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.
Duodenal alkaline phosphatase of juvenile (11-day-old) mice, like other non-specific alkaline phosphatases, has the ability to hydrolyse PP(i). When a constant Mg(2+)/PP(i) concentration ratio is maintained, plots of velocity as a function of PP(i) concentration are consistent with Michaelis-Menten kinetics. Mg(2+) activates pyrophosphate hydrolysis and maximal activity is obtained at a constant Mg(2+)/PP(i) concentration ratio of 0.66. At higher ratios there is strong inhibition. At constant concentrations of Mg(2+) and increasing concentrations of PP(i), the velocity-substrate (PP(i)) concentration plots show sigmoidal dependence. By assuming that the true substrate is MgP(2)O(7) (2-) complex, and using complexity constants, the concentrations of free Mg(2+), Mg(2)P(2)O(7) and MgP(2)O(7) (2-) were calculated in assay mixtures ranging in PP(i) concentration from 0.1 to 2.5mm and in total Mg(2+) concentration from 0.6 to 2.6mm. From these data, the concentrations of added Mg(2+) and PP(i) in the assay mixtures were selected so that the velocity could be measured (1) at three fixed concentrations of free Mg(2+) ions with varied concentrations of MgP(2)O(7) (2-) and (2) at four fixed concentrations of Mg(2)P(2)O(7) with varied concentrations of MgP(2)O(7) (2-). Lineweaver-Burk and Hill plots from these data showed that the inhibition is caused by free Mg(2+) ions, of a mixed type and consistent with Michaelis-Menten kinetics. The sigmoidal dependence observed between velocity and PP(i) concentration at constant concentration of total Mg(2+) is therefore not due to allosteric inhibition. It is due to a combined effect of (1) inhibition by free Mg(2+) ions, (2) depletion of the true substrate, MgP(2)O(7) (2-), owing to the formation of Mg(2)P(2)O(7) and (3) the manner in which the concentrations of these three molecular or ionic species change when PP(i) concentration is increased maintaining the total Mg(2+) concentration constant.  相似文献   

2.
Enhanced absorption is observed in the (15)N NMR spectra of (15)NO(-)(3) during decomposition of peroxynitrite and the peroxynitrite-CO(2) adduct at pH 5.25, indicating the formation of (15)NO(-)(3) in radical pairs [(15)NO(*)(2), HO(*)] and [(15)NO(*)(2), CO(*-)(3)]. During the reaction of peroxynitrite and the peroxynitrite-CO(2) adduct with L-tyrosine, the (15)N NMR signal of the nitration product 3-nitrotyrosine exhibits emission showing a radical pathway of its formation. The nuclear polarization is built up in radical pairs [(15)NO(*)(2), tyr(*)] generated by free radical encounters of nitrogen dioxide and tyrosinyl radicals. The (15)N NMR signal of (15)NO(-)(2) formed during reaction of peroxynitrite with L-tyrosine appears in emission. It is concluded that tyrosinyl radicals are generated by reaction of nitrogen dioxide with L-tyrosine. In contrast to this, (15)NO(-)(2) does not show (15)N chemically induced dynamic nuclear polarization (CIDNP) during reaction of the peroxynitrite-CO(2) adduct with L-tyrosine, indicating a different reaction mechanism, which is assumed to be a hydrogen transfer between CO(*-)(3) and L-tyrosine. Emission is also observed in the (15)N NMR signals of 2-nitro-4-fluorophenol, 3-nitro-4-hydroxyphenylacetic acid, 2-nitrophenol, and 4-nitrophenol during reaction of 4-fluorophenol, 4-hydroxyphenylacetic acid, and phenol with peroxynitrite and the peroxynitrite-CO(2) adduct. 3-Nitro-4-hydroxyphenylacetic acid is also observed in emission during reaction of phenylacetic acid with peroxynitrite, but is not formed with the peroxynitrite-CO(2) adduct. The magnitude of the (15)N CIDNP effect during reaction of peroxynitrite with 4-fluorophenol and of the peroxynitrite-CO(2) adduct with 4-fluorophenol and phenol is determined. It excludes the occurrence of nonradical reactions. Only weak emission signals are observed during the reaction of peroxynitrite with phenol in (15)NO(-)(2), 2-nitrophenol, and 4-nitrophenol. 2-Nitrophenol is only formed in traces, and 4-nitrophenol is only formed in higher yields. The latter might be generated in part via a nonradical pathway.  相似文献   

