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 78 毫秒
1.
Although considerably more oxidation-resistant than other P-type ATPases, the yeast PMA1 H+-ATPase of Saccharomyces cerevisiae SY4 secretory vesicles was inactivated by H2O2, Fe2+, Fe- and Cu-Fenton reagents. Inactivation by Fe2+ required the presence of oxygen and hence involved auto-oxidation of Fe2+ to Fe3+. The highest Fe2- (100 μM) and H2O2 (100 mM) concentrations used produced about the same effect. Inactivation by the Fenton reagent depended more on Fe2+ content than on H2O2 concentration, occurred only when Fe2+ was added to the vesicles first and was only slightly reduced by scavengers (mannitol, Tris, NaN3, DMSO) and by chelators (EDTA, EGTA, DTPA, BPDs, bipyridine, 1, 10-phenanthroline). Inactivation by Fe- and Cu- Fenton reagent was the same; the identical inactivation pattern found for both reagents under anaerobic conditions showed that both reagents act via OH·. The lipid peroxidation blocker BHT prevented Fenton-induced rise in lipid peroxidation in both whole cells and in isolated membrane lipids but did not protect the H+-ATPase in secretory vesicles against inactivation. ATP partially protected the enzyme against peroxide and the Fenton reagent in a way resembling the protection it afforded against SH-specific agents. The results indicate that Fe2+ and the Fenton reagent act via metal-catalyzed oxidation at specific metal-binding sites, very probably SH-containing amino acid residues. Deferrioxamine, which prevents the redox cycling of Fe2+, blocked H+-ATPase inactivation by Fe2+ and the Fenton reagent but not that caused by H2O2, which therefore seems to involve a direct non-radical attack. Fe-Fenton reagent caused fragmentation of the H+-ATPase molecule, which, in Western blots, did not give rise to defined fragments bands but merely to smears.  相似文献   

2.
Oxidative reactions can result in the formation of electronically excited species that undergo radiative decay depending on electronic transition from the excited state to the ground state with subsequent ultra‐weak photon emission (UPE). We investigated the UPE from the Fe2+–EDTA (ethylenediaminetetraacetic acid)–AA (ascorbic acid)–H2O2 (hydrogen peroxide) system with a multitube luminometer (Peltier‐cooled photon counter, spectral range 380–630 nm). The UPE, of 92.6 μmol/L Fe2+, 185.2 μmol/L EDTA, 472 μmol/L AA, 2.6 mmol/L H2O2, reached 1217 ± 118 relative light units during 2 min measurement and was about two times higher (P < 0.001) than the UPE of incomplete systems (Fe2+–AA–H2O2, Fe2+–EDTA–H2O2, AA–H2O2) and medium alone. Substitution of Fe2+ with Cr2+, Co2+, Mn2+ or Cu2+ as well as of EDTA with EGTA (ethylene glycol‐bis(β‐aminoethyl ether)‐N,N,N′,N′‐tetraacetic acid) or citrate powerfully inhibited UPE. Experiments with scavengers of reactive oxygen species (dimethyl sulfoxide, mannitol, sodium azide, superoxide dismutase) revealed the dependence of UPE only on hydroxyl radicals. Dimethyl sulfoxide at the concentration of 0.74 mmol/L inhibited UPE by 79 ± 4%. Plant phenolics (ferulic, chlorogenic and caffec acids) at the concentration of 870 μmol/L strongly enhanced UPE by 5‐, 13.9‐ and 46.8‐times (P < 0.001), respectively. It is suggested that augmentation of UPE from Fe2+–EDTA–AA–H2O2 system can be applied for detection of these phytochemicals.  相似文献   

