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.
2.
We report on a rather unknown feature of oligonucleotides, namely, their potent antioxidant activity. Previously, we showed that nucleotides are potent antioxidants in FeII/CuI/II–H2O2 systems. Here, we explored the potential of 2′-deoxyoligonucleotides as inhibitors of the FeII/CuI/II-induced ·OH formation from H2O2. The oligonucleotides [d(A)5,7,20; d(T)20; (2′-OMe-A)5] proved to be highly potent antioxidants with IC50 values of 5–17 or 48–85 μM in inhibiting FeII/CuI- or CuII-induced H2O2 decomposition, respectively, thus representing a 40–215-fold increase in potency as compared with Trolox, a standard antioxidant. The antioxidant activity is only weakly dependent on the oligonucleotides’ length or base identity. We analyzed by matrix-assisted laser desorption/ionization time of flight mass spectrometry and 1H-NMR spectroscopy the composition of the d(A)5 solution exposed to the aforementioned oxidative conditions for 4 min or 24 h. We concluded that the primary (rapid) inhibition mechanism by oligonucleotides is metal ion chelation and the secondary (slow) mechanism is radical scavenging. We characterized the CuI–d(A)5 and CuII–d(A)7 complexes by 1H-NMR and 31P-NMR or frozen-solution ESR spectroscopy, respectively. CuI is probably coordinated to d(A)5 via N1 and N7 of two adenine residues and possibly also via two phosphate/bridging water molecules. The ESR data suggest CuII chelation through two nitrogen atoms of the adenine bases and two oxygen atoms (phosphates or water molecules). We conclude that oligonucleotides at micromolar concentrations prevent FeII/CuI/II-induced oxidative damage, primarily through metal ion chelation. Furthermore, we propose the use of a short, metabolically stable oligonucleotide, (2′-OMe-A)5, as a highly potent and relatively long lived (t 1/2 ~ 20 h) antioxidant.  相似文献   

3.
1. Rubredoxin isolated from the green photosynthetic bacterium Chloropseudomonas ethylica was similar in composition to those from anaerobic fermentative bacteria. Amino acid analysis indicated a minimum molecular weight of 6352 with one iron atom per molecule. 2. The circular-dichroism and electron-paramagnetic-resonance spectra of Ch. ethylica rubredoxin showed many similarities to those of Clostridium pasteurianum, but suggested that there may be subtle differences in the protein conformation about the iron atom. 3. Mössbauer-effect measurements on rubredoxin from Cl. pasteurianum and Ch. ethylica showed that in the oxidized state the iron (high-spin Fe3+) has a hyperfine field of 370±3kG, whereas in the reduced state (high-spin Fe2+) the hyperfine field tensor is anisotropic with a component perpendicular to the symmetry axis of the ion of about −200kG. For the reduced protein the sign of the electric-field gradient is negative, i.e. the ground state of the Fe2+ is a [unk] orbital. There is a large non-cubic ligand-field splitting (Δ/k=900°K), and a small spin-orbit splitting (D~+4.4cm−1) of the Fe2+ levels. 4. The contributions of core polarization to the hyperfine field in the Fe3+ and Fe2+ ions are estimated to be −370 and −300kG respectively. 5. The significance of these results in interpretation of the Mössbauer spectra of other iron–sulphur proteins is discussed.  相似文献   

4.
HbA O2 reacts readily with FeII(CN)5H2O3? to form aquometHb and peroxide via a second order process: rate=k[HbO2][FeII(CN)5H2O3?]. A slight enchancement in the rate of metHb formation due to the H2O2 produced can be prevented by addition of catalase. The reaction is free from complications exhibited by other reductants. The hexacyanide, ferrocyanide, reacts with HbA O2 but at only ca. 0.02% the rate and with formation of cyanometHb. Reductants such as phenols and sulfa drugs may produce radicals that can enter into side reactions. FeII(CN)5H2O3? shows promise as an effective probing reagent for the characterization of H2O2 production from oxygenated heme and other proteins.  相似文献   