3.
Bacterioferritin (EcBFR) of Escherichia coli is an iron-mineralizing hemoprotein composed of 24 identical subunits, each containing a dinuclear metal-binding site known as the "ferroxidase center." The chemistry of Fe(II) binding and oxidation and Fe(III) hydrolysis using H(2)O(2) as oxidant was studied by electrode oximetry, pH-stat, UV-visible spectrophotometry, and electron paramagnetic resonance spin trapping experiments. Absorption spectroscopy data demonstrate the oxidation of two Fe(II) per H(2)O(2) at the ferroxidase center, thus avoiding hydroxyl radical production via Fenton chemistry. The oxidation reaction with H(2)O(2) corresponds to [Fe(II)(2)-P](Z) + H(2)O(2) --> [Fe(III)(2)O-P](Z) + H(2)O, where [Fe(II)(2)-P](Z) represents a diferrous ferroxidase center complex of the protein P with net charge Z and [Fe(III)(2)O-P](Z) a micro-oxo-bridged diferric ferroxidase complex. The mineralization reaction is given by 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2FeOOH((core)) + 4H(+), where two Fe(II) are again oxidized by one H(2)O(2). Hydrogen peroxide is shown to be an intermediate product of dioxygen reduction when O(2) is used as the oxidant in both the ferroxidation and mineralization reactions. Most of the H(2)O(2) produced from O(2) is rapidly consumed in a subsequent ferroxidase reaction with Fe(II) to produce H(2)O. EPR spin trapping experiments show that the presence of EcBFR greatly attenuates the production of hydroxyl radical during Fe(II) oxidation by H(2)O(2), consistent with the ability of the bacterioferritin to facilitate the pairwise oxidation of Fe(II) by H(2)O(2), thus avoiding odd electron reduction products of oxygen and therefore oxidative damage to the protein and cellular components through oxygen radical chemistry.  相似文献   

4.
1. Pigeon heart mitochondria produce H(2)O(2) at a maximal rate of about 20nmol/min per mg of protein. 2. Succinate-glutamate and malate-glutamate are substrates which are able to support maximal H(2)O(2) production rates. With malate-glutamate, H(2)O(2) formation is sensitive to rotenone. Endogenous substrate, octanoate, stearoyl-CoA and palmitoyl-carnitine are by far less efficient substrates. 3. Antimycin A exerts a very pronounced effect in enhancing H(2)O(2) production in pigeon heart mitochondria; 0.26nmol of antimycin A/mg of protein and the addition of an uncoupler are required for maximal H(2)O(2) formation. 4. In the presence of endogenous substrate and of antimycin A, ATP decreases and uncoupler restores the rates of H(2)O(2) formation. 5. Reincorporation of ubiquinone-10 and ubiquinone-3 to ubiquinone-depleted pigeon heart mitochondria gives a system in which H(2)O(2) production is linearly related to the incorporated ubiquinone. 6. The generation of H(2)O(2) by pigeon heart mitochondria in the presence of succinate-glutamate and in metabolic state 4 has an optimum pH value of 7.5. In states 1 and 3u, and in the presence of antimycin A and uncoupler, the optimum pH value is shifted towards more alkaline values. 7. With increase of the partial pressure of O(2) to the hyperbaric region the formation of H(2)O(2) is markedly increased in pigeon heart mitochondria and in rat liver mitochondria. With rat liver mitochondria and succinate as substrate in state 4, an increase in the pO(2) up to 1.97MPa (19.5atm) increases H(2)O(2) formation 10-15-fold. Similar pO(2) profiles were observed when rat liver mitochondria were supplemented either with antimycin A or with antimycin A and uncoupler. No saturation of the system with O(2) was observed up to 1.97MPa (19.5atm). By increasing the pO(2) to 1.97MPa (19.5atm), H(2)O(2) formation in pigeon heart mitochondria with succinate as substrate increased fourfold in metabolic state 4, with antimycin A added the increase was threefold and with antimycin A and uncoupler it was 2.5-fold. In the last two saturation of the system with oxygen was observed, with an apparent K(m) of about 71kPa (0.7-0.8atm) and a V(max.) of 12 and 20nmol of H(2)O(2)/min per mg of protein. 8. It is postulated that in addition to the well-known flavin reaction, formation of H(2)O(2) may be due to interaction with an energy-dependent component of the respiratory chain at the cytochrome b level.  相似文献   

