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 31 毫秒
1.
The relative effectiveness of oxidizing (.OH, H2O2), ambivalent (O2?) and reducing free radicals (e? and CO2?) in causing damage to membranes and membrane-bound glyceraldehyde-3-phosphate dehydrogenase of resealed erythrocyte ghosts has been determined. The rates of damage to membranebound glyceraldehyde-3-phosphate dehydrogenase (R(enz)) were measured and the rates of damage to membranes (R(mb)) were assessed by measuring changes in permeability of the resealed ghosts to the relatively low molecular weight substrates of glyceraldehyde-3-phosphate dehydrogenase. Each radical was selectively isolated from the mixture produced during gamma-irradiation, using appropriate mixtures of scavengers such as catalase, superoxide dismutase and formate. .OH, O2? and H2 O2 were approximately equally effective in inactivating membrane-bound glyceraldehyde-3-phosphate dehydrogenase, while e? and CO2? were the least effective. R(enz) values of O2? and H2O2 were 10-times and of .OH 15-times that of e?. R(mb) values were quite similar for e? and H2O2 (about twice that of O2?), while that of .OH was 3-times that of O2?. Hence, with respect to R(mb): .OH >e? = H2O2 >O2? , and with respect to R(enz): .OH >O2? = H2O2 >e?. The difference between the effectiveness of the most damaging and the least damaging free radicals was more than 10-fold greater in damage to the enzyme than to the membranes. Comparison between H2O2 added as a chemical reagent and H2O2 formed by irradiation showed that membranes and membrane-bound glyceraldehyde-3-phosphate dehydrogenase were relatively inert to reagent H2O2 but markedly susceptible to the latter.  相似文献   

2.
R.L. Pan  S. Izawa 《BBA》1979,547(2):311-319
NH2OH-treated, non-water-splitting chloroplasts can oxidize H2O2 to O2 through Photosystem II at substantial rates (100–250 μequiv · h?1 · mg?1 chlorophyll with 5 mM H2O2) using 2,5-dimethyl-p-benzoquinone as an electron acceptor in the presence of the plastoquinone antagonist dibromothymoquinone. This H2O2 → Photosystem II → dimethylquinone reaction supports phosphorylation with a Pe2 ratio of 0.25–0.35 and proton uptake with H+e values of 0.67 (pH 8)–0.85 (pH 6). These are close to the Pe2 value of 0.3–0.38 and the H+e values of 0.7–0.93 found in parallel experiments for the H2O → Photosystem II → dimethylquinone reaction in untreated chloroplasts. Semi-quantitative data are also presented which show that the donor → Photosystem II → dibromothymoquinone (→O2) reaction can support phosphorylation when the donor used is a proton-releasing reductant (benzidine, catechol) but not when it is a non-proton carrier (I?, ferrocyanide).  相似文献   

3.
(1) H+/electron acceptor ratios have been determined with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing endogenous substrates during reduction of O2, NO?2 or N2O. Under optimal H+-translocation conditions, the ratios H+O, H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were 6.0–6.3, 4.02, 5.79 and 3.37, respectively. (2) With ascorbate/N,N,N′,N′-tetramethyl-p-phenylenediamine as exogenous substrate, addition of NO?2 or N2O to an anaerobic cell suspension resulted in rapid alkalinization of the outer bulk medium. H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were ?0.84, ?2.33 and ?1.90, respectively. (3) The H+oxidant ratios, mentioned in item 2, were not altered in the presence of valinomycinK+ and the triphenylmethylphosphonium cation. (4) A simplified scheme of electron transport to O2, NO?2 and N2O is presented which shows a periplasmic orientation of the nitrite reductase as well as the nitrous oxide reductase. Electrons destined for NO?2, N2O or O2 pass two H+-translocating sites. The H+electron acceptor ratios predicted by this scheme are in good agreement with the experimental values.  相似文献   

4.
The interaction of |CnH2n+1N+(CH3)3| · I? (n = 3, 6, 9, 12, 14, 16 or 18) with egg-yolk phosphatidylcholine-water dispersions has been studied by 31P-NMR spectroscopy. It is shown that the effective anisotropy of 31P chemical shift (?Δσeff) of the lamellar phospholipid liquid-crystalline phase Lα increases with increasing concentration and alkyl chain length of the drug. Addition of |C6H13N+(CH3)3| ·I ? or |C9H19N+(CH3)3I? to the phospholipid-water dispersion at a molar ratio ammonium salt:phospholipid > 0.8 induces in the dispersion a structure with an effective isotropic phospholipid motion. This structure is unstable and slowly transforms into the hexagonal phase. These effects have not been observed in phospholipid-water dispersions mixed with the ammonium derivatives with the longer alkyl chains n  12, 14, 16 or 18. It is proposed that these results might explain the effects of the investigated drugs on the nerve, muscle and bacterial cells.  相似文献   