3.
A novel thermoacidophilic iron-reducing Archaeon, strain NA−1, was isolated from a hot fumarole in Manza, Japan. Strain NA-1 could grow autotrophically using H2 or S0 as an electron donor and Fe3+ as an electron acceptor, and also could grow heterotrophically using some organic compounds. Fe3+ and O2 served as electron acceptors for growth. However, S0, NO3 , NO2 , SO4 2−, Mn4+, fumarate, and Fe2O3 did not serve as electron acceptors. The ranges of growth temperature and pH were 60–90°C (optimum: 80°C) and pH 1.0–5.0 (optimum: pH 1.2–1.5), respectively. Cells were nearly regular cocci with an envelope comprised of the cytoplasmic membrane and a single outer S-layer. The crenarchaeal-specific quinone (cardariellaquinone) was detected, and the genomic DNA G + C content was 29.9 mol%. From 16S rDNA analysis, it was determined that strain NA-1 is closely related to Acidianus ambivalens (93.1%) and Acidianus infernus (93.0%). However, differences revealed by phylogenetic and phenotypic analyses clearly show that strain NA-1 represents a new species, Acidianus manzaensis, sp. nov., making it the first identified thermoacidophilic iron-reducing microorganism (strain NA-1T = NBRC 100595 = ATCC BAA 1057). Strain NA-1 has been deposited in the culture collections of the National Institute of Technology and Evolution (NBRC 100595) and American Type Culture Collection (ATCC BAA 1057). The 16S rDNA sequence has been deposited at GenBank under accession number AB182498.  相似文献   

4.
The activity of 1-aminocyclopropane-1-carboxylic acid synthase (ACC synthase, ACS) and the concentrations of superoxide radical (O2−.) and hydrogen peroxide (H2O2) were measured in etiolated mungbean seedlings following their transfer to a growth chamber at 25°C after a 5-h-chilling treatment at 5°C. All of these variables increased dramatically after the transfer, and strong correlations were found between ACS activity and the concentrations of superoxide and H2O2. Exogenous applications of two generators of superoxide radicals, methylviologen (MV) and xanthine–xanthine oxidase (X–XOD), enhanced ACS activity in seedlings, but their effects were inhibited by exogenous applications of specific scavengers of O2−.. However, applications of H2O2 or specific H2O2-scavengers had no significant effects on seedlings ACS activity. The results indicate that O2−. was involved in the chilling-induced increases in ACS activity, but not H2O2. ACS activity peaked ca. 8 h after the transfer, and then declined, but the decline could be counteracted by exogenous applications of specific O2−. scavengers, this suggests that damage was caused by superoxide radicals influencing ACS activity in etiolated mungbean seedlings. Further analysis of changes in two key kinetic parameters of ACS activity—V max (maximum velocity) and K m (the Michaelis constant)—in the seedlings indicated that the presence of O2−. may reduce K m, i.e. increase substrate (S-adenosyl methionine, SAM) affinity. That would be the main mechanism responsible for the observed chilling-induced increases in ACS activity in etiolated mungbean seedlings.  相似文献   

5.
Abstract : Alkaline phosphatase, one of the enzymes responsible for the conversion of phosphocholine into choline, was purified from bovine brain membrane, where the phosphatase is bound as glycosylphosphatidylinositollinked protein, and subjected to oxidative inactivation. The phosphatase activity, based on the hydrolysis of p-nitrophenyl phosphate and phosphocholine, decreased slightly after the exposure to H2O2. Inclusion of Cu2+ in the incubation with 1 mM H2O2 led to a rapid decrease of activity in a time- and concentration-dependent manner. In comparison, the H2O2/Cu2+ system was much more effective than the H2O2/Fe2+ system in inactivating brain phosphatase. In a further study, it was observed that the hydroxy radical scavengers mannitol, ethanol, or benzoate failed to prevent against H2O2/Cu2+-induced inactivation of the phosphatase, excluding the involvement of extraneous hydroxy radicals in metalcatalyzed oxidation. In addition, it was found that both substrates, p-nitrophenyl phosphate and phosphocholine, and an inhibitor, phosphate ion, at their saturating concentrations exhibited a remarkable, although incomplete, protection against the inactivating action of H2O2/Cu2+. A similar protection was also expressed by divalent metal ions such as Mg2+ or Mn2+. Separately, it was found that H2O2/Fe2+-induced inactivation was prevented by p-nitrophenyl phosphate or Mg2+ but not phosphate ions. Thus, it is implied that phosphocholine-hydrolyzing alkaline phosphatase in brain membrane might be one of enzymes susceptible to metal-catalyzed oxidation.  相似文献   