5.
Volume-sensitive outwardly rectifying (VSOR) Cl channels are critical for the regulatory volume decrease (RVD) response triggered upon cell swelling. Recent evidence indicates that H2O2 plays an essential role in the activation of these channels and that H2O2 per se activates the channels under isotonic isovolumic conditions. However, a significant difference in the time course for current onset between H2O2-induced and hypotonicity-mediated VSOR Cl activation is observed. In several cell types, cell swelling induced by hypotonic challenges triggers the release of ATP to the extracellular medium, which in turn, activates purinergic receptors and modulates cell volume regulation. In this study, we have addressed the effect of purinergic receptor activation on H2O2-induced and hypotonicity-mediated VSOR Cl current activation. Here we show that rat hepatoma cells (HTC) exposed to a 33% hypotonic solution responded by rapidly activating VSOR Cl current and releasing ATP to the extracellular medium. In contrast, cells exposed to 200 μm H2O2 VSOR Cl current onset was significantly slower, and ATP release was not detected. In cells exposed to either 11% hypotonicity or 200 μm H2O2, exogenous addition of ATP in the presence of extracellular Ca2+ resulted in a decrease in the half-time for VSOR Cl current onset. Conversely, in cells that overexpress a dominant-negative mutant of the ionotropic receptor P2X4 challenged with a 33% hypotonic solution, the half-time for VSOR Cl current onset was significantly slowed down. Our results indicate that, at high hypotonic imbalances, swelling-induced ATP release activates the purinergic receptor P2X4, which in turn modulates the time course of VSOR Cl current onset in a extracellular Ca2+-dependent manner.  相似文献   

6.
[FeFe] hydrogenases are H2-evolving enzymes that feature a diiron cluster in their active site (the [2Fe]H cluster). One of the iron atoms has a vacant coordination site that directly interacts with H2, thus favoring its splitting in cooperation with the secondary amine group of a neighboring, flexible azadithiolate ligand. The vacant site is also the primary target of the inhibitor O2. The [2Fe]H cluster can span various redox states. The active-ready form (Hox) attains the FeIIFeI state. States more oxidized than Hox were shown to be inactive and/or resistant to O2. In this work, we used density functional theory to evaluate whether azadithiolate-to-iron coordination is involved in oxidative inhibition and protection against O2, a hypothesis supported by recent results on biomimetic compounds. Our study shows that Fe–N(azadithiolate) bond formation is favored for an FeIIFeII active-site model which disregards explicit treatment of the surrounding protein matrix, in line with the case of the corresponding FeIIFeII synthetic system. However, the study of density functional theory models with explicit inclusion of the amino acid environment around the [2Fe]H cluster indicates that the protein matrix prevents the formation of such a bond. Our results suggest that mechanisms other than the binding of the azadithiolate nitrogen protect the active site from oxygen in the so-called H ox inact state.  相似文献   

7.
1. The Mössbauer spectra of Scenedesmus ferredoxin enriched in 57Fe were measured and found to be identical with those of two other plant-type ferredoxins (from spinach and Euglena) that had been previously measured. Better resolved Mössbauer spectra of spinach ferredoxin are also reported from protein enriched in 57Fe. All these iron–sulphur proteins are known to contain two iron atoms in a molecule that takes up one electron on reduction. 2. The Mössbauer spectra at 195°K have electric hyperfine structure only and show that on reduction the electron goes to one of the iron atoms, the other appearing to remain unchanged. 3. In the oxidized state, both iron atoms are in a similar chemical state, which appears from the chemical shift and quadrupole splitting to be high-spin Fe3+, but they are in slightly different environments. In the reduced state the iron atoms are different and the molecule appears to contain one high-spin Fe2+ and one high-spin Fe3+ atom. 4. At lower temperatures (77 and 4.2°K) the spectra of both iron atoms in the reduced proteins show magnetic hyperfine structure which suggests that the iron in the oxidized state also has unpaired electrons. This provides experimental evidence for earlier suggestions that in the oxidized state there is antiferromagnetic exchange coupling, which would result in a low value for the magnetic susceptibility. 5. In a small magnetic field the spectrum of the reduced ferredoxin shows a Zeeman splitting with hyperfine field (Hn) of 180kG at the nuclei. On application of a strong magnetic field H the spectrum splits into two spectra with effective fields Hn±H, thus confirming the presence of the two antiferromagnetically coupled iron atoms. 6. These results are in agreement with the model proposed by Gibson, Hall, Thornley & Whatley (1966); in the oxidized state there are two Fe3+ atoms (high spin) antiferromagnetically coupled and on reduction of the ferredoxin by one electron one of the ferric atoms becomes Fe2+ (high spin).  相似文献   