5.
Over the last decade, an expanding diversity of secreted phospholipases A(2) (sPLA(2)s) has been identified in mammals. Here, we report the cloning in mice of three additional sPLA(2)s called mouse group IIE (mGIIE), IIF (mGIIF), and X (mGX) sPLA(2)s, thus giving rise to eight distinct sPLA(2)s in this species. Both mGIIE and mGIIF sPLA(2)s contain the typical cysteines of group II sPLA(2)s, but have relatively low levels of identity (less than 51%) with other mouse sPLA(2)s, indicating that these enzymes are novel group II sPLA(2)s. However, a unique feature of mGIIF sPLA(2) is the presence of a C-terminal extension of 23 amino acids containing a single cysteine. mGX sPLA(2) has 72% identity with the previously cloned human group X (hGX) sPLA(2) and displays similar structural features, making it likely that mGX sPLA(2) is the ortholog of hGX sPLA(2). Genes for mGIIE and mGIIF sPLA(2)s are located on chromosome 4, and that of mGX sPLA(2) on chromosome 16. Northern and dot blot experiments with 22 tissues indicate that all eight mouse sPLA(2)s have different tissue distributions, suggesting specific functions for each. mGIIE sPLA(2) is highly expressed in uterus, and at lower levels in various other tissues. mGIIF sPLA(2) is strongly expressed during embryogenesis and in adult testis. mGX sPLA(2) is mostly expressed in adult testis and stomach. When the cDNAs for the eight mouse sPLA(2)s were transiently transfected in COS cells, sPLA(2) activity was found to accumulate in cell medium, indicating that each enzyme is secreted and catalytically active. Using COS cell medium as a source of enzymes, pH rate profile and phospholipid headgroup specificity of the novel sPLA(2)s were analyzed and compared with the other mouse sPLA(2)s.  相似文献   

6.
Agents mobilising Ca(2+) from the endoplasmic reticulum are known to activate apoptosis. Whatever means are used, the release of Ca(2+) is often followed by a store-dependent entry of Ca(2+). Whether apoptosis is triggered by the depletion of the stores or by the subsequent store-dependent entry of Ca(2+) is still a matter of controversy. Here we studied apoptosis in CHO cells transfected with the rat neurotensin (NT) receptor, in which the store-dependent entry of Ca(2+) is abolished by repressing the transient receptor potential channel 2 (TRPC2) by an antisense oligonucleotide strategy (TRPC2(-) cells) [Cell Calcium 30 (2001) 157]. When stimulated with thapsigargin (TG), apoptosis occurred in both TRPC2(+) and TRPC2(-) cells but 12h earlier in TRPC2(+) cells, suggesting that store-dependent entry of Ca(2+) can accelerate the process. The expression and localisation of caspase-12, an enzyme that has been involved in the apoptosis triggered by a stress on the endoplasmic reticulum, was not different in TRPC2(+) and TRPC2(-) cells. On the contrary, the expression of GADD153 (Growth Arrest and DNA Damage inducible gene 153) triggered by TG treatment depended on external Ca(2+) and occurred earlier in TRPC2(+) than in TRPC2(-) cells. In these cells, we also noted the presence of K(+) channels activated by Ca(2+) (K(Ca) channels). Stimulation of TRPC2(+) cells with TG or with NT triggered a long sustained K(+) current, parallel to [Ca(2+)](i) transients, and resulting in a sustained hyperpolarisation of the cell membrane. K(+) current and hyperpolarisation were transient and not sustained in TRPC2(-) cells. Inhibition of K(Ca) channels with charybdotoxin dramatically reduced the K(+) current and also significantly brought down the level of apoptosis, suggesting that a prolonged efflux of K(+) could be involved in the apoptosis process. We conclude that in CHO cells, store-dependent entry of Ca(2+) can accelerate apoptosis by accelerating the expression of GADD153 and by inducing a prolonged efflux of K(+) out of the cell.  相似文献   