6.
Hydrogen peroxide (H2O2) inactivates mushroom tyrosinase in a biphasic manner, with the rate being faster in the first phase than in the second one. The inactivation of the enzyme is dependent on H2O2 concentration (in the range of 0.05–5.0 mM), but independent of the pH (in the range of 4.5–8.0). The rate of inactivation of mushroom tyrosinase by H2O2 is faster under anaerobic conditions (nitrogen) than under aerobic ones (air). Substrate analogues such as L-mimosine, L-phenylalanine, p-fluorophenylalanine and sodium benzoate protect the enzyme against inactivation by H2O2. Copper chelators such as tropolone and sodium azide also protect the enzyme. Under identical conditions, apotyrosinase is not inactivated by H2O2, unlike holotyrosinase. The inactivation of mushroom tyrosinase is not accelerated by an OH?dot generating system (Fe2+-EDTA-H2O2) nor is it protected by OHdot scavengers such as mannitol, urate, sodium formate and histidine. Exhaustive dialysis or incubation with catalase does not restore the activity of H2O2-inactivated enzyme. The data suggest that Cu2+ at the active site of mushroom tyrosinase is essential for the inactivation by H2O2. The inactivation does not occur via the OHdot radical in the bulk phase but probably via an enzyme-bound OHdot.  相似文献   

7.
Wang W  Gao P 《Biodegradation》2002,13(6):383-394
A special low-molecular-weight peptide named Gt factor, was isolated and purified from the extracellular culture of brown-rot fungi Gloeophyllum trabeum via gel filtration chromatography and HPLC. It has been shown to reduce Fe3+ to Fe2+. Electron paramagnetic resonance (EPR) spectroscopy revealed Gt factor was able to drive H2O2 generation via a superoxide anion O2 .- intermediate and mediate the formation of hydroxyl radical HO. in the presence of O2. All the results indicated that Gt factor could oxidize the cellulose, disrupt the inter- and intrahydrogen bonds in cellulose chains by a HO. -involved mechanism. This resulted in depolymerization of the cellulose, which made it accessible for further enzymatic hydrolysis.  相似文献   

8.
Supplementation with CaCl2·2H2O (50 mg l−1) or CuSO4·5H2O (10 mg l−1) improved mannitol production by Candida magnoliae by 14.5 and 18.6% (25 and 32 g/L), respectively. When used in combination, they acted synergistically: Ca2+ decreased the intracellular concentration of mannitol 30%, whereas Cu2+ increased the intracellular activity of mannitol dehydrogenase 1.6-times more than control. Ca2+ probably works by altering the permeability of cells to mannitol, whereas, Cu2+ increases the activity of an enzyme responsible for mannitol biosynthesis.  相似文献   

9.
In this work, the effects of iron ion intercalations on lead–tellurate glasses were investigated via FTIR, Raman and UV-Vis spectroscopies. This homogeneous glass system has compositions xFe2O3·(100−x)[4TeO2·PbO2], where x = 0–60 mol%. The presented observations in these mechanisms show that the lead ions have a pronounced affinity towards [TeO3] structural units, resulting in the deformation of the Te–O–Te linkages, and leading to the intercalation of [PbO n ] (n = 3, 4) and [FeO n ] (n = 4, 6) entities in the [TeO4] chain network. The formation of negatively charged [FeO4]1− structural units implies the attraction of Pb2+ ions in order to compensate for this electrical charge. Upon increasing the Fe2O3 content to 60 mol%, the network can accommodate an excess of oxygen through the formation of [FeO6] structural units and the conversion of [TeO4] into [TeO3] structural units. For even higher Fe2O3 contents, Raman spectra indicate a greater degree of depolymerization of the vitreous network than FTIR spectra do. The bands due to the Pb–O bond vibrations are very strongly polarized and the [TeO4] structural units convert into [TeO3] units via an intermediate coordination stage termed “[TeO3+1]” structural units. Our UV-Vis spectroscopic data show two mechanisms: (i) the conversion of the Fe3+ to Fe2+ at the same time as the oxidation of Pb2+ to Pb+4 ions for samples with low Fe2O3 contents; (ii) when the Fe2O3 content is high (x ≥ 50 mol%), the Fe2+ ions capture positive holes and are transferred to Fe3+ ions through a photochemical reaction, while the Pb2+ ions are formed by the reduction of Pb4+ ions. DFT calculations show that the addition of Fe2O3 to lead–tellurate glasses seems to break the axial Te–O bonds, and the [TeO4] structural units are gradually transformed into [TeO3+1]- and [TeO3]-type polyhedra. Analyzing these data further indicates a gradual conversion of the lead ions from covalent to ionic environment. There is then a charge transfer between the tri- and tetracoordinated tellurium atoms due to the capacity of the lead–tellurate network to form the appropriate coordination environments containing structural units of opposite charge, such as iron ions, [FeO4]1−.  相似文献   