8.
Metals such as CuI and FeII generate hydroxyl radical (OH) by reducing endogenous hydrogen peroxide (H2O2). Because antioxidants can ameliorate metal-mediated oxidative damage, we have quantified the ability of glutathione, a primary intracellular antioxidant, and other biological sulfur-containing compounds to inhibit metal-mediated DNA damage caused hydroxyl radical. In the CuI/H2O2 system, six sulfur compounds, including both reduced and oxidized glutathione, inhibited DNA damage with IC50 values ranging from 3.4 to 12.4 μM. Glutathione and 3-carboxypropyl disulfide also demonstrated significant antioxidant activity with FeII and H2O2. Additional gel electrophoresis and UV-vis spectroscopy studies confirm that antioxidant activity for sulfur compounds in the CuI system is attributed to metal coordination, a previously unexplored mechanism. The antioxidant mechanism for sulfur compounds in the FeII system, however, is unlike that of CuI. Our results demonstrate that glutathione and other sulfur compounds are potent antioxidants capable of preventing metal-mediated oxidative DNA damage at well below their biological concentrations. This novel metal-binding antioxidant mechanism may play a significant role in the antioxidant behavior of these sulfur compounds and help refine understanding of glutathione function in vivo.  相似文献   

9.

Background and Aims

Experimental evidence in the literature suggests that O2•− produced in the elongation zone of roots and leaves by plasma membrane NADPH oxidase activity is required for growth. This study explores whether growth changes along the root tip induced by hyperosmotic treatments in Zea mays are associated with the distribution of apoplastic O2•−.

Methods

Stress treatments were imposed using 150 mm NaCl or 300 mm sorbitol. Root elongation rates and the spatial distribution of growth rates in the root tip were measured. Apoplastic O2•− was determined using nitro blue tetrazolium, and H2O2 was determined using 2′, 7′-dichlorofluorescin.

Key Results

In non-stressed plants, the distribution of accelerating growth and highest O2•− levels coincided along the root tip. Salt and osmotic stress of the same intensity had similar inhibitory effects on root elongation, but O2•− levels increased in sorbitol-treated roots and decreased in NaCl-treated roots.

Conclusions

The lack of association between apoplastic O2•− levels and root growth inhibition under hyper-osmotic stress leads us to hypothesize that under those conditions the role of apoplastic O2•− may be to participate in signalling processes, that convey information on the nature of the substrate that the growing root is exploring.Key words: Root tip growth, Zea mays, salt stress, reactive oxygen species, ROS  相似文献   

10.
Adenovirus expressing ClC-3 (Ad-ClC-3) induces Cl/H+ antiport current (IClC-3) in HEK293 cells. The outward rectification and time dependence of IClC-3 closely resemble an endogenous HEK293 cell acid-activated Cl current (IClacid) seen at extracellular pH ≤ 5.5. IClacid was present in smooth muscle cells from wild-type but not ClC-3 null mice. We therefore sought to determine whether these currents were related. IClacid was larger in cells expressing Ad-ClC-3. Protons shifted the reversal potential (Erev) of IClC-3 between pH 8.2 and 6.2, but not pH 6.2 and 5.2, suggesting that Cl and H+ transport become uncoupled at low pH. At pH 4.0 Erev was completely Cl dependent (55.8 ± 2.3 mV/decade). Several findings linked ClC-3 with native IClacid; 1) RNA interference directed at ClC-3 message reduced native IClacid; 2) removal of the extracellular “fast gate” (E224A) produced large currents that were pH-insensitive; and 3) wild-type IClC-3 and IClacid were both inhibited by (2-sulfonatoethyl)methanethiosulfonate (MTSES; 10–500 μm)-induced alkanethiolation at exposed cysteine residues. However, a ClC-3 mutant lacking four extracellular cysteine residues (C103_P130del) was completely resistant to MTSES. C103_P130del currents were still acid-activated, but could be distinguished from wild-type IClC-3 and from native IClacid by a much slower response to low pH. Thus, ClC-3 currents are activated by protons and ClC-3 protein may account for native IClacid. Low pH uncouples Cl/H+ transport so that at pH 4.0 ClC-3 behaves as an anion-selective channel. These findings have important implications for the biology of Cl/H+ antiporters and perhaps for pH regulation in highly acidic intracellular compartments.  相似文献   