7.
Kinetic analysis and mechanistic aspects of autoxidation of catechins   总被引:3,自引:0,他引:3  
A peroxidase-based bioelectrochemical sensor of hydrogen peroxide (H(2)O(2)) and a Clark-type oxygen electrode were applied to continuous monitoring and kinetic analysis of the autoxidation of catechins. Four major catechins in green tea, (-)-epicatechin, (-)-epicatechin gallate, (-)-epigallocatechin, and (-)-epigallocatechin gallate, were used as model compounds. It was found that dioxygen (O(2)) is quantitatively reduced to H(2)O(2). The initial rate of autoxidation is suppressed by superoxide dismutase and H(+), but is independent of buffer capacity. Based on these results, a mechanism of autoxidation is proposed; the initial step is the one-electron oxidation of the B ring of catechins by O(2) to generate a superoxide anion (O(2)(*-)) and a semiquinone radical, as supported in part by electron spin resonance measurements. O(2)(*-) works as a stronger one-electron oxidant than O(2) against catechins and is reduced to H(2)O(2). The semiquinone radical is more susceptible to oxidation with O(2) than fully reduced catechins. The autoxidation rate increases with pH. This behavior can be interpreted in terms of the increase in the stability of O(2)(*-) and the semiquinone radical with increasing pH, rather than the acid dissociation of phenolic groups. Cupric ion enhances autoxidation; most probably it functions as a catalyst of the initial oxidation step of catechins. The product cuprous ion can trigger a Fenton reaction to generate hydroxyl radical. On the other hand, borate ion suppresses autoxidation drastically, due to the strong complex formation with catechins. The biological significance of autoxidation and its effectors are also discussed.  相似文献   

8.
4-Pyrimidinone ribofuranoside (H(2)o(4)U) and 4-pyrimidinone 2'-deoxyribofuranoside (dH(2)o(4)U) were synthesized by the oxidative desulfurization of parent 2-thiouracil nucleosides with m-chloroperbenzoic acid. The crystal structures of H(2)o(4)U and dH(2)o(4)U and their conformations in solution were determined and compared with corresponding 2-thiouracil and uracil nucleosides. The absence of a large 2-thiocarbonyl/2-carbonyl group in the nucleobase moiety results in C2'-endo puckering of the ribofuranose ring (S conformer) in the crystal structure of H(2)o(4)U, which is not typical of RNA nucleosides. Interestingly, the hydrogen bonding network in the crystals of dH(2)o(4)U stabilizes the sugar moiety conformation in the C3'-endo form (N conformer), rarely found in DNA nucleosides. In aqueous solution, dH(2)o(4)U reveals a similar population of the C2'-endo conformation (65%) to that of 2'-deoxy-2-thiouridine (62%), while the 62% population of the S conformer for H(2)o(4)U is significantly different from that of the parent 2-thiouridine, for which the N conformer is dominant (71%). Such a difference may be of biological importance, as the desulfurization process of natural tRNA 2-thiouridines may occur under conditions of oxidative stress in the cell and may influence the decoding process.  相似文献   

9.
Apoptosis was studied under conditions that mimic the steady state of H(2)O(2) in vivo. This is at variance with previous studies involving a bolus addition of H(2)O(2), a procedure that disrupts the cellular homeostasis. The results allowed us to define three phases for H(2)O(2)-induced apoptosis in Jurkat T-cells with reference to cytosolic steady state concentrations of H(2)O(2) [(H(2)O(2))(ss)]: (H(2)O(2))(ss) values below 0.7 microM elicited no effects; (H(2)O(2))(ss) approximately 0.7-3 microM induced apoptosis; and (H(2)O(2))(ss) > 3 microM yielded no additional apoptosis and a gradual shift towards necrosis as the mode of cell death were observed. H(2)O(2)-induced apoptosis was not affected by either BCNU, an inhibitor of glutathione reductase, or diamide, a compound that reacts both with low-molecular weight and protein thiols, or selenols. Glutathione depletion, accomplished by incubating cells either with buthionine sulfoximine or in cystine-free medium, rendered cells more sensitive to H(2)O(2)-induced apoptosis, but did not change the threshold and saturating concentrations of H(2)O(2) that induced apoptosis. Two unrelated metal chelators, desferrioxamine and dipyridyl, strongly protected against H(2)O(2)-induced apoptosis. It may be concluded that, under conditions of H(2)O(2) delivery that mimic in vivo situations, the oxidative event that triggers the induction of apoptosis by H(2)O(2) is a Fenton-type reaction and is independent of the thiol or selenium states of the cell.  相似文献   

