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.
There are five oxidation-reduction states of horseradish peroxidase which are interconvertible. These states are ferrous, ferric, Compound II (ferryl), Compound I (primary compound of peroxidase and H2O2), and Compound III (oxy-ferrous). The presence of heme-linked ionization groups was confirmed in the ferrous enzyme by spectrophotometric and pH stat titration experiments. The values of pK were 5.87 for isoenzyme A and 7.17 for isoenzymes (B + C). The proton was released when the ferrous enzyme was oxidized to the ferric enzyme while the uptake of the proton occurred when the ferrous enzyme reacted with oxygen to form Compound III. The results could be explained by assuming that the heme-linked ionization group is in the vicinity of the sixth ligand and forms a stable hydrogen bond with the ligand.The measurements of uptake and release of protons in various reactions also yielded the following stoichiometries: Ferric peroxidase + H2O2 → Compound I, Compound I + e? + H+ → Compound II, Compound II + e? + H+ → ferric peroxidase, Compound II + H2O2 → Compound III, Compound III + 3e? + 3H+ → ferric peroxidase.Based on the above stoichiometries and assuming the interaction between the sixth ligand and heme-linked ionization group of the protein, it was possible to picture simple models showing structural relations between five oxidation-reduction states of peroxidase. Tentative formulae are as follows: [Pr·Po·Fe-(II) $?PrH+·Po·Fe(II)] is for the ferrous enzyme, Pr·Po·Fe(III)OH2 for the ferric one, Pr·Po·Fe(IV)OH? for Compound II, Pr(OH?)·Po+·Fe(IV)OH? for Compound I, and PrH+·Po·Fe(III)O2? for Compound III, in which Pr stands for protein and Po for porphyrin. And by Fe(IV)OH?, for instance, is meant that OH? is coordinated at the sixth position of the heme iron and the formal oxidation state of the iron is four.  相似文献   

2.
To evaluate the biological preference of [Yb(phen)2(OH2)Cl3](H2O)2 (phen is 1,10-phenanthroline) for DNA, interaction of Yb(III) complex with DNA in Tris–HCl buffer is studied by various biophysical and spectroscopic techniques which reveal that the complex binds to DNA. The results of fluorescence titration reveal that [Yb(phen)2(OH2)Cl3](H2O)2 has strongly quenched in the presence of DNA. The binding site number n, apparent binding constant K b, and the Stern–Volmer quenching constant K SV are determined. ΔH 0, ΔS 0, and ΔG 0 are obtained based on the quenching constants and thermodynamic theory (ΔH 0?>?0, ΔS 0?>?0, and ΔG 0?<?0). The experimental results show that the Yb(III) complex binds to DNA by non-intercalative mode. Groove binding is the preferred mode of interaction for [Yb(phen)2(OH2)Cl3](H2O)2 to DNA. The DNA cleavage results show that in the absence of any reducing agent, Yb(III) complex can cleave DNA. The antimicrobial screening tests are also recorded and give good results in the presence of Yb(III) complex.  相似文献   

3.
The preparation of the planar yellow [Ni([8]aneN2)2](ClO4)2 is described. The complex dissociates in basic solution, with rate = kOH[NiL][OH?] (L = 1,5-diazacyclo-octane). At 25 °C, kOH = 4.5 x 10?2 M?1 s?1 and the corresponding activation parameters are ΔH = 69.2 kJ mol?1 and ΔS298 = ?38.6 J K?1 mol?1. Acid catalysed dissociation in quite slow even in strongly acidic solutions. The kinetic data in this case can be fitted to the expression Kobs = ko + KH[H+], where ko relates to a solvolytic pathway and kH to the acid catalysed pathway. At 60 °C, Ko = 2 x 10?5 s?1 and kH is 2 x 10?5 M?1 s?1. Possible mechanisms for these reactions are considered.The Ni(II)/Ni(III) redox couple for NiLn+ is irreversible on Pt using MeCN as solvent.  相似文献   