11.
Hydrogen peroxide (H2O2) is commonly formed in microbial habitats by either chemical oxidation processes or host defense responses. H2O2 can penetrate membranes and damage key intracellular biomolecules, including DNA and iron-dependent enzymes. Bacteria defend themselves against this H2O2 by inducing a regulon that engages multiple defensive strategies. A previous microarray study suggested that yaaA, an uncharacterized gene found in many bacteria, was induced by H2O2 in Escherichia coli as part of its OxyR regulon. Here we confirm that yaaA is a key element of the stress response to H2O2. In a catalase/peroxidase-deficient (Hpx) background, yaaA deletion mutants grew poorly, filamented extensively, and lost substantial viability when they were cultured in aerobic LB medium. The results from a thyA forward mutagenesis assay and the growth defect of the yaaA deletion in a recombination-deficient (recA56) background indicated that yaaA mutants accumulated high levels of DNA damage. The growth defect of yaaA mutants could be suppressed by either the addition of iron chelators or mutations that slowed iron import, indicating that the DNA damage was caused by the Fenton reaction. Spin-trapping experiments confirmed that Hpx yaaA cells had a higher hydroxyl radical (HO) level. Electron paramagnetic resonance spectroscopy analysis showed that the proximate cause was an unusually high level of intracellular unincorporated iron. These results demonstrate that during periods of H2O2 stress the induction of YaaA is a critical device to suppress intracellular iron levels; it thereby attenuates the Fenton reaction and the DNA damage that would otherwise result. The molecular mechanism of YaaA action remains unknown.  相似文献   

12.
Studies of the Uptake of Nitrate in Barley : IV. Electrophysiology   总被引:17,自引:5,他引:12       下载免费PDF全文
Transmembrane electrical potential differences (Δψ) of epidermal and cortical cells were measured in intact roots of barley (Hordeum vulgare L. cv Klondike). The effects of exogenous NO3 on Δψ (in the concentration range from 100 micromolar to 20 millimolar) were investigated to probe the mechanisms of nitrate uptake by the high-affinity (HATS) and low-affinity (LATS) transport systems for NO3 uptake. Both transport systems caused depolarization of Δψ, demonstrating that the LATS (like the HATS) for NO3 uptake is probably mediated by an electrogenic cation (H+?) cotransport system. Membrane depolarization by the HATS was “inducible” by NO3, and saturable with respect to exogenous [NO3]. By contrast, depolarization by the LATS was constitutive, and first-order in response to external [NO3]. H+ fluxes, measured in 200 micromolar and in 5 millimolar Ca(NO3)2 solutions, failed to alkalinize external media as anticipated for a 2 H+:1 NO3 symport. However, switching from K2SO4 solutions (which were strongly acidifying) to KNO3 solutions at the same K+ concentration caused marked reductions in H+ efflux. These observations are consistent with NO3 uptake by the HATS and the LATS via 2 H+:1 NO3 symports. These observations establish that the HATS for nitrate uptake by barley roots is essentially similar to those reported for Lemna and Zea mays by earlier workers. There are, nevertheless, distinct differences between barley and corn in their quantitative responses to external NO3.  相似文献   