10.
The hydrolysis mechanism of CS(2) was studied using density functional theory. By analyzing the structures of the reactant, transition states, intermediates, and products, it can be concluded that the hydrolysis of CS(2) occurs via two mechanisms, one of which is a two-step mechanism (CS(2) first reacts with an H(2)O, leading to the formation of the intermediate COS, then COS reacts with another H(2)O, resulting in the formation of H(2)S and CO(2)). The other is a one-step mechanism, where CS(2) reacts with two H(2)O molecules continuously, leading to the formation of H(2)S and CO(2). By analyzing the thermodynamics and the change in the kinetic function, it can be concluded that the rate-determining step involves H and OH in H(2)O attacking S and C in CS(2), respectively, causing the C=S double bond to change into a single bond. The two mechanisms are competitive. When performing the hydrolysis of CS(2) with a catalyst, the optimal temperature is below 252°C.  相似文献   

11.
Yamazaki S  Morioka C  Itoh S 《Biochemistry》2004,43(36):11546-11553
Tyrosinase is a copper monooxygenase containing a coupled dinuclear copper active site (type-3 copper), which catalyzes oxygenation of phenols (phenolase activity) as well as dehydrogenation of catechols (catecholase activity) using O(2) as the oxidant. In this study, catalase activity (conversion of H(2)O(2) to (1/2)O(2) and H(2)O) and peroxygenase activity (H(2)O(2)-dependent oxygenation of substrates) of mushroom tyrosinase have been examined kinetically by using amperometric O(2) and H(2)O(2) sensors. The catalase activity has been examined by monitoring the initial rate of O(2) production from H(2)O(2) in the presence of a catalytic amount of tyrosinase in 0.1 M phosphate buffer (pH 7.0) at 25 degrees C under initially anaerobic conditions. It has been found that the catalase activity of mushroom tyrosinase is three-order of magnitude greater than that of mollusk hemocyanin. The higher catalase activity of tyrosinase could be attributed to easier accessibility of H(2)O(2) to the dinuclear copper site of tyrosinase. Mushroom tyrosinase has also been demonstrated for the first time to catalyze oxygenation reaction of phenols with H(2)O(2) (peroxygenase activity). The reaction has been investigated kinetically by monitoring the H(2)O(2) consumption rate in 0.5 M borate buffer (pH 7.0) under aerobic conditions. Similarity of the substituent effects of a series of p-substituted phenols in the peroxygenase reaction with H(2)O(2) to those in the phenolase reaction with O(2) as well as the absence of kinetic deuterium isotope effect with a perdeuterated substrate (p-Cl-C(6)D(4)OH vs p-Cl-C(6)H(4)OH) clearly demonstrated that the oxygenation mechanisms of phenols in both systems are the same, that is, the electrophilic aromatic substitution reaction by a (micro-eta(2):eta(2)-peroxo)dicopper(II) intermediate of oxy-tyrosinase.  相似文献   

12.
Nitrite and SCN(-) in saliva can mixes with H(2)O(2) in the stomach. The mixing can result in the formation of ONOOH. It is not yet known how salivary SCN(-) reacts with ONOOH. An objective of the present study was to elucidate the reaction between ONOOH and SCN(-). In nitrite/H(2)O(2) systems at pH 2, SCN(-) inhibited the consumption of nitrite and the formation of O(3)(-). SCN(-) enhanced the decomposition of ONOOH and H(2)O(2) in HNO(2)/H(2)O(2) systems. Accompanying the reactions, sulfate was formed, suggesting that ONOOH oxidized SCN(-). SCN(-) inhibited the nitration of phenolics induced by HNO(2)/H(2)O(2). The inhibition is discussed taking SCN(-)-dependent reduction of ONOOH to HNO(2) into consideration. SCN(-) also inhibited H(2)O(2)-induced consumption of nitrite and nitration of phenolics in acidified saliva. The result obtained in this study suggests that salivary SCN(-) can reduce ONOOH to O(2)(-)/HNO(2) inhibiting nitrating reactions in the stomach.  相似文献   