4.
The biological effects of ultraviolet radiation (UV), such as DNA damage, mutagenesis, cellular aging, and carcinogenesis, are in part mediated by reactive oxygen species (ROS). The major intracellular ROS intermediate is hydrogen peroxide, which is synthesized from superoxide anion (O2) and further metabolized into the highly reactive hydroxyl radical. In this study, we examined the involvement of mitochondria in the UV‐induced H2O2 accumulation in a keratinocyte cell line HaCaT. Respiratory chain blockers (cyanide‐p‐trifluoromethoxy‐phenylhydrazone and oligomycin) and the complex II inhibitor (theonyltrifluoroacetone) prevented H2O2 accumulation after UV. Antimycin A that inhibits electron flow from mitochondrial complex III to complex IV increased the UV‐induced H2O2 synthesis. The same effect was seen after incubation with rotenone, which blocks electron flow from NADH‐reductase (complex I) to ubiquinone. UV irradiation did not affect mitochondrial transmembrane potential (ΔΨm). These data indicate that UV‐induced ROS are produced at complex III via complex II (succinate‐Q‐reductase). J. Cell. Biochem. 80:216–222, 2000. © 2000 Wiley‐Liss, Inc.  相似文献   

5.
The Oxygen activating mechanism of Fusarium lipoxygenase, a heme-containing dioxygenase, was studied. The enzyme did not require any cofactors, such as H2O2, however, both superoxide dismutase and catalase inhibited linoleate peroxidation by Fusarium lipoxygenase. A low concentration of H2O2 caused a distinct acceleration in enzymatic peroxidation. These results indicate that both O2? and H2O2 are produced as essential intermediates of oxygen activation during formation of linoleate hydroperoxides by Fusarium lipoxygenase. This peroxidation reaction was also prevented by scavengers of singlet oxygen (1O2), but not by scavengers of hydroxy 1 radical (OH). Generation of O2? in the enzyme reaction was detected by its ability to oxidize epinephrine to adrenochrome. Moreover, the rate of peroxide formation was greater in the D2O than in the H2O buffer system. These results suggest that the Haber–Weiss reaction (O2?+H2O2→OH?+OH·+1O2) is taking part in linoleate peroxidation by Fusarium lipoxygenase, and the 1O2 evolved could be responsible for the peroxidation of linoleate. H2O2 produced endogenously in the enzyme reaction might act as an activating factor for the enzyme. This possible mechanism of oxygen activation can explain the absence of a need for exogenous cofactors with Fusarium lipoxygenase in contrast to an other heme-containing dioxygenase, tryptophan pyrrolase, which requires an exogenous activating factor, such as H2O2.  相似文献   

6.
In order to evaluate biological potential of a novel synthesized complex [Nd(dmp)2Cl3.OH2] where dmp is 29-dimethyl 110-phenanthroline, the DNA-binding, cleavage, BSA binding, and antimicrobial activity properties of the complex are investigated by multispectroscopic techniques study in physiological buffer (pH 7.2).The intrinsic binding constant (Kb) for interaction of Nd(III) complex and FS–DNA is calculated by UV–Vis (Kb = 2.7 ± 0.07 × 105) and fluorescence spectroscopy (Kb = 1.13 ± 0.03 × 105). The Stern–Volmer constant (KSV), thermodynamic parameters including free energy change (ΔG°), enthalpy change (?H°), and entropy change (?S°), are calculated by fluorescent data and Vant’ Hoff equation. The experimental results show that the complex can bind to FS–DNA and the major binding mode is groove binding. Meanwhile, the interaction of Nd(III) complex with protein, bovine serum albumin (BSA), has also been studied by using absorption and emission spectroscopic tools. The experimental results show that the complex exhibits good binding propensity to BSA. The positive ΔH° and ?S° values indicate that the hydrophobic interaction is main force in the binding of the Nd(III) complex to BSA, and the complex can quench the intrinsic fluorescence of BSA remarkably through a static quenching process. Also, DNA cleavage was investigated by agarose gel electrophoresis that according to the results cleavage of DNA increased with increasing of concentration of the complex. Antimicrobial screening test gives good results in the presence of Nd(III) complex system.  相似文献   

7.
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   

8.
In the present study, we investigated the alteration of reactive oxygen species production along the longitudinal axis of barley root tips during Cd treatment. In unstressed barley root tips, H2O2 production decreased from the root apex towards the differentiation zone where again, a slight increase was observed towards the more mature region of root. An opposite pattern was observed for O 2 ?? and OH? generation. The amount of both O 2 ?? and OH? was highest in the elongation zone, decreased in the root apex and at the differentiation zone of root, then increased again towards the more mature region of root. An elevated Cd-induced O 2 ?? production started in the elongation zone and increased further along the differentiation zone of barley root tip. In contrast, Cd-induced H2O2 production was localised to the root elongation zone and to the beginning of the differentiation zone. In contrast to Cd-induced H2O2 and O 2 ?? production, Cd reduced OH? production along the whole barley root tip. Our results suggest that not only an increase but also the spatial distribution of reactive oxygen species production is involved in the Cd-induced stress response of barley root tip.  相似文献   