13.
A simple strategy for the induction of extracellular hydroxyl radical (OH) production by white-rot fungi is presented. It involves the incubation of mycelium with quinones and Fe3+-EDTA. Succinctly, it is based on the establishment of a quinone redox cycle catalyzed by cell-bound dehydrogenase activities and the ligninolytic enzymes (laccase and peroxidases). The semiquinone intermediate produced by the ligninolytic enzymes drives OH production by a Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+). H2O2 production, Fe3+ reduction, and OH generation were initially demonstrated with two Pleurotus eryngii mycelia (one producing laccase and versatile peroxidase and the other producing just laccase) and four quinones, 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and 2-methyl-1,4-naphthoquinone (menadione [MD]). In all cases, OH radicals were linearly produced, with the highest rate obtained with MD, followed by DBQ, MBQ, and BQ. These rates correlated with both H2O2 levels and Fe3+ reduction rates observed with the four quinones. Between the two P. eryngii mycelia used, the best results were obtained with the one producing only laccase, showing higher OH production rates with added purified enzyme. The strategy was then validated in Bjerkandera adusta, Phanerochaete chrysosporium, Phlebia radiata, Pycnoporus cinnabarinus, and Trametes versicolor, also showing good correlation between OH production rates and the kinds and levels of the ligninolytic enzymes expressed by these fungi. We propose this strategy as a useful tool to study the effects of OH radicals on lignin and organopollutant degradation, as well as to improve the bioremediation potential of white-rot fungi.White-rot fungi are unique in their ability to degrade a wide variety of organopollutants (36, 47), mainly due to the secretion of a low-specificity enzyme system whose natural function is the degradation of lignin (11). Components of this system include laccase and/or one or two types of peroxidase, such as lignin peroxidase (LiP), manganese peroxidase (MnP), and versatile peroxidase (VP) (31). Besides acting directly, the ligninolytic enzymes can bring about lignin and pollutant degradation through the generation of low-molecular-weight extracellular oxidants, including (i) Mn3+, (ii) free radicals from some fungal metabolites and lignin depolymerization products (7, 22), and (iii) oxygen free radicals, mainly hydroxyl radicals (OH) and lipid peroxidation radicals (21). Although OH radicals are the strongest oxidants found in cultures of white-rot fungi (1), studies of their involvement in pollutant degradation are scarce. One of the reasons is that the mechanisms proposed for OH production still await in vivo validation.Several potential sources of extracellular OH based on the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+) have been postulated for white-rot fungi. In one case, an extracellular fungal glycopeptide has been shown to reduce O2 and Fe3+ to H2O2 and Fe2+ (45). Enzymatic sources include cellobiose dehydrogenase, LiP, and laccase. Among these, only cellobiose dehydrogenase is able to directly catalyze the formation of Fenton''s reagent (33). The ligninolytic enzymes, however, act as an indirect source of OH through the generation of Fe3+ and O2 reductants, such as formate (CO2) and semiquinone (Q) radicals. The first time evidence was provided that a ligninolytic enzyme was involved in OH production, oxalate was used to generate CO2 in a LiP reaction mediated by veratryl alcohol (4). The proposed mechanism consisted of the following cascade of reactions: production of veratryl alcohol cation radical (Valc+) by LiP, oxidation of oxalate to CO2 by Valc+, reduction of O2 to O2 by CO2, and a superoxide-driven Fenton reaction (Haber-Weiss reaction) in which Fe3+ was reduced by O2. The OH production mechanism assisted by Q was inferred from the oxidation of 2-methoxy-1,4-benzohydroquinone (MBQH2) and 2,6-dimethoxy-1,4-benzohydroquinone (DBQH2) by Pleurotus eryngii laccase in the presence of Fe3+-EDTA. The ability of Q radicals to reduce both Fe3+ to Fe2+ and O2 to O2, which dismutated to H2O2, was demonstrated (14). In this case, OH radicals were generated by a semiquinone-driven Fenton reaction, as Q radicals were the main agents accomplishing Fe3+ reduction. The first evidence of the likelihood of this OH production mechanism being operative in vivo had been obtained from incubations of P. eryngii with 2-methyl-1,4-naphthoquinone (menadione [MD]) and Fe3+-EDTA (15). Extracellular OH radicals were produced on a constant basis through quinone redox cycling, consisting of the reduction of MD by a cell-bound quinone reductase (QR) system, followed by the extracellular oxidation of the resulting hydroquinone (MDH2) to its semiquinone radical (MD). The production of extracellular O2 and H2O2 by P. eryngii via redox cycling involving laccase was subsequently confirmed using 1,4-benzoquinone (BQ), 2-methyl-1,4-benzoquinone, and 2,3,5,6-tetramethyl-1,4-benzoquinone (duroquinone), in addition to MD (16). However, the demonstration of OH production based on the redox cycling of quinones other than MD was still required.In the present paper, we describe the induction of extracellular OH production by P. eryngii upon its incubation with BQ, 2-methoxy-1,4-benzoquinone (MBQ), 2,6-dimethoxy-1,4-benzoquinone (DBQ), and MD in the presence of Fe3+-EDTA. The three benzoquinones were selected because they are oxidation products of p-hydroxyphenyl, guaiacyl, and syringyl units of lignin (MD was included as a positive control). Along with laccase, the involvement of P. eryngii VP in the production of O2 and H2O2 from hydroquinone oxidation has also been reported (13). Since hydroquinones are substrates of all known ligninolytic enzymes, quinone redox cycling catalysis could involve any of them. Here, we demonstrate OH production by P. eryngii under two different culture conditions, leading to the production of laccase or laccase and VP. We also show that quinone redox cycling is widespread among white-rot fungi by using a series of well-studied species that produce different combinations of ligninolytic enzymes.  相似文献   

14.
L-amino acid oxidase (LAAO) has important biological roles in many organisms, thus attracting great attention from researchers to establish its detection methods. In this study, a new quantitative in-gel determination of LAAO activity based on ferric-xylenol orange (FeIIIXO) formation was established. This method showed that due to the conversion of FeII to FeIII by H2O2 and subsequent formation of FeIIIXO complex halo in agar medium, the logarithm of H2O2 concentration from 5 to 160 µM was linearly correlated to the diameter of purplish red FeIIIXO halo. By extracting the LAAO-generated H2O2 concentration, the LAAO activity can be quantitatively determined. This FeIIIXO agar assay is highly sensitive to detect H2O2 down to micromolar range. More importantly, it is easy to handle, cheap, reproducible, convenient and accurate. Coupled with SDS-PAGE, it can directly be used to determine the number and approximate molecular weight of LAAO in one assay. All these features make this in-gel FeIIIXO assay useful and convenient as a general procedure for following enzyme purification, assaying fractions from a column, or observing changes in activity resulting from enzyme modifications, hence endowing this method with broad applications.  相似文献   