13.
The thionucleoside 2-thiocytidine (C2S) occurs in nature in transfer RNAs; it receives attention in diverse fields like drug research and nanotechnology. By potentiometric pH titrations we measured the acidity constants of H(C2S)(+) and the stability constants of the M(C2S)(2+) and M(C2S-H)(+) complexes (M(2+) = Zn(2+), Cd(2+)), and we compared these results with those obtained previously for its parent nucleoside, cytidine (Cyd). Replacement of the (C2)=O unit by (C2)=S facilitates the release of the proton from (N3)H(+) in H(C2S)(+) (pK (a) = 3.44) somewhat, compared with H(Cyd)(+) (pK (a) = 4.24). This moderate effect of about 0.8 pK units contrasts with the strong acidification of about 4 pK units of the (C4)NH(2) group in C2S (pK (a) = 12.65) compared with Cyd (pK (a) approximately 16.7); the reason for this result is that the amino-thione tautomer, which dominates for the neutral C2S molecule, is transformed upon deprotonation into the imino-thioate form with the negative charge largely located on the sulfur. In the M(C2S)(2+) complexes the (C2)S group is the primary binding site rather than N3 as is the case in the M(Cyd)(2+) complexes, though owing to chelate formation N3 is to some extent still involved in metal ion binding. Similarly, in the Zn(C2S-H)(+) and Cd(C2S-H)(+) complexes the main metal ion binding site is the (C2)S(-) unit (formation degree above 99.99% compared with that of N3). However, again a large degree of chelate formation with N3 must be surmised for the M(C2S-H)(+) species in accord with previous solid-state studies of related ligands. Upon metal ion binding, the deprotonation of the (C4)NH(2) group (pK (a) = 12.65) is dramatically acidified (pK (a) approximately 3), confirming the very high stability of the M(C2S-H)(+) complexes. To conclude, the hydrogen-bonding and metal ion complex forming capabilities of C2S differ strongly from those of its parent Cyd; this must have consequences for the properties of those RNAs which contain this thionucleoside.  相似文献   

14.
Modulation of mitochondrial function by hydrogen peroxide   总被引:16,自引:0,他引:16  
During normal cellular metabolism, mitochondrial electron transport results in the formation of superoxide anion (O(2)) and subsequently hydrogen peroxide (H(2)O(2)). Because H(2)O(2) increases in concentration under certain physiologic and pathophysiologic conditions and can oxidatively modify cellular components, it is critical to understand the response of mitochondria to H(2)O(2). In the present study, treatment of isolated rat heart mitochondria with H(2)O(2) resulted in a decline and subsequent recovery of state 3 NADH-linked respiration. Alterations in NADH levels induced by H(2)O(2) closely paralleled changes in the rate of state 3 respiration. Assessment of electron transport chain complexes and Krebs cycle enzymes revealed that alpha-ketoglutarate dehydrogenase (KGDH), succinate dehydrogenase (SDH), and aconitase were susceptible to H(2)O(2) inactivation. Of particular importance, KGDH and SDH activity returned to control levels, concurrent with the recovery of state 3 respiration. Inactivation is not because of direct interaction of H(2)O(2) with KGDH and SDH. In addition, removal of H(2)O(2) alone is not sufficient for reactivation. Enzyme activity does not recover unless mitochondria remain intact. The sensitivity of KGDH and SDH to H(2)O(2)-mediated inactivation and the reversible nature of inactivation suggest a potential role for H(2)O(2) in the regulation of KGDH and SDH.  相似文献   