9.
Effects of deficient (20mmol m?3) and sufficient (1000 mmol m?3) magnesium (Mg) supply and of varied light intensity (100 μmol m?2 s?1 to 580 μmol m?2 s?1) on paraquat-dependent chlorophyll destruction in bean (Phaseolus vulgaris) plants grown in nutrient solution were studied over a 12-d period using leaf discs or intact primary leaves. Treatment of leaf discs with 10mmol m 3 paraquat for 15h caused severe chlorophyll loss, especially with increasing light intensity. This chlorophyll destruction by paraquat was very much higher in Mg-sufficient than Mg-deficient leaves. The occurrence of paraquat resistance in Mg deficient leaves was already apparent after 6d growth in nutrient solution, i.e. before any decrease in chlorophyll or growth by Mg deficiency was evident. Also, following foliar application of paraquat (10–140 mmol m?3) to intact plants, Mg-deficient plants were much more resistant to paraquat, even following longer exposure duration (72 h) and four to 14 times higher paraquat concentrations than those received by Mg sufficient plants. From experiments where exogenous scavengers of superoxide radical (O2.-), hydroxyl radical (OH·) and singlet oxygen (1O2) were applied to leaf discs, it appears that O2.-, and partly, OH· are the main O2 species which contribute to chlorophyll destruction by paraquat. The results demonstrate that Mg-deficient bean plants become highly resistant to O2.--mediated and light-induced paraquat injury. The mode of this paraquat resistance is attributed to well-known stimulative effects of Mg deficiency on O2.- and H2O2 scavenging enzymes and antioxidants.  相似文献   

10.
The scavenging of superoxide radical by manganous complexes: in vitro   总被引:22,自引:0,他引:22  
Dialyzable manganese has been shown to be present in millimolar concentrations within cells of Lactobacillus plantarum and related lactic acid bacteria. This unusual accumulation of Mn appears to serve the same function as Superoxide dismutase (SOD), conferring hyperbaric oxygen and Superoxide tolerance on these SOD-free organisms. The form of the Mn in the lactic acid bacteria and the mechanisms whereby it protects the cell from oxygen damage are unknown. This report examines the mechanisms by which Mn catalytically scavenges O2?, both in the xanthine oxidase/cytochrome c SOD assay and in a number of in vitro systems relevant to the in vivo situation. In all the reaction mixtures examined, Mn(II) is first oxidized by O2? to Mn(III), and H2O2 is formed. In pyrophosphate buffer the Mn(III) thus formed is re-reduced to Mn(II) by a second O2?, making the reaction a true metal-catalyzed dismutation like that catalyzed by SOD. Alternatively, if the reaction takes place in orthophosphate or a number of other buffers, the Mn(III) is preferentially reduced largely by reductants other than O2?, such as thiols, urate, hydroquinone, or H2O2. H2O2, a common product of the lactic acid bacteria, reacted rapidly with Mn(III) to form O2, apparently without intermediate O2 release. Free hexaquo Mn(II) ions were shown by electron spin resonance spectroscopy and activity assays in noncomplexing buffers to be poorly reactive with O2?. In contrast, Mn(II) formed complexes having a high catalytic activity in scavenging O2? with a number of organic acids, including malate, pyruvate, propionate, succinate, and lactate, with the Mn-lactate complex showing the greatest activity.  相似文献   

11.
A simple chemical system consisting of FeSO4 and H2O2 (Fenton's reagent) was shown to emit light (chemiluminescence). The addition of tryptophan to the reaction markedly enhanced light production. Very little chemiluminescence was observed when H2O2 was omitted from the reaction and when ferric, instead of ferrous, ions were used. Hydroxyl radical (OH.) and singlet oxygen (1ΔgO2) quenchers suppressed chemiluminescence of the FeSO4 + tryptophan + H2O2 system; and, deuterium oxide (2H2O) enhanced chemiluminescence of both FeSO4 reactions. These observations suggest that a radical chain reaction involving both OH. and 1ΔgO2 is responsible for the chemiluminescent reactions. Six iron-containing proteins, some of which are located within granulocytes, all emitted light in the presence of H2O2. Since iron and H2O2 are present in metabolically stimulated granulocytes, it is likely that chemiluminescent reactions similar to the ones demonstrated in this study account for part of the chemiluminescence of activated granulocytes.  相似文献   