10.
DNA-binding proteins from nutrient-starved cells (DPS) protect cells from oxidative stress by removing H2O2 and iron. A new class of DPS-like proteins has recently been identified, with DPS-like protein from Sulfolobus solfataricus (SsDPS) being the best characterized to date. SsDPS protects cells from oxidative stress and is upregulated in response to H2O2 but also in response to iron depletion. The ferroxidase active site of SsDPS is structurally similar to the active sites of manganese catalase and rat liver arginase. The present work shows that the ferroxidase center in SsDPS binds two Mn2+ ions with K D = (1/K 1 K 2)1/2 = 48(3) μM. The binding constant of the second Mn2+ is significantly higher than that of the first, inducing dinuclear Mn(II) cluster formation for all but the lowest concentrations of added Mn2+. In competition experiments, equimolar amounts of Fe2+ were unable to displace the bound manganese. EPR spectroscopy of the Mn2 2+ cluster showed signals comparable to those of other characterized dimanganese clusters. The exchange coupling for the cluster was determined, J = −1.4(3) cm−1 (H = −2JS 1 S 2), and is within the range expected for a μ1,1-carboxylato bridge between the manganese ions. Manganese-bound SsDPS showed catalase activity at a rate 10–100 times slower than for manganese catalases. EPR spectra of SsDPS after addition of H2O2 showed the appearance of an intermediate in the reaction with H2O2.  相似文献   

11.
Summary Hydroxyl radicals (OH') can be formed in aqueous solution by direct reaction of hydrogen peroxide (H2O2) with ferrous salt (Fenton reaction). OH' damage to deoxyribose, measured as formation of thiobarbituric acid-reactive material, was evaluated at different pHs to study the mechanism of action of classical OH' scavengers. OH' scavenger effect on Fe2+ oxidation was also evaluated in the same experimental conditions. In the absence of OH' scavengers, OH' damage to deoxyribose is higher at acidic compared to neutral and moderately basic pH. At acidic pH deoxiribose is per se able to inhibit Fe2+ oxidation by H202. Most of OH' scavengers tested inhibit deoxyribose damage and Fe2+ oxidation in a similar manner: both inhibitions are most relevant at acidic pH and decrease by increasing the pH. These results are not due to OH' scavenger inhibition of Fenton reaction. The influence of pH on the parameters studied appears to be due to the competition of deoxyribose and OH' scavengers for iron. These results suggest the prominent role of iron binding in the degradation of deoxyribose and in the OH' scavenging ability of different compounds. Results obtained with triethylenetetramine, a iron chelator with a low rate constant with OH', confirm that both deoxyribose and the OH' scavengers interact with iron bringing about a site specific Fenton reaction; that the OH' formed at these sites oxidize these molecules to their radical forms which in turn reduce the Fe3– produced by Fenton reaction. The results presented indicate that most of classical OH' scavengers exert their effect predominantly by preventing the site specific reaction between Fe2+ and H202 on the deoxyribose molecule.  相似文献   

12.
Human copper-zinc superoxide dismutase undergoes inactivation when exposed to O2? and H2O2 generated during the oxidation of acetaldehyde by xanthine oxidase at pH 7.4 and 37° C. In contrast, human manganese superoxide dismutase is not inactivated under the same conditions. Catalase and Mn-superoxide dismutase protect CuZn superoxide dismutase from inactivation. Similar protection is observed with hydroxyl radical (OH.) scavengers, such as formate and mannitol. In contrast, other OH. scavengers such as ethanol and tert-butyl alcohol, have no protective action. The latter results indicate that “free OH.” is not responsible for the inactivation. Furthermore, H2O2 generated during the oxidation of glucose by glucose oxidase, i.e., without production of O2?, does not induce CuZn superoxide dismutase inactivation. A mechanism accounting for this O2?H2O2-dependent inactivation of CuZn superoxide dismutase is proposed.  相似文献   