5.
A thermodynamic characterization of the Na+-H+ exchange system in Halobacterium halobium was carried out by evaluating the relevant phenomenological parameters derived from potential-jump measurements. The experiments were performed with sub-bacterial particles devoid of the purple membrane, in 1 M NaCl, 2 M KCl, and at pH 6.5–7.0. Jumps in either pH or pNa were brought about in the external medium, at zero electric potential difference across the membrane, and the resulting relaxation kinetics of protons and sodium flows were measured. It was found that the relaxation kinetics of the proton flow caused by a pH-jump follow a single exponential decay, and that the relaxation kinetics of both the proton and the sodium flows caused by a pNa-jump also follow single exponential decay patterns. In addition, it was found that the decay constants for the proton flow caused by a pH-jump and a pNa-jump have the same numerical value. The physical meaning of the decay constants has been elucidated in terms of the phenomenological coefficients (mobilities) and the buffering capacities of the system. The phenomenological coefficients for the Na+-H+ flows were determined as differential quantities. The value obtained for the total proton permeability through the particle membrane via all available channels, LH = (?JH +pH)Δψ,ΔpNa, was in the range of 850–1150 nmol H+·(mg protein)?1·h?1·(pH unit)?1 for four different preparations; for the total Na+ permeability, LNa = (?JNa+pNa)Δψ,ΔpH, it was 1620–2500 nmol Na+·(mg protein)?1·h?1·(pNa unit)?1; and for the proton ‘cross-permeability’, LHNa = (?JH+pNa)Δψ,ΔpH, it was 220–580 nmol H+·(mg protein)?1·h?1·(pNa unit)?1, for different preparations. From the above phenomenological parameters, the following quantities have been calculated: the degree of coupling (q), the maximal efficiency of Na+-H+ exchange (ηmax), the flow and force efficacies (?) of the above exchange, and the admissible range for the values of the molecular stoichiometry parameter (r). We found q ? 0.4; ηmax ? 5%; 0.36 ? r ? 2; ?JNa+ ? 1.3 · 105μmol · (RT unit)?1 at JNa = 1 μmolNa+ · (mgprotein)?1 · h?1; and ?ΔpNa ? 5 · 104 ΔpNa · (mg protein) · h · (RT unit)?1 at ΔpNa = 1 unit, for different preparations.  相似文献   

6.
The action of xanthine oxidase upon acetaldehyde or xanthine at pH 10.2 has been shown to be accompanied by substantial accumulation of O2? during the first few minutes of the reaction. H2O2 decreases this accumulation of O2? presumably because of the Haber-Weiss reaction (H2O2+O2?OH?+OH+O2) and very small amounts of superoxide dismutase eliminate it. This accumulation of O2? was demonstrated in terms of a burst of reduction of cytochrome c, seen when the latter compound was added after aerobic preincubation of xanthine oxidase with its substrate. The kinetic peculiarities of the luminescence seen in the presence of luminol, which previously led to the proposal of H2O4?, can now be satisfactorily explained entirely on the basis of known radical intermediates.  相似文献   

7.
The stoichiometry of free NADPH oxidation in phenobarbital induced rabbit liver microsomes was measured by means of registering the rates of NADPH, H+ and O2 consumption and O2? and H2O2 production. ΔO2?:ΔH2O2 ratio is approximately I indicating that about half H2O2 results from O2? dismutation, the second half being formed directly. ΔNADPH:ΔH2O2 and ΔO2:ΔH2O2 ratios exceed I and therefore another product of the reaction is water. The fact that the ratio (ΔNADPH-ΔH2O2):(ΔO2-ΔH2O2) is 2 allows one to consider direct 4-electron O2 reduction as the major way of water formation rather than endogenous substrate hydroxylation.  相似文献   