12.
Acid-base regulation during nitrate assimilation in Hydrodictyon africanum   总被引:8,自引:5,他引:3  
Abstract The acid-base balance during NO3? assimilation in Hydrodictyon africanum has been investigated during growth from (1) an analysis of the elemental composition of the cells, (2) the alkalinity of the ash and (3) the net H+ changes in the medium during growth. These investigations agree in showing that some 0.25 excess organic negative charges are generated per N assimilation from No3? as N-source and C02 as C-source; the excess OH? (0.75 OH? per NO3? assimilated) appears in the medium. Approximately half of the excess organic negative charge is attributable to cell wall uronates; the remainder is intracellular. All of the excess OH? appearing in the medium must have crossed the plasmalemma (as net downhill H+ influx or OH? efflux). Previous work has shown that the value of ψco is more negative than ψK+ during NO3? assimilation, suggesting that the active electrogenic H+ extrusion pump is still operative despite the net downhill H+ influx. The interpretation of this in terms of H+?NO3? symport which causes the entry of more H+ than is consumed in NO3? metabolism, with extrusion of the excess H+via the active, electrogenic H+ pump, was tested by measuring short-term H+ influx upon addition of NO?3. A net H+ influx occurs before NOa assimilation (as indicated by additional O2 evolution in the light) has commenced, suggesting a mechanistic relation of H+ and NO3? influxes. This is consistent with the interpretation suggested above. Determinations of cytoplasmic pH showed no significant effect of NO3? assimilation, suggesting that cytoplasmic pH changes sufficient to change the ‘pH-regulating’ H+ fluxes are smaller than the errors in the determination of cytoplasmic pH.  相似文献   

13.
Summary

Reactive oxygen species (ROS) such as hydrogen peroxide (H2O2), superoxide anion (O2?), and hydroxyl radical (OH?) have been implicated in mediating various pathological events such as cancer, atherosclerosis, diabetes, ischemia, inflammatory diseases, and the aging process. The glutathione (GSH) redox cycle and antioxidant enzymes—superoxide dismutase (SOD) and catalase (CAT)—play an important role in scavenging ROS and preventing cell injury. Pycnogenol has been shown to protect endothelial cells against oxidant-induced injury. The present study determined the effects of pycnogenol on cellular metabolism of H2O2 and O2? and on glutathione-dependent and -independent antioxidant enzymes in bovine pulmonary artery endothelial cells (PAEC). Confluent monolayers of PAEC were incubated with pycnogenol, and oxidative stress was triggered by hypoxanthine and xanthine oxidase or H2O2. Pycnogenol caused a concentration-dependent enhancement of H2O2 and O2? clearance. It increased the intracellular GSH content and the activities of GSH peroxidase and GSH disulfide reductase. It also increased the activities of SOD and CAT. The results suggest that pycnogenol promotes a protective antioxidant state by upregulating important enzymatic and nonenzymatic oxidant scavenging systems.  相似文献   

14.
Bis-Methyl N,N-diethylcarbamylmethylenephosphonato dysprosium thiocyanate, Dy[O2P(OCH3)CH2C(O)N(C2H5)2]2(NCS) was prepared from the combination of ethanolic solutions of Dy(NCS)3·xH2O and (CH3O)2P(O)CH2C(O)N(C2H5)2. The complex was characterized by infrared and NMR spectroscopy, and single crystal X-ray diffraction methods. The crystal structure was determined at 25 °C from 3727 independent reflections by using a standard automated diffractometer. The complex was found to crystallize in the monoclinic space group P21/c with a = 13.282(4) Å, b = 19.168(5) Å, c = 9.648(2) Å, β = 90.09(2)°, Z = 4, V = 2456.4 Å3 and ?cald = 1.72 g cm?3. The structure was solved by standard heavy atom techniques, and blocked least-squares refinement converged with Rf = 4.7% and RwF = 4.9%. The Dy atom is seven coordinate and bonded in a bidentate fashion to two anionic phosphonate ligands [O2P(OCH3)CH2C(O)N(C2H5)2?] through the carbonyl oxygen atoms and one of two phosphonate oxygen atoms. In addition, each Dy atom is coordinated to two phosphonate oxygen atoms from two neighboring complexes and to the nitrogen atom of a thiocyanate ion. This coordination scheme gives rise to a two-dimensional polymeric structure. Some important bond distances include DyNCS 2.433(8) Å, DyO(carbonyl)avg 2.39(2) Å, DyO(equat. phosphoryl)avg 2.303(8) Å, DyO(axial phosphoryl)avg 2.25(2), PO(phosphoryl)avg 1.493(3) Å and CO(carbonyl)avg 1.25(1) Å.  相似文献   