13.
Freshly-added iron only slightly affected the growth of iron-sufficient cells of the green alga Scenedesmus incrassatulus Bohl, strain R-83, but induced accumulation of malondialdehyde (MDA) in cells and excretion of MDA in the medium. These effects were stronger in response to Fe2+ as compared to Fe3+, but Fe3+ induced the release of more iron-binding chelators from these cells than Fe2+. Fe3+ added either in dark or in light induced release of equal concentrations of iron-complexing agents, part of which formed strong chelates with iron in the medium. Exogenously added hydrogen peroxide inhibited iron-induced release of chelators but the effect was removed by addition of the hydroxyl radical scavenger dimethylsulfoxide (DMSO). Malondialdehyde also inhibited the release of chelators. Release of chelators was induced in the absence of iron salts by photoexcited chlorophyll (Chl). The Chl-induced release was efficiently inhibited by singlet oxygen scavengers such as dimethylfuran, -carotene, sodium azide and vitamin B6, and stimulated in D2O or DMSO. Exogenously added catalase inhibited the release more than added superoxide dismutase. The Fe3-induced release of chelators was also inhibited by scavengers of singlet oxygen, but was not affected by sodium azide and by ethanol. Hence both H2O2 and singlet oxygen were involved in induction of chelator release in the absence of iron in light. The induction of chelator release by iron in dark involved H2O2, but not singlet oxygen.  相似文献   

14.
 Dithionite has been found to reduce directly (without mediators) the Escherichia coli R2 subunit of ribonucleotide reductase. With dithionite (∼10 mM) in large excess, the reaction at 25  °C is complete in ∼10 h. Preparations of E. coli R2 have an FeIII 2 (met-R2) component in this work at ∼40% levels, alongside the fully active enzyme FeIII 2 . . . Tyr*, which has a tyrosyl radical at Tyr-122. In the pH range studied (7–8) the kinetics are biphasic. Rate laws for both phases give [S2O4 2–] and not [S2O4 2–]1/2 dependencies, and saturation kinetics are observed for the first time in R2 studies. No dependence on pH was detected. The kinetics (25  °C) of the first phase are reproduced in separate experiments using only met-R2, with association of S2O4 2– to met-R2, K=330 M–1, occurring prior to electron transfer, k et=4.8×10–4 s–1, I=0.100 M (NaCl). The second phase assigned to the reaction of FeIII 2 . . . Tyr* with S2O4 2– gives K=800 M–1 and k et=5.6×10–5 s–1. Bearing in mind the substantially smaller reduction potential for FeIII 2 compared to Tyr*, this is a quite remarkable finding, with implications similar to those already reported for the reaction of R2 with hydrazine, but with additional information provided by the saturation kinetics. The similarity in rates for the two phases (∼fourfold difference) suggests that reduction of FeIII 2 is occurring in both cases, and since S2O4 2– is involved a two-equivalent change is proposed with the formation of FeII 2 . . . Tyr* in the case of active R2. As a sequel to the second phase, intramolecular reduction of the strongly oxidising Tyr* by the FeII 2 is rapid, and further decay of FeIIFeIII is also fast. There is no stable mouse met-R2 form, and the single-phase reaction with dithionite gives saturation kinetics with K=208 M–1 and k et=1.7±10–3 s–1. Mechanistic implications, including the applicability of a pathway for electron transfer via FeA, are considered. Received: 25 February 1998 / Received: 20 August 1998  相似文献   

15.
Summary The effect of low concentrations of hydrogen peroxide (H2O2) (5 × 10−7−9.5 × 10−7 M) on cell growth and antibody production was investigated with murine hybridoma cells (Mark 3 and anti-hPL) in culture. Cell growth, measured by flow cytometry with morphological parameters, was significantly stimulated by H2O2 (8 × 10−7 M) but H2O2 concentration of 7 × 10−6 M and above increased cell death. H2O2 stimulation of antibody production was nonsignificant. The metabolism of cells treated with 8 × 10−7 or 1 × 10−5 M H2O2 was similar to that of the control in terms of glucose and glutamine consumption, lactate and ammonia production, and amino acid concentrations in the medium. The concentrations of lactate dehydrogenase, a marker of cell death, in test and control cells were similar. However, concentrations of intracellular free radicals measured by flow cytometry with dihydrorhodamine 123 (DHR 123) and dichlorofluorescein diacetate (DCFH-DA) as fluorochromes were different. The reactive oxygen species content of cells in 8 × 10−7 M H2O2 was similar to that of the controls, but there was a sudden, marked production of superoxide anions (detected with DHR 123) and H2O2 or peroxides (detected with DCFH-DA) by cells incubated with 1 × 10−5 M H2O2 which increased with increasing H2O2 until cell death.  相似文献   