15.
A new tri-cyanometalate building block for heterometallic complexes, [PPh4]2[FeII(Tpms)(CN)3] (2) (PPh4 = tetraphenylphosphonium; Tpms = tris(pyrazolyl) methanesulfonate), has been prepared. Using it as a building block, a one-dimensional chain compound, {[FeII(Tpms)(CN)3][MnII(H2O)2( DMF)2]} · DMF (3), has been synthesized and structurally characterized. The magnetic properties of 3 correspond to a ferromagnetic chain with weak long-range superexchanged magnetic interaction between the high-spin manganese(II) ions.  相似文献   

16.
Liu YL  Chiang YH  Liu GY  Hung HC 《PloS one》2011,6(6):e21314
Peptidylarginine deiminase 4 (PAD4) is a homodimeric enzyme that catalyzes Ca2+-dependent protein citrullination, which results in the conversion of arginine to citrulline. This paper demonstrates the functional role of dimerization in the regulation of PAD4 activity. To address this question, we created a series of dimer interface mutants of PAD4. The residues Arg8, Tyr237, Asp273, Glu281, Tyr435, Arg544 and Asp547, which are located at the dimer interface, were mutated to disturb the dimer organization of PAD4. Sedimentation velocity experiments were performed to investigate the changes in the quaternary structures and the dissociation constants (K d) between wild-type and mutant PAD4 monomers and dimers. The kinetic data indicated that disrupting the dimer interface of the enzyme decreases its enzymatic activity and calcium-binding cooperativity. The K d values of some PAD4 mutants were much higher than that of the wild-type (WT) protein (0.45 µM) and were concomitant with lower k cat values than that of WT (13.4 s−1). The K d values of the monomeric PAD4 mutants ranged from 16.8 to 45.6 µM, and the k cat values of the monomeric mutants ranged from 3.3 to 7.3 s−1. The k cat values of these interface mutants decreased as the K d values increased, which suggests that the dissociation of dimers to monomers considerably influences the activity of the enzyme. Although dissociation of the enzyme reduces the activity of the enzyme, monomeric PAD4 is still active but does not display cooperative calcium binding. The ionic interaction between Arg8 and Asp547 and the Tyr435-mediated hydrophobic interaction are determinants of PAD4 dimer formation.  相似文献   

17.
The asymmetrically coordinated complex [{L(Ph2acac)FeIII}(μ-O){FeIII(Cl4-cat)L}](BPh4)·1.5toluene has been synthesized and structurally characterized (Ph2acac=1,3-diphenylpropane-1,3-dionate, Cl4-cat2–=tetrachlorocatecholate, L=1,4,7-trimethyl-1,4,7-triazacyclononane). This species can be electrochemically oxidized and reduced by one electron, respectively, yielding two species which both have an S=1/2 ground state. It is shown that the oxidation is ligand-centered, affording a coordinated semiquinonate(1–) ligand with S=1/2 which is antiferromagnetically coupled to a high-spin FeIII ion (S=5/2) yielding an S=2 state which, in turn, is antiferromagnetically coupled (through the oxo bridge) to the second high-spin FeIII ion (S=5/2) yielding the observed S=1/2 ground state. In contrast, the reduction is metal-centered generating a mixed-valent species with an [FeIII-O-FeII]3+ core; intramolecular antiferromagnetic coupling again produces an S=1/2 ground state. The symmetrical complex [{LFeIII(Ph2acac)}2(μ-O)](ClO4)2 has also been synthesized, as have the mononuclear species [LFeII(Ph2acac)Cl] and [LFeIII(aacac)Cl](ClO4)·1 mesitylene [aacac=3-(9-anthryl)acetylacetonate(1–)], all of which have been characterized by X-ray crystallography. The magnetism, the Mössbauer-, EPR-, and UV-VIS-spectra and the electrochemistry of complexes are reported.  相似文献   