15.
《Inorganica chimica acta》1987,133(2):347-352
When crystals of [Dy(OH2)7(OHMe)] [DyCl(OH2)2(18- crown-6)]2Cl7·2H2O [1] are allowed to warm from 5 °C to ambient temperature (22 °C) under the original solvent mixture (1:3 CH3OH: CH3CN), they redissolve and the title complex can be isolated by slow evaporation of the resulting solution. The crystal structure of this complex, [Dy(OH2)8]Cl3·18-crown-6·4H2O, has been determined. It crystallizes in the monoclinic space group, P21/c, with a = 10.395(1), b = 18.684(1), c = 16.259- (3) Å, β= 102.56(1)°, and Dcalc = 1.61 g cm−3 for Z = 4. A final conventional R value of 0.041 was obtained by least-squares refinement using 3453 independent observed [Fo⩾5σ(Fo)] reflections. The [Dy(OH2)8]3+ cations and crown ether molecules are hydrogen bonded in a polymeric chain with the crown molecules separating the cations and a total of seven DyOH2···O(crown ether) hydrogen bonds. The chains are connected by a hydrogen bonding network consisting of the cations, chloride ions, and uncoordinated water molecules. The geometry of the cation is best described as a bicapped trigonal prism with distortions on the reaction pathway toward dodecahedral symmetry. The two capping atoms average 2.41(1) Å from Dy, the remaining DyO distances average 2.38(2) Å. The 18-crown-6 molecule has the D3d conformation normally observed except for a distortion of one OCCO unit containing the oxygen atom accepting two hydrogen bonds.  相似文献   

16.
We studied anionic inhibition of the reaction CO2 + OH?? HCO3? catalyzed by human red cell carbonic anhydrase B (I) and C (II), using iodide and cyanate. In the forward reaction with respect to CO2 as the substrate, inhibition was mixed but favoring noncompetitive; the back reaction, with HCO3? as the substrate, yielded strict competitive kinetics. Mean inhibition constants, KI, in the pH range 7.2–7.5 are: iodide, 0.5 mm for enzyme B and 16 mm for C; cyanate, 0.8 μm for B and 20 μm for C. When OH? was considered as the substrate for the forward reaction, cyanate and chloride behaved as competitive inhibitors. The true inhibition constant (KI0) for cyanate (calculated for infinitely low OH?) is 0.4 μm for enzyme B and 4 μm for C. Apart from the difference in anion affinity and some 10-fold higher activity of C > B, the isozymes showed similar patterns of inhibition. Data agree with generally proposed mechanisms describing the active site as ZnH2O with pKa of about 7.  相似文献   

17.
The extraction of U(VI)with dicyclohexano-18-crown-6 (mixed isomers or isomer A) from HCl medium is effective and selective, and can be used for separating and analysing uranium and thorium. However, little is known of the properties of the extraction complex of uranium with crown-ether in organic phase. In this paper we report the preparation, characteristic and structure of the crystalline extraction complex IaUO2Cl2HClH2O, Iabeing isomer A of dicyclohexano-18-crown-6.After extracting uranium(VI) from aqueous hydrochloric acid solution with Ia in 1,2-dichloroethane, the crystalline product of the extraction complex was prepared from the organic phase by diluting with a non-polar solvent at 25 °C. The content of uranium, crown-ether and HCl was determined. The IR spectrum of the crystals shows that the strong hydronium-crown ether/oxygen hydrogen bond absorption is found in the region 2300–2400 cm−1. The chemical shift in the range 9–12 ppm was observed. The 1H NMR signal of hydronium protons appears at 9.890 ppm. The results of assay correspond to the formula Ia2·(H3O+)2·UO2Cl42−.Crystal structure of the extraction complex has been determined by X-ray crystallography. Crystals are monoclinic, space group C2/c (No. 15) a=32.464, b=10.203, c=21.616 Å, β=119.73° and Z=4. In the complex each of the two H3O+ cations is anchored in the crown-ether cavity by three stronger hydrogen bonds (distances approximately 2.65 Å), whereas uranium forms UO2Cl42− with Cl as counterion about 8 Å away from the H3O+.  相似文献   