16.
Chitosan, CN, or H2O2 caused the death of epidermal cells (EC) in the epidermis of pea leaves that was detected by monitoring the destruction of cell nuclei; chitosan induced chromatin condensation and marginalization followed by the destruction of EC nuclei and subsequent internucleosomal DNA fragmentation. Chitosan did not affect stoma guard cells (GC). Anaerobic conditions prevented the chitosan-induced destruction of EC nuclei. The antioxidants nitroblue tetrazolium or mannitol suppressed the effects of chitosan, H2O2, or chitosan + H2O2 on EC. H2O2 formation in EC and GC mitochondria that was determined from 2′,7′-dichlorofluorescein fluorescence was inhibited by CN and the protonophoric uncoupler carbonyl cyanide m-chlorophenylhydrazone but was stimulated by these agents in GC chloroplasts. The alternative oxidase inhibitors propyl gallate and salicylhydroxamate prevented chitosan- but not CN-induced destruction of EC nuclei; the plasma membrane NADPH oxidase inhibitors diphenylene iodonium and quinacrine abolished chitosan- but not CN-induced destruction of EC nuclei. The mitochondrial protein synthesis inhibitor lincomycin removed the destructive effect of chitosan or H2O2 on EC nuclei. The effect of cycloheximide, an inhibitor of protein synthesis in the cytoplasm, was insignificant; however, it was enhanced if cycloheximide was added in combination with lincomycin. The autophagy inhibitor 3-methyladenine removed the chitosan effect but exerted no influence on the effect of H2O2 as an inducer of EC death. The internucleosome DNA fragmentation in conjunction with the data on the 3-methyladenine effect provides evidence that chitosan induces programmed cell death that follows a combined scenario including apoptosis and autophagy. Based on the results of an inhibitor assay, chitosan-induced EC death involves reactive oxygen species generated by the NADPH oxidase of the plasma membrane.  相似文献   

17.
The ability of paraquat radicals (PQ+.) generated by xanthine oxidase and glutathione reductase to give H2O2-dependent hydroxyl radical production was investigated. Under anaerobic conditions, paraquat radicals from each source caused chain oxidation of formate to CO2, and oxidation of deoxyribose to thiobarbituric acid-reactive products that was inhibited by hydroxyl radical scavengers. This is in accordance with the following mechanism derived for radicals generated by γ-irradiation [H. C. Sutton and C. C. Winterbourn (1984) Arch. Biochem. Biophys.235, 106–115] PQ+. + Fe3+ (chelate) → Fe2+ (chelate) + PQ++ H2O2 + Fe2+ (chelate) → Fe3+ (chelate) + OH? + OH.. Iron-(EDTA) and iron-(diethylenetriaminepentaacetic acid) (DTPA) were good catalysts of the reaction; iron complexed with desferrioxamine or transferrin was not. Extremely low concentrations of iron (0.03 μm) gave near-maximum yields of hydroxyl radicals. In the absence of added chelator, no formate oxidation occurred. Paraquat radicals generated from xanthine oxidase (but not by the other methods) caused H2O2-dependent deoxyribose oxidation. However, inhibition by scavengers was much less than expected for a reaction of hydroxyl radicals, and this deoxyribose oxidation with xanthine oxidase does not appear to be mediated by free hydroxyl radicals. With O2 present, no hydroxyl radical production from H2O2 and paraquat radicals generated by radiation was detected. However, with paraquat radicals continuously generated by either enzyme, oxidation of both formate and deoxyribose was measured. Product yields decreased with increasing O2 concentration and increased with increasing iron(DTPA). These results imply a major difference in reactivity between free and enzymatically generated paraquat radicals, and suggest that the latter could react as an enzyme-paraquat radical complex, for which the relative rate of reaction with Fe3+ (chelate) compared with O2 is greater than is the case with free paraquat radicals.  相似文献   