18.
The induction of hydroxyl radical (OH) production via quinone redox cycling in white-rot fungi was investigated to improve pollutant degradation. In particular, we examined the influence of 4-methoxybenzaldehyde (anisaldehyde), Mn2+, and oxalate on Pleurotus eryngii OH generation. Our standard quinone redox cycling conditions combined mycelium from laccase-producing cultures with 2,6-dimethoxy-1,4-benzoquinone (DBQ) and Fe3+-EDTA. The main reactions involved in OH production under these conditions have been shown to be (i) DBQ reduction to hydroquinone (DBQH2) by cell-bound dehydrogenase activities; (ii) DBQH2 oxidation to semiquinone (DBQ) by laccase; (iii) DBQ autoxidation, catalyzed by Fe3+-EDTA, producing superoxide (O2) and Fe2+-EDTA; (iv) O2 dismutation, generating H2O2; and (v) the Fenton reaction. Compared to standard quinone redox cycling conditions, OH production was increased 1.2- and 3.0-fold by the presence of anisaldehyde and Mn2+, respectively, and 3.1-fold by substituting Fe3+-EDTA with Fe3+-oxalate. A 6.3-fold increase was obtained by combining Mn2+ and Fe3+-oxalate. These increases were due to enhanced production of H2O2 via anisaldehyde redox cycling and O2 reduction by Mn2+. They were also caused by the acceleration of the DBQ redox cycle as a consequence of DBQH2 oxidation by both Fe3+-oxalate and the Mn3+ generated during O2 reduction. Finally, induction of OH production through quinone redox cycling enabled P. eryngii to oxidize phenol and the dye reactive black 5, obtaining a high correlation between the rates of OH production and pollutant oxidation.The degradation of lignin and pollutants by white-rot fungi is an oxidative and rather nonspecific process based on the production of substrate free radicals (36). These radicals are produced by ligninolytic enzymes, including laccase and three kinds of peroxidases: lignin peroxidase, manganese peroxidase, and versatile peroxidase (VP) (23). The H2O2 required for peroxidase activities is provided by several oxidases, such as glyoxal oxidase and aryl-alcohol oxidase (AAO) (9, 18). This free-radical-based degradative mechanism leads to the production of a broad variety of oxidized compounds. Common lignin depolymerization products are aromatic aldehydes and acids, and quinones (34). In addition to their high extracellular oxidation potential, white-rot fungi show strong ability to reduce these lignin depolymerization products, using different intracellular and membrane-bound systems (4, 25, 39). Since reduced electron acceptors of oxidized compounds are donor substrates for the above-mentioned oxidative enzymes, the simultaneous actions of both systems lead to the establishment of redox cycles (35). Although the function of these redox cycles is not fully understood, they have been hypothesized to be related to further metabolism of lignin depolymerization products that require reduction to be converted in substrates of the ligninolytic enzymes (34). A second function attributed to these redox cycles is the production of reactive oxygen species, i.e., superoxide anion radicals (O2), H2O2, and hydroxyl radicals (OH), where lignin depolymerization products and fungal metabolites act as electron carriers between intracellular reducing equivalents and extracellular oxygen. This function has been studied in Pleurotus eryngii, whose ligninolytic system is composed of laccase (26), VP (24), and AAO (9). Incubation of this fungus with different aromatic aldehydes has been shown to provide extracellular H2O2 on a constant basis, due to the establishment of a redox cycle catalyzed by an intracellular aryl-alcohol dehydrogenase (AAD) and the extracellular AAO (7, 10). The process was termed aromatic aldehyde redox cycling, and 4-methoxybenzaldehyde (anisaldehyde) serves as the main Pleurotus metabolite acting as a cycle electron carrier (13). A second cyclic system, involving a cell-bound quinone reductase activity (QR) and laccase, was found to produce O2 and H2O2 during incubation of P. eryngii with different quinones (11). The process was described as the cell-bound divalent reduction of quinones (Q) by QR, followed by extracellular laccase oxidation of hydroquinones (QH2) into semiquinones (Q), which autoxidized to some extent, producing O2 (Q + O2 ⇆ Q + O2). H2O2 was formed by O2 dismutation (O2 + HO2 + H+ → O2 + H2O2). In an accompanying paper, we describe the extension of this O2 and H2O2 generation mechanism to OH radical production by the addition of Fe3+-EDTA to incubation mixtures of several white-rot fungi with different quinones (6). Among them, those derived from 4-hydroxyphenyl, guaiacyl, and syringyl lignin units were used: 1,4-benzoquinone (BQ), 2-methoxy-1,4-benzoquinone (MBQ), and 2,6-dimethoxy-1,4-benzoquinone (DBQ), respectively. Semiquinone autoxidation under these conditions was catalyzed by Fe3+-EDTA instead of being a direct electron transfer to O2. The intermediate Fe2+-EDTA reduced not only O2, but also H2O2, leading to OH radical production by the Fenton reaction (H2O2 + Fe2+ → OH + OH + Fe3+).Although OH radicals are the strongest oxidants produced by white-rot fungi (2, 14), studies of their involvement in pollutant degradation are quite scarce. In this context, the objectives of this study were to (i) determine possible factors enhancing the production of OH radicals by P. eryngii via quinone redox cycling and (ii) test the validity of this inducible OH production mechanism as a strategy for pollutant degradation. Our selection of possible OH production promoters was guided by two observations (6). First, the redox cycle of benzoquinones working with washed P. eryngii mycelium is rate limited by hydroquinone oxidation, since the amounts of the ligninolytic enzymes that remained bound to the fungus under these conditions were not large. Second, H2O2 is the limiting reagent for OH production by the Fenton reaction.With these considerations in mind, anisaldehyde and Mn2+ were selected to increase H2O2 production. As mentioned above, anisaldehyde induces H2O2 production in P. eryngii via aromatic aldehyde redox cycling (7). Mn2+ has been shown to enhance H2O2 production during the oxidation of QH2 by P. eryngii laccase by reducing the O2 produced in the semiquinone autoxidation reaction (Mn2+ + O2 → Mn3+ + H2O2 + 2 H+) (26). Mn2+ was also selected to increase the hydroquinone oxidation rate, since this reaction has been shown to be propagated by the Mn3+ generated in the latter reaction (QH2 + Mn3+ → Q + Mn2+ + 2 H+). The replacement of Fe3+-EDTA by Fe3+-oxalate was also planned in order to increase the QH2 oxidation rate above that resulting from the action of laccase. Oxalate is a common extracellular metabolite of wood-rotting fungi to which the function of chelating iron and manganese has been attributed (16, 45). The use of Fe3+-oxalate and nonchelated Fe3+, both QH2 oxidants, has been proven to enable quinone redox cycling in fungi that do not produce ligninolytic enzymes, such as the brown-rot fungus Gloeophyllum trabeum (17, 40, 41). Finally, phenol and the azo dye reactive black 5 (RB5) were selected as model pollutants.  相似文献   