8.
(1) Treatment of (Na+ + K+)-ATPase from rabbit kidney outer medulla with the γ-35S labeled thio-analogue of ATP in the presence of Na+ + Mg2+ and the absence of K+ leads to thiophosphorylation of the enzyme. The Km value for [γ-S]ATP is 2.2 μM and for Na+ 4.2 mM at 22°C. Thiophosphorylation is a sigmoidal function of the Na+ concentration, yielding a Hill coefficient nH = 2.6. (2) The thio-analogue (Km = 35 μM) can also support overall (Na+ + K+)-ATPase activity, but Vmax at 37°C is only 1.3 γmol · (mg protein)? · h?1 or 0.09% of the specific activity for ATP (Km = 0.43 mM). (3) The thiophosphoenzyme intermediate, like the natural phosphoenzyme, is sensitive to hydroxylamine, indicating that it also is an acylphosphate. However, the thiophosphoenzyme, unlike the phosphoenzyme, is acid labile at temperatures as low as 0°C. The acid-denatured thiophosphoenzyme has optimal stability at pH 5–6. (4) The thiophosphorylation capacity of the enzyme is equal to its phosphorylation capacity, indicating the same number of sites. Phosphorylation by ATP excludes thiophosphorylation, suggesting that the two substrates compete for the same phosphorylation site. (5) The (apparent) rate constants of thiophosphorylation (0.4 s?1 vs. 180 s?1), spontaneous dethiophosphorylation (0.04 s?1 vs. 0.5 s?1) and K+-stimulated dethiophosphorylation (0.54 s?1 vs. 230 s?1) are much lower than those for the corresponding reactions based on ATP. (6) In contrast to the phosphoenzyme, the thiophosphoenzyme is ADP-sensitive (with an apparent rate constant in ADP-induced dethiophosphorylation of 0.35 s?1, KmADP = 48 μM at 0.1 mM ATP) and is relatively K+-insensitve. The Km for K+ in dethiophosphorylation is 0.9 mM and in dephosphorylation 0.09 mM. The thiophosphoenzyme appears to be for 75–90% in the ADP-sensitive E1-conformation.  相似文献   

9.
A method for calculating the rate constant (KA1A2) for the oxidation of the primary electron acceptor (A1) by the secondary one (A2) in the photosynthetic electron transport chain of purple bacteria is proposed.The method is based on the analysis of the dark recovery kinetics of reaction centre bacteriochlorophyll (P) following its oxidation by a short single laser pulse at a high oxidation-reduction potential of the medium. It is shown that in Ectothiorhodospira shaposhnikovii there is little difference in the value of KA1A2 obtained by this method from that measured by the method of Parson ((1969) Biochim. Biophys. Acta 189, 384–396), namely: (4.5±1.4) · 103s?1 and (6.9±1.2) · 103 s?1, respectively.The proposed method has also been used for the estimation of the KA1A2 value in chromatophores of Rhodospirillum rubrum deprived of constitutive electron donors which are capable of reducing P+ at a rate exceeding this for the transfer of electron from A1 to A2. The method of Parson cannot be used in this case. The value of KA1A2 has been found to be (2.7±0.8) · 103 s?1.The activation energies for the A1 to A2 electron transfer have also been determined. They are 12.4 kcal/mol and 9.9 kcal/mol for E. shaposhnikovii and R. rubrum, respectively.  相似文献   

10.
G. Peters  M.A.J. Rodgers 《BBA》1981,637(1):43-52
Laser flash photolysis techniques have yielded rate constants for physical and reactive quenching modes of O2(1Δg) by nicotine, nicotinamide adenine dinucleotide (oxidized and reduced forms) and the reduced forms of nicotinamide mononucleotide, nicotinamide adenine dinucleotide phosphate and nicotinamide hypoxanthine dinucleotide. In the case of the last four named compounds, kinetic spectroscopy furnished evidence for one-electron transfers to O2(1Δg). Specifically, production of O?2 was demonstrated unequivocally by reaction with 1,4-benzoquinone. Quantitative determinations revealed the extent of reactive quenching to be near 60% in each case.  相似文献   

11.
The mechanism of the reaction of bis-(salicylato)-copper(II) with superoxide anion has been studied by utilizing electron paramagnetic resonance and polarographic techniques. The proposed reaction sequence is as follows:
Cu(II) + O2?Cu(II)O2?Cu(I)O2?Cu(I) + O2
Cu(I) + O2?Cu(I)O2?Cu(II)O22?2H+Cu(II)O + H2O2
Using the xanthine-xanthine oxidase system as a superoxide generator, it was found that the concentration of this copper complex for 50% inhibition of the xanthine-cytochrome c reductase activity was about 1000 times more per mole of copper than that of erythrocyte superoxide dismutase.  相似文献   