15.
Reactions of human oxyhemoglobin A with iron(II) compounds have been investigated. Human oxyhemoglobin (HbO2) reacts with aquopentacyanoferrate(II), Fe(II)(CN)5H2O3-, to yield hydrogen peroxide, aquomethemoglobin and Fe(III)(CN)5H2O2-. The reaction follows a second order rate law, first order in the pentacyanide and in HbO2. Since reaction rates are lower in the presence of catalase, the H2O2 produced must promote metHb formation in reactions independent of pentacyanide. Changes in concentrations of effectors (e.g. H+, inositol hexaphosphate, Cl-, and Zn2+), alkylation of beta-93 cysteine with N-ethylmaleimide, and substitution at distal histidine (as in Hb Zurich with beta-63 His----Arg) in each case can markedly affect pentacyanide reaction rates demonstrating a fine control of rates by protein structure. Hexacyanoferrate(II) (ferrocyanide) reacts with HbO2 to produce cyano-metHb as well as aquo-metHb but the reaction with the hexacyanide is much slower than with the aquopentacyanide. Iron(II) EDTA converts HbO2 to deoxy-Hb with no evidence for formation of metHb as an intermediate. These findings support a mechanism in which the pentacyanide anion reacts directly with coordinated dioxygen. One-electron transfers to O2 from both pentacyanide iron(II) and heme iron(II) result in the formation of a mu-peroxo intermediate, HbFe(III)-O-O-Fe(III) (CN)5(3-). Hydrolysis of this intermediate yields metHb . H2O, H2O2, and FeIII(CN)5H2O2-. The reaction of HbO2 with Fe(CN)6(4-) must follow an outer sphere electron transfer mechanism. However, the very slow rate that is seen with Fe(CN)6(4-) could arise entirely from the pentacyanide produced from loss of one cyanide ligand from the hexacyanide. Fe(II)EDTA reacts rapidly with free O2 in solution but can not interact directly with the heme-bound O2 of HbAO2. The dynamic character of the O2 binding sites apparently permits access of the Fe2+ of the pentacyanide to coordinated dioxygen but the protein structure is not sufficiently flexible to allow the larger Fe2+EDTA molecule to react with bound O2. It is necessary for maintenance of the oxygen transport function of the red cell for reductants such as the methemoglobin reductase system, glutathione, and ascorbate to be able to reduce metHb to deoxy-Hb. It is also important for these reductants to be unable to donate an electron to HbO2 to yield H2O2 and metHb. Thus, a mechanistic requirement for the delivery of one-electron directly to the dioxygen ligand, if peroxide is to be produced, enables the protein to protect the oxygenated species from those electron donors normally present in the cell by denying these reductants steric access to coordinated O2.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
The recognition reactions between a synthetic disaccharide alpha-Kdo-(2-->4)-alpha-Kdo-(2-->O)-allyl and two monoclonal antibodies (mAbs) were studied by NMR, yielding two distinct bound conformations of the carbohydrate ligand. One mAb, S23-24, recognizes the disaccharides alpha-Kdo-(2-->4)-alpha-Kdo-(2-->O)-allyl and alpha-Kdo-(2-->8)-alpha-Kdo-(2-->O)-allyl with similar affinities, whereas mAb S25-2 binds to the disaccharide alpha-Kdo-(2-->8)-alpha-Kdo-(2-->O)-allyl with an approximately 10-fold higher affinity than to the disaccharide alpha-Kdo-(2-->4)-alpha-Kdo-(2-->O)-allyl. Compared to S25-2, S23-24 binds to alpha-Kdo-(2-->4)-alpha-Kdo-(2-->O)-allyl with an approximately 50-fold increased affinity. We used NMR experiments that are based on the transferred NOE effect, specifically, trNOESY, trROESY, QUIET-trNOESY, and MINSY experiments, to show that the (2-->8)-specific mAb, S25-2, stabilizes a conformation of the alpha-(2-->4)-linked disaccharide that is not highly populated in solution. S23-24 recognizes two conformations of alpha-Kdo-(2-->4)-alpha-Kdo-(2-->O)-allyl, one that is highly populated in aqueous solution and another conformation that is similar to the one bound by S25-2. This is the first example where it is experimentally shown that a carbohydrate ligand may adopt different bioactive conformations upon interaction with mAbs with different fine specificities. Our NMR studies indicate that a careful examination of spin diffusion is critical for the analysis of bioactive conformations of carbohydrate ligands.  相似文献   

17.
Alkaline solutions of hydrogen peroxide partially delignify wheat straw and other lignocellulosic materials, leaving a cellulosic residue that is highly susceptible to enzymatic digestion by cellulase. The delignification reaction is strongly dependent upon the pH of the reaction mixture, with an optimum at pH 11.5-11.6, pKa for the dissociation H(2)O(2) right harpoon over left harpoon H(+) + HOO(-). The data are consistent with a mechanism in which H(2)O(2) decomposition products such as .OH and O(2) (-)., rather than H(2)O(2) or HOO(-), are the primary lignin oxidizing species. During the course of the delignification reaction, O(2) is evolved from the reaction mixture indicating active H(2)O(2) decomposition. At a given concentration of H(2)O(2), the rate of O(2) evolution is proportional to the amount of lignocellulosic substrate present in the reaction mixture. However, the total amount of O(2) evolved is inversely proportional to the amount of substrate present, indicating that some of the peroxide oxygen becomes incorporated into lignin degradation products. The amount of peroxide oxygen incorporated can range as high as 2 O(2) per lignin C(9) unit, depending upon the initial concentration of lignocellulosic substrate.  相似文献   

