首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Exchange of iron by gallium in siderophores   总被引:2,自引:0,他引:2  
T Emery 《Biochemistry》1986,25(16):4629-4633
Siderophores are iron transport compounds produced by numerous microorganisms and which strongly chelate Fe(III), but not Fe(II). Other trivalent metals, such as Al(III), Cr(III), or Ga(III), are not capable of significantly displacing iron from siderophores. However, I demonstrate here that Ga(III) can effectively displace iron under reducing conditions. With ascorbate as reductant and ferrozine as Fe(II) trapping agent, the kinetics of reductive displacement of iron by Ga(III) were followed spectroscopically by the increase of absorbance at 562 nm due to formation of the Fe(II)-ferrozine complex. No significant reduction of siderophore occurred in the absence of Ga(III). With excess Ga(III), the displacement was quantitative and very rapid. The rate of metal exchange was pseudo first order with respect to Ga(III) concentration and highly pH dependent, suggesting that siderophore ligands are displaced from the iron in a concerted mechanism by Ga(III) and protonation to expose the Fe(III) to reduction by ascorbate. Reaction rates were dependent upon the structure of the siderophore, being greatest for ferric rhodotorulic acid and slowest for ferrichrome A at pH 5.4. The pH profile for ferric rhodotorulic acid was unusual in that it showed a maximum at pH 6.5, while all other siderophores examined showed an increase in rate as pH was lowered from 7.0. The physiological significance of this reaction to the clinical use of gallium is discussed.  相似文献   

2.
1. A comparative study shows that ferrous ions give a much better yield of Fe(III)-bleomycin than ferric ions, when iron salt is added to bleomycin in a buffer solution (pH 7.2). 2. The amount of Fe(III)-bleomycin formed after addition of ferric ions was markedly increased in the presence of ferric ion binding compounds (BSA, citrate) or reducing agents (ascorbate, cysteine).  相似文献   

3.
Fe(II)- and Fe(III)-induced lipid peroxidation of rabbit small intestinal microvillus membrane vesicles was studied. Ferrous ammonium sulphate, ferrous ascorbate at a molar ratio of 10:1, and ferric citrate, at molar ratios of 1:1 and 1:20, did not stimulate lipid peroxidation. Ferrous ascorbate, 1:1, induced low stimulation, while ferrous ascorbate, 1:20 gave higher stimulation of lipid peroxidation. These results show that in our experimental system, ascorbate is a promotor rather than an inhibitor of lipid peroxidation. Ferric nitrilotriacetate (at molar ratios of 1:2 and 1:10), at an iron concentration of 200 microM, was by far the most effective in inducing lipid peroxidation. Superoxide dismutase, mannitol and glutathione had no effect, while catalase, thiourea and vitamin E markedly decreased ferrous ascorbate 1:20-induced lipid peroxidation. Ferric nitrilotriacetate-induced lipid peroxidation was slightly reduced by catalase and mannitol, significantly reduced by superoxide dismutase, and completely inhibited by thiourea. Glutathione caused a 100% increase in the ferric nitrilotriacetate-induced lipid peroxidation. These results suggest that Fe(II) in the presence of trace amounts of Fe(III), or an oxidizing agent and Fe(III) in the presence of Fe(II) or a reducing agent, are potent stimulators of lipid peroxidation of microvillus membrane vesicles. Addition of deferoxamine completely inhibited both ferrous ascorbate, 1:20 and ferric nitrilotriacetate-induced lipid peroxidation, demonstrating the requirement for iron for its stimulation. Iron-induced peroxidation of microvillus membrane may have physiological significance because it could already be demonstrated at 2 microM iron concentration.  相似文献   

4.
Based on the ability of phytosiderophores to chelate other heavy metals besides iron (Fe), phytosiderophores were suggested to prevent graminaceous plants from cadmium (Cd) toxicity. To assess interactions between Cd and phytosiderophore-mediated Fe acquisition, maize (Zea mays) plants were grown hydroponically under limiting Fe supply. Exposure to Cd decreased uptake rates of 59Fe(III)-phytosiderophores and enhanced the expression of the Fe-phytosiderophore transporter gene ZmYS1 in roots as well as the release of the phytosiderophore 2'-deoxymugineic acid (DMA) from roots under Fe deficiency. However, DMA hardly mobilized Cd from soil or from a Cd-loaded resin in comparison to the synthetic chelators diaminetriaminepentaacetic acid and HEDTA. While nano-electrospray-high resolution mass spectrometry revealed the formation of an intact Cd(II)-DMA complex in aqueous solutions, competition studies with Fe(III) and zinc(II) showed that the formed Cd(II)-DMA complex was weak. Unlike HEDTA, DMA did not protect yeast (Saccharomyces cerevisiae) cells from Cd toxicity but improved yeast growth in the presence of Cd when yeast cells expressed ZmYS1. When supplied with Fe-DMA as a Fe source, transgenic Arabidopsis (Arabidopsis thaliana) plants expressing a cauliflower mosaic virus 35S-ZmYS1 gene construct showed less growth depression than wild-type plants in response to Cd. These results indicate that inhibition of ZmYS1-mediated Fe-DMA transport by Cd is not related to Cd-DMA complex formation and that Cd-induced phytosiderophore release cannot protect maize plants from Cd toxicity. Instead, phytosiderophore-mediated Fe acquisition can improve Fe uptake in the presence of Cd and thereby provides an advantage under Cd stress relative to Fe acquisition via ferrous Fe.  相似文献   

5.
Poaceae plants release 2′‐deoxymugineic acid (DMA) and related phytosiderophores to chelate iron (Fe), which often exists as insoluble Fe(III) in the rhizosphere, especially under high pH conditions. Although the molecular mechanisms behind the biosynthesis and secretion of DMA have been studied extensively, little information is known about whether DMA has biological roles other than chelating Fe in vivo. Here, we demonstrate that hydroponic cultures of rice (Oryza sativa) seedlings show almost complete restoration in shoot height and soil‐plant analysis development (SPAD) values after treatment with 3–30 μm DMA at high pH (pH 8.0), compared with untreated control seedlings at normal pH (pH 5.8). These changes were accompanied by selective accumulation of Fe over other metals. While this enhanced growth was evident under high pH conditions, DMA application also enhanced seedling growth under normal pH conditions in which Fe was fairly accessible. Microarray and qRT‐PCR analyses revealed that exogenous DMA application attenuated the increased expression levels of various genes related to Fe transport and accumulation. Surprisingly, despite the preferential utilization of ammonium over nitrate as a nitrogen source by rice, DMA application also increased nitrate reductase activity and the expression of genes encoding high‐affinity nitrate transporters and nitrate reductases, all of which were otherwise considerably lower under high pH conditions. These data suggest that exogenous DMA not only plays an important role in facilitating the uptake of environmental Fe, but also orchestrates Fe and nitrate assimilation for optimal growth under high pH conditions.  相似文献   

6.
Electrons utilized in the heme oxygenase (HO) reaction are provided by NADPH-cytochrome P450 