12.
The permeability of the lysosomal membrane to small anions and cations was studied at 37°C and pH 7.0 in a lysosomal-mitochondrial fraction isolated from the liver of untreated rats. The extent of osmotic lysis following ion influx was used as a measure of ion permeancy. In order to preserve electroneutrality, anion influx was coupled to an influx of K+ in the presence of valinomycin, and cation influx was coupled to an efflux of H+ using the protonophore 3-tert-butyl-5,2′-dichloro-4′-nitrosalicilylanilide. Lysosomal lysis was monitored by observing the loss of latency of two lysosomal hydrolases.The order of permeability of the lysosomal membrane to anions was found to be SCN? > I? > CH3COO? > Cl? ≈ HCO?3 ≈ Pi > SO42? and that to cations Cs+ > K+ > Na+ > H+. These orders are largely in agreement with the lyotropic series of anions and cations.The implications of these findings for the mechanism by means of which a low intralysosomal pH is produced and maintained are discussed.  相似文献   

13.
14.
Cultured epithelial monolayers of MDCK cells grown upon Millipore filter supports and mounted in Ussing chambers for transport studies respond to addition of 5 · 10?7M adrenalin from only the basal bathing solution by an increased short-circuit current, due both to an increased transmonolayer potential difference (basal solution electropositive) and an increased transmonolayer conductance. Measurement of tracer Na+, K+ and Cl? fluxes demonstrate that the adrenalin-stimulated short-circuit current results primarily from basal to apical net Cl? secretion. Half-maximal stimulation of the short-circuit current was observed at (3.1 ± 0.3) · 10?8M adrenalin; the order of potency of adrenergic agonists for short-circuit current stimulation was isoprenalin >adrenalin >noradrenalin, consistent with adrenalin action being mediated by a β-adrenergic receptor. The adrenalin-stimulated short-circuit current was sensitive to inhibition (75%) by basal additions of furosemide (1 · 10?4M); phloretin inhibition (54%, 57%) was observed from both epithelial surfaces. Amiloride (10?4 M) and 4-acetamido-4-isothiocyanostilbene-2, 2′-disulphonic acid (SITS) (10 μM) were ineffective as inhibitors of the adrenalin response. The increased short-circuit current was sensitive to replacement of medium Na+ by choline (87%) and Tris (93%). Li+ was a partially effective substitute cation for Na+ · NO3?, and isethionate were ineffective substitutes for Cl? whereas Br? was partially effective. Partial replacement of medium Na+ by choline gave an upward-curving non-saturable dependence of the adrenalin-stimulated short-circuit current upon [Na]; partial replacement of Cl? by NO3? in contrast gave a saturable increase with a K12 of approx. 65 mM Cl?.  相似文献   

15.
Unidirectional fluxes of [14C]lactose by whole cells of Escherichia coli under highly energized and partially de-energized (in the presence of CN?) conditions are analyzed kinetically.When the cells are energized, the value for V influx is 0.45 ± 0.01 mM internal concentration increment/s and Kt is 0.26 ± 0.03 mM. At an external concentration of 0.61 mM the steady-state internal concentration is 0.25 M, reached after about 1h. The maximum steady-state concentration ratio is 2 · 103.The efflux process under these conditions is non-saturable, being linearly dependent upon internal concentration over the range 25–250 mM with a first-order rate constant of 8.8 ± 0.2 · 10?4 s?1.The transport in the presence of CN? is active, with a maximum concentration ratio (internal concentration/external concentration) of 104, and the uptake is mimicked by anoxia (< 70 ppm O2).The effects of CN? are to lower the V for influx and to change the efflux from a non-saturable to a saturable process with a value for Kt (60 mM) intermediate between that for energized efflux (> 250 mM) and influxe (0.3–0.6 mM), the latter value not changing appreciably. Partial de-energization thus affects both the influx and efflux processes.  相似文献   

16.
When certain ferrous chelates are added to lipid, peroxidation of the lipid occurs following a short lag. This suggests that a product of ferrous autoxidation is required to initiate lipid peroxidation. This autoxidation product is apparently ferric iron, rather than the oxygen radicals which also result from ferrous autoxidation. Studies with oxy-radical scavengers and catalase suggest that O2? H2O2, or the ·OH are not involved in the initiation reactions, therefore, we propose that a ferrous-dioxygen-ferric chelate complex may be the initiating species.  相似文献   