18.
Incubation of 3-(2-phenylethyl)-4-methylsydnone (PMS) with liver microsomes from phenobarbital-pretreated rats or with reconstituted cytochrome P-450b results in loss of the enzyme chromophore. Chromophore loss is NADPH-dependent even though the sydnone decomposes by an oxygen- but not enzyme-dependent process to give pyruvic acid and, presumably, the (2-phenylethyl)diazonium cation. N-(2-Phenylethyl)protoporphyrin IX and N-(2-phenylethenyl)protoporphyrin IX have been isolated from the livers of rats treated with PMS. Both deuteriums are retained in the N-(2-phenylethyl) adduct derived from 3-(2-phenyl[1,1-2H]ethyl)-4-methylsydnone, but one deuterium is lost in the N-(2-phenylethenyl) adduct. The N-(2-phenylethyl) to N-(2-phenylethenyl) adduct ratio is increased by deuterium substitution. No spectroscopically detectable intermediates precede chromophore loss in incubations of reconstituted cytochrome P-450b with PMS. Electron paramagnetic resonance (EPR)-spin trapping studies show that carbon radicals are formed in incubations of the sydnones with liver microsomes but by a process that is independent of chromophore destruction. It is proposed that the 2-phenylethyl radical formed by electron transfer to the sydnone-derived (2-phenylethyl)diazonium cation adds to the prosthetic heme group to give the N-(2-phenylethyl) adduct. This alkylation reaction is similar to that observed with (2-phenylethyl)hydrazine. Autoxidation of the Fe-CH(CH2Ph)-N bridged species expected from insertion of 2-phenyldiazoethane into one of the heme Fe-N bonds is proposed to explain the unprecedented introduction of a double bond into the N-(2-phenylethenyl) adduct.  相似文献   

19.
We report here that the Leishmania major ascorbate peroxidase (LmAPX), having similarity with plant ascorbate peroxidase, catalyzes the oxidation of suboptimal concentration of ascorbate to monodehydroascorbate (MDA) at physiological pH in the presence of added H(2)O(2) with concurrent evolution of O(2). This pseudocatalatic degradation of H(2)O(2) to O(2) is solely dependent on ascorbate and is blocked by a spin trap, alpha-phenyl-n-tert-butyl nitrone (PBN), indicating the involvement of free radical species in the reaction process. LmAPX thus appears to catalyze ascorbate oxidation by its peroxidase activity, first generating MDA and H(2)O with subsequent regeneration of ascorbate by the reduction of MDA with H(2)O(2) evolving O(2) through the intermediate formation of O(2)(-). Interestingly, both peroxidase and ascorbate-dependent pseudocatalatic activity of LmAPX are reversibly inhibited by SCN(-) in a concentration dependent manner. Spectral studies indicate that ascorbate cannot reduce LmAPX compound II to the native enzyme in presence of SCN(-). Further kinetic studies indicate that SCN(-) itself is not oxidized by LmAPX but inhibits both ascorbate and guaiacol oxidation, which suggests that SCN(-) blocks initial peroxidase activity with ascorbate rather than subsequent nonenzymatic pseudocatalatic degradation of H(2)O(2) to O(2). Binding studies by optical difference spectroscopy indicate that SCN(-) binds LmAPX (Kd = 100 +/- 10 mM) near the heme edge. Thus, unlike mammalian peroxidases, SCN(-) acts as an inhibitor for Leishmania peroxidase to block ascorbate oxidation and subsequent pseudocatalase activity.  相似文献   

20.
The nature of the acid-volatile selenium in the liver of the male rat   总被引:2,自引:1,他引:1  
1. The properties of rat liver acid-volatile selenium have been compared with those of H(2)Se and (CH(3))(2)Se. 2. In model experiments oxidation-sensitive H(2) (75)Se was trapped quantitatively under anaerobic conditions in 0.1m-AgNO(3), and (CH(3))(2) (75)Se was trapped quantitatively in 8m-HNO(3). The acid-labile selenium of a liver homogenate, and of a microsomal fraction, was found to behave quite unlike (CH(3))(2) (75)Se and in a manner indistinguishable from H(2) (75)Se. 3. It was concluded that the acid-volatile material is certainly not (CH(3))(2)Se and that it is probably H(2)Se. 4. The significance of these findings is discussed in relation to current knowledge about the metabolism and detoxication of selenium, and a scheme is proposed which incorporates this knowledge with recent observations on the interactions between trace amounts of selenium and tocopherol, and the production of acute selenium deficiency by Ag(+) in vitamin E-deficient rats.  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