx2 is found to be a conserved feature within the deuterostomes and appears to be the only additional conserved intramolecular disulfide within the glutaredoxins.  相似文献   

16.
John H. Golbeck  Bessel Kok 《BBA》1979,547(2):347-360
The primary photochemical quencher Q and the secondary electron acceptor pool in Photosystem II have been titrated. We used particles of Scenedesmus mutant No. 8 that lack System I and allowed the system to equilibrate with external redox mediators in darkness prior to measurement of the fluorescence rise curve.The titration of Q, as indicated by the dark level of Fi, occurs in two discrete steps. The high-potential component (Qh) has a midpoint potential of +68 mV (pH 7.2) and accounts for ~67% of Q. The pH sensitivity of the midpoint potential is ?60 mV, indicating the involvement of 1 H+e. The low-potential component (Q1) accounts for the remaining 33% of Q and shows a midpoint potential near?300 mV (pH 7.2).The plastoquinone pool, assayed as the half-time of the fluorescence rise curve, titrates as a single component with a midpoint potential 30–40 mV more oxidizing than that of Qh, i.e., at 106 mV (pH 7.2). The Em shows a pH sensitivity of ?60 mV/pH unit, indicating the involvement of 1 H+e. The observation that all 12–14 electron equivalents in the pool titrate as a single component indicates that the heterogeneity otherwise observed in the secondary acceptor system is a kinetic rather than a thermodynamic property.Illumination causes peculiar, and as yet unclarified, changes of both Q and the secondary pool under anaerobic conditions that are reversed by oxygen.  相似文献   

17.
A. Telfer  J. Barber  P. Heathcote  M.C.W. Evans 《BBA》1978,504(1):153-164
1. Photosystem I particles enriched in P-700 prepared by Triton X-100 treatment of chloroplasts show a light-induced increase in fluorescence yield of more than 100% in the presence of dithionite but not in its absence.2. Steady state light maintains the P-700, of these particles, in the oxidised state when ascorbate is present but in the presence of dithionite only a transient oxidation occurs.3. EPR data show that, in these particles, the primary electron acceptor (X) is maintained in the reduced state by light at room temperature only when the dithionite is also present. In contrast, the secondary electron acceptors are reduced in the dark by dithionite.4. Fluorescence emission and excitation spectra and fluorescence lifetime measurements for the constant and variable fluorescence indicate a heterogeneity of the chlorophyll in these particles.5. It is concluded that the variable fluorescence comes from those chlorophylls which can transfer their energy to the reaction centre and that the states PX and P+X are more effective quenchers of chlorophyll fluorescence than PX?, where P is P-700.  相似文献   

18.
We have previously measured the individual content of immunoreactive vasopressin (AVP), oxytocin (OT) and vasotocin (AVT) in 155 human pineal glands, and report here identification and measurement of the neurophysin (Np) content of the same glands, using specific homologous human neurophysin I (HNp I) and neurophysin II (HNp II) radioimmunoassays. Median values for HNp I were for men 47 ng/gland (range, 5 to 1360) and for women 24 ng/gland (range, 5 to 1000); median values for HNp II were respectively 7 ng/gland (range, 2 to 191) and 15 ng/gland (range, 2 to 356) with no significant difference between men and women for HNp I and HNp II but a significant difference (p<0.001) between HNp I and HNp II for both sexes. Gel filtration showed that pineal neurophysins were eluted at the same volume as both standard Np and Np from human posthypophyses used as controls. HNp I correlated both with AVP (rs=0.54 for men and 0.55 for women) and OT (rs=0.86 for men and 0.57 for women) but not with AVT, while HNp II correlated with AVP (rs=0.52 for men and 0.53 for women) and OT (rs=0.92 for men and 0.50 for women) but not with AVT. This study thus confirms the presence of two neurophysins in the human pineal gland and further indicates that they are related to AVP and OT concentrations in the same gland. The results also imply, however, that the presence of immunoreactive AVT (more probably a closely related peptide) is independent of the neurophysins.  相似文献   

19.
Action spectra for photosystem II (PSII)-driven oxygen evolution and of photosystem I (PSI)-mediated H2 photoproduction and photoinhibition of respiration were used to determine the participation of chlorophyll (Chl) a/b-binding Pcb proteins in the functions of pigment apparatus of Prochlorothrix hollandica. Comparison of the in situ action spectra with absorption spectra of PSII and PSI complexes isolated from the cyanobacterium Synechocystis 6803 revealed a shoulder at 650 nm that indicated presence of Chl b in the both photosystems of P. hollandica. Fitting of two action spectra to absorption spectrum of the cells showed a chlorophyll ratio of 4:1 in favor of PSI. Effective antenna sizes estimated from photochemical cross-sections of the relevant photoreactions were found to be 192 ± 28 and 139 ± 15 chlorophyll molecules for the competent PSI and PSII reaction centers, respectively. The value for PSI is in a quite good agreement with previous electron microscopy data for isolated Pcb-PSI supercomplexes from P. hollandica that show a trimeric PSI core surrounded by a ring of 18 Pcb subunits. The antenna size of PSII implies that the PSII core dimers are associated with ∼ 14 Pcb light-harvesting proteins, and form the largest known Pcb-PSII supercomplexes.  相似文献   

20.
The paper describes molecular dynamics (MD) simulations on the crystal structures of the Iβ and II phases of cellulose. Structural proposals for each of these were made in the 1970s on the basis of X-ray diffraction data. However, due to the limited resolution of these data some controversies remained and details on hydrogen bonding could not be directly obtained. In contrast to structure factor amplitudes in X-ray diffraction, energies, as obtained from MD simulations, are very sensitive to the positions of the hydroxyl hydrogen atoms. Therefore the latter technique is very suitable for obtaining such structural details. MD simulations of the Iβ phase clearly shows preference for one of the two possible models in which the chains are packed in a parallel orientation. Only the parallel-down mode (in the definition of Gardner and Blackwell (1974) J Biopolym 13: 1975-2001) presents a stable structure. The hydrogen bonding consists of two intramolecular hydrogen bonds parallel to the glycosidic linkage for both chains, and two intralayer hydrogen bonds. The layers are packed hydrophobically. All hydroxymethyl group are positioned in the tg conformation. For the cellulose II form it was found that, in contrast to what seemed to emerge from the X-ray fibre diffraction data, both independent chains had the gt conformation. This idea already existed because of elastic moduli calculations and 13C-solid state NMR data. Recently, the structure of cellotetraose was determined. There appear to be a striking similarity between the structure obtained from the MD simulations and this cellotetraose structure in terms of packing of the two independent molecules, the hydrogen bonding network and the conformations of the hydroxymethyl group, which were also gt for both molecules. The structure forms a 3D hydrogen bonded network, and the contribution from electrostatics to the packing is more pronounced than in case of the Iβ structure. In contrast to what is expected, in view of the irreversible transition of the cellulose I to II form, the energies of the Iβ form is found to be lower than that of II by 1 kcal mol-1 per cellobiose. This revised version was published online in November 2006 with corrections to the Cover Date.  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