18.
《Free radical research》2013,47(1-2):77-83
Nitro-tyrosine considerably promotes the degradation of DNA, when incubated with Cu2+ and ascorbate in oxygenated aqueous solution. This deleterious process requires oxygen and can be inhibited with catalase, indicating that H2O2 is involved, via the reduction of oxygen. Menadione and 2,4,6-trinitro-benzenesulfonate, known to catalyze particularly fast such reduction of oxygen, were only slightly more active than nitro-tyrosine. Degradation of DNA can be explained by a site-specific Fenton type reaction of H2O2 with the DNA-Cu+ complex.

DNA-Cu+ + H2O2 → DNA' ' 'OH + Cu2+ + OH?

Copper-chelating agents (EDTA and penicillamine) prevent DNA degradation, whereas OH-scavengers (t-butanol) are ineffective. The deleterious activity of nitro-tyrosine (and of other nitroaromatics) in the DNA model system may indicate important toxicological implications, since aromatic nitration is a significant mode of action of nitrogen dioxide.  相似文献   

19.
G. Bottu 《Luminescence》1989,3(2):59-65
The chemiluminescence of luminol and lucigenin is often used to detect the production of reactive oxygen derivatives by phagocytic cells. Also, several quenchers and enzyme inhibitors are used to determine which oxygen derivatives are responsible for the observed effects. In the present work we have assessed the reliability of dimethylthiourea and cysteamine (OH. quenchers), desferrioxamine (iron chelator) and diethyldithiocarbamate (superoxide dismutase inhibitor). They all react with CIO? and are also strong inhibitors of the luminescence of luminol catalysed by horseradish peroxidase (HRP); cysteamine and diethyldithiocarbamate also react with H2O2. NaN3 is an inhibitor of myeloperoxidase and a quencher of singlet O2, but we found that under certain conditions it can amplify the the luminescence of luminol triggered by CIO? or Fenton's reagent. A complex of copper and penicillamine that had been proposed as an $ {\rm O}_{\rm 2} ^{\bar .} $ quencher, quenches all luminescent reactions studied. On the other hand, we were able to confirm the relative specificity of other quenchers: taurine for CIO?, benzoate for OH. and mannitol for both OH. and ‘crypto-OH.’.  相似文献   

20.
In a chilling-sensitive plant, cucumber, chilling of leaves in the light results in irreversible damage to PSI. Recent in vitro studies suggested that hydroxyl radicals, which are formed in the presence of H2O2 and reduced Fe-S centers, are involved in the PSI inhibition. We therefore examined this possibility in vivo. Chilling of leaves at 5°C in the light caused a temporary increase in H2O2 concentration, which was probably due to the net H2O2 production in vivo. The activity, measured at 5°C, of the thylakoid ascorbate peroxidase (APX), a key enzyme of the H2O2-scavenging system, was about 20% of that measured at 25°C. The isolated thylakoids retaining high thylakoid APX activity did not show light-dependent net H2O2 production at 25°C. However, at 5°C, net production of H2O2 was observed. Since the rate of electron flow to molecular oxygen in the isolated thylakoids was ca 5 mmol e? mol?1 Chl s?1 at 5°C, the H2O2-scavenging capacity was below this level. When intact leaves were illuminated at 5°C at an irradiance of 100 µmol m?2 s?1, the rate of electron transport through PSII was ca 20 mmol e? mol?1 Chl s?1 and more than 80% of QA was in the reduced state. Since thylakoids are uncoupled in cucumber leaves at 5°C in the light. ATP is not formed and energy dissipation in the form of heat is suppressed. Therefore, the electron flow to molecular oxygen would be greater than 5 mmol e? mol?1 Chl s?1. Moreover, under such conditions, components in the electron transport chain, including Fe-S centers in PSI, were probably reduced. These features indicate that, when cucumber leaves are chilled in the light, hydroxyl radicals can be produced by the Fenton reaction and cause damage to PSI.  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