17.
The reactivities of anionic nitroalkanes with 2-nitropropane dioxygenase of Hansenula mrakii, glucose oxidase of Aspergillus niger, and mammalian d-amino acid oxidase have been compared kinetically. 2-Nitropropane dioxygenase is 1200 and 4800 times more active with anionic 2-nitropropane than d-amino acid oxidase and glucose oxidase, respectively. The apparent Km values for anionic 2-nitropropane are as follows: 2-nitropropane dioxygenase, 1.61 mm; glucose oxidase, 16.7 mm; and d-amino acid oxidase, 11.1 mm. Anionic 2-nitropropane undergoes an oxygenase reaction with 2-nitropropane dioxygenase and glucose oxidase, and an oxidase reaction with d-amino acid oxidase. In contrast, anionic nitroethane is oxidized through an oxygenase reaction by 2-nitropropane dioxygenase, and through an oxidase reaction by glucose oxidase. All nitroalkane oxidations by these three flavoenzymes are inhibited by Cu and Zn-superoxide dismutase of bovine blood, Mn-superoxide dismutases of bacilli, Fe-superoxide dismutase of Serratia marcescens, and other O2? scavengers such as cytochrome c and NADH, but are not affected by hydroxyl radical scavengers such as mannitol. None of the O2? scavengers tested affected the inherent substrate oxidation by glucose oxidase and d-amino acid oxidase. Furthermore, the generation of O2? in the oxidation of anionic 2-nitropropane by 2-nitropropane dioxygenase was revealed by ESR spectroscoy. The ESR spectrum of anionic 2-nitropropane plus 2-nitropropane dioxygenase shows signals at g1 = 2.007 and g11 = 2.051, which are characteristic of O2?. The O2? generated is a catalytically essential intermediate in the oxidation of anionic nitroalkanes by the enzymes.  相似文献   

18.
The redox potential of ?O2? was determined based on the dependence of the electron transfer reaction from ?O2? upon the known redox potential of various acceptors A (including a range of quinones, dyes and ferricyanide). The efficiencies and the rates of these electron transfer processes were determined, using the technique of pulse radiolysis, by monitoring the formation kinetics of the semiquinone radical anions at the appropriate wavelength. From the percentage efficiency versus Eo′ plot, an Eo′ value of + 0.15 ± 0.01 V at pH 7.0 and 22°C, or Eo = + 0.57 ± 0.01 V, for the ?O2?O2 couple was obtained. The rate k(?O2? + A → O2 + ?A?) = 9.8 × 108 M?1 sec?1 where A = p-benzoquinone. The Eo value for the ?HO2 radical is > 1.0 V. It was also found that hydroquinone can quantitatively reduce ?O2? to H2O2, k(?O2? + QH2 → QH? + HO2?) = 1.6 ± 0.1 × 107 M?1 sec?1 at pH 7.0 and 22°C.  相似文献   

19.
N-Phenylhydroxylamine is oxidized in aqueous phosphate buffer to nitrosobenzene, nitrobenzene, and azoxybenzene. Degradation is O2 dependent and shows general catalysis by H2PO4? (k1 = 2.3 M?2 sec?1) and PO4?3 (k2 = 2.3 × 105M?2 sec?1) or kinetically equivalent terms. Evidence is presented suggesting the intermediacy of a highly reactive species leading to these products.  相似文献   

20.
(1) Aqueous solutions of 1–10 μM ferricytochrome c treated with 100 μM–100 mM H2O2 at pH 8.0 emit chemiluminescence with quantum yield Ф ? 10?9 and absolute maximum intensity Imax ? 105 hv/s per cm3 (λ = 440), and exhibit exponential decay with a rate constant of 0.15 s?1. (2) The emission spectrum of the chemiluminescence covers the range 380–620 nm with the maximum at 460 ± 10 nm. (3) Neither cytochrome c nor haemin fluoresce in the spectral region of the chemiluminescence. In the reaction course with H2O2, a weak fluorescence in the region 400–620 nm with λmax = 465–510 nm (λexc 315–430 nm) gradually arises. This originates from tryptophan oxidation products of the formylkynurenine type or from imidazole derivatives, respectively. (4) Frozen solutions (77 K) of cytochrome c exhibit phosphorescence typical of tryptophan (λexc = 280 nm, λem = 450 nm). During the peroxidation, an additional phosphorescence gradually appears in the range 480–620 nm with λmax = 530 nm (λexc = 340 nm). This originates from oxidative degradation products of tryptophan. (5) There are no red bands in the chemiluminescence spectra of cytochrome c or haemin. This result suggests that singlet molecular oxygen O2(1Δg) is not involved in either peroxidation or chemiluminescence. (6) The haem Fe3+ group and H2O2 appear to be crucial for the chemiluminescence. It is suggested that the generation of electronically excited, light-emitting states is coupled to the production of conformational out-of-equilibrium states of peroxy-Fe-protoporphyrin IX compounds.  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