19.
Three new chiral ligands bearing an O,O′,N donor set (OmethoxyOhydroxyNpyridine) were synthesised and coordinated to FeIII, FeII, NiII, CuII and ZnII to yield complexes with the general formula [M(OON)Clx]y. While the pyridine N and the hydroxy O atoms coordinate strongly to all applied metal ions, the methoxy donor seems not to be involved in coordination, although some evidence for a weak interaction between OMe and the ZnII were found in NMR spectra. In the bidentate O′,N coordination mode the new ligands exhibit several coordination geometries as analysed in the solid compounds by XRD, EXAFS and EPR and in solution by UV-Vis absorption, cyclic voltammetry, EXAFS, EPR or NMR spectroscopy.  相似文献   

20.
We examined the effect of reactive oxygen species (ROS) on MicroRNAs (miRNAs) expression in endothelial cells in vitro, and in mouse skeletal muscle following acute hindlimb ischemia. Human umbilical vein endothelial cells (HUVEC) were exposed to 200 μM hydrogen peroxide (H2O2) for 8 to 24 h; miRNAs profiling showed that miR-200c and the co-transcribed miR-141 increased more than eightfold. The other miR-200 gene family members were also induced, albeit to a lower level. Furthermore, miR-200c upregulation was not endothelium restricted, and occurred also on exposure to an oxidative stress-inducing drug: 1,3-bis(2 chloroethyl)-1nitrosourea (BCNU). miR-200c overexpression induced HUVEC growth arrest, apoptosis and senescence; these phenomena were also induced by H2O2 and were partially rescued by miR-200c inhibition. Moreover, miR-200c target ZEB1 messenger RNA and protein were downmodulated by H2O2 and by miR-200c overexpression. ZEB1 knockdown recapitulated miR-200c-induced responses, and expression of a ZEB1 allele non-targeted by miR-200c, prevented miR-200c phenotype. The mechanism of H2O2-mediated miR-200c upregulation involves p53 and retinoblastoma proteins. Acute hindlimb ischemia enhanced miR-200c in wild-type mice skeletal muscle, whereas in p66ShcA −/− mice, which display lower levels of oxidative stress after ischemia, upregulation of miR-200c was markedly inhibited. In conclusion, ROS induce miR-200c and other miR-200 family members; the ensuing downmodulation of ZEB1 has a key role in ROS-induced apoptosis and senescence.  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