18.
V. A. Bouryi 《Neurophysiology》1998,30(4-5):301-304
Barium currents through ion channels formed by α1-subunit of L-type Ca2+ channel (I α1) were recorded from cultured chinese hamster ovary (CHO) cells. The cells were stably transfected with either a cardiac or a smooth muscle (SM) variant of α1-subunit. TheI α1 in both cases exhibited similar fast voltage-dependent activation kinetics and slow apparent inactivation kinetics. With 10 mM Ba2+ in the bath solution,I α1 was activated at potentials more positive than −40 mV, peaked between 0 and +10 mV, and reversed at about +50 mV. In addition to slow apparent inactivation of inward current, both subunits provided an extremely slow voltage-dependent inactivation at potentials more positive than −100 mV, with half-maximum inactivation at −43.4 mV for cardiac and −41.4 mV for SM α1-subunits. The onset of inactivation as well as recovery from this process were within a time range of minutes. The voltage dependence of steady-state inactivation could be fitted by the sum of two Boltzmann's equations with slope factors of about 12 mV and 5 mV. A less sloped component has its midpoints at −75.6 and −63.7 mV, and a steeper component has its midpoints at −42.8 and −37.7 mV for cardiac and SM α1-subunits, respectively. Relative contribution of the steeper component was higher in both subunits (0.86 and 0.66 for cardiac and SM subunits, respectively). For comparison, the inactivation curves for 5-sec-long conditioning prepulses could be fitted by single Boltzmann's distribution with a 20 mV more positive midpoint and a slope factor of about 13 mV. In contrast to the steady-state inactivation curves, they showed considerable overlap with the steady-state activation curve. Our results reflect functional consequences of known sequence differences between α1-subunits of the cardiac and SM L-type Ca2+ channels and could be used in structural modeling of Ca2+ channel gating. In addition, they show that depolarization-induced window current has a transient nature and decays with the development of extremely slow inactivation. This is the first demonstration that slow inactivation of the L-type Ca2+ channel is an intrinsic property of its α1-subunits.  相似文献   

19.
In a greenhouse study, with and without rice plants, of five flooded Philippine rice soils whose organic C (OC) content varied from 0.5 to 3.6%, incorporation ofSesbania rostrata, Azolla microphylla and rice straw affected the kinetics of soil solution NH 4 + −N, K+, Fe2+, Mn2+, Zn2+, and P. Sesbania and Azolla increased NH 4 + −N concentration above the control treatment, whereas rice straw depressed it. In all soils Azolla released less NH 4 + −N than Sesbania. The apparent net N release depended on the soil and ranged from 44–81% for Sesbania and 27–52% for Azolla. These effects persisted throughout the growth of IR36. Soil solution and exchangeable NH 4 + −N increased initially but levelled off between 30 to 80 days and between 20 to 40 days after flooding (DF), respectively. With rice, soil solution NH 4 + −N concentration, reached a peak at 15–40 DF and declined to very low levels (<4mg L−1). In the 3 soils of low OC content nitrogen derived from green manure ranged from 34–53% and the apparent revovery of added green manure N varied from 29–67%. Almost all N released from both Azolla and Sesbania were recovered in the rice plant in all soils except Concepcion with only 77%. The concentration of K+, Fe2+, Mn2+ and P in the soil solution were higher with rice straw than Sesbania and Azolla in all soils except Hanggan which showed no change in Fe2+ and Mn2+ but increased K+ and P. In general, rice straw, Sesbania and Azolla decreased Zn2+ concentration in all soils.  相似文献   

20.
The relationship between the production of reactive oxygen species and the hypersensitive response (HR) of tobacco (Nicotiana tabacum L.) toward an incompatible race of the Oomycete Phytophthora parasitica var nicotianae has been investigated. A new assay for superoxide radical (O2) production based on reduction of the tetrazolium dye sodium,3′-(1-[phenylamino-carbonyl]-3,4-tetrazolium)-bis(4-methoxy-6-nitro) benzene-sulfonic acid hydrate (XTT) has enabled the quantitative estimation of perhydroxyl/superoxide radical acid-base pair (HO2·/O2) production during the resistant response. Tobacco suspension cells were inoculated with zoospores from compatible or incompatible races of the pathogen. Subsequent HO2·/O2 production was monitored by following the formation of XTT formazan. In the incompatible interaction only, HO2·/O2 was produced in a minor burst between 0 and 2 h and then in a major burst between 8 and 10 h postinoculation. During this second burst, rates of XTT reduction equivalent to a radical flux of 9.9 × 10−15 mol min−1 cell−1 were observed. The HO2·/O2 scavengers O2 dismutase and Mn(III)desferal each inhibited dye reduction. An HR was observed in challenged, resistant cells immediately following the second burst of radical production. Both scavengers inhibited the HR when added prior to the occurrence of either radical burst, indicating that O2 production is a necessary precursor to the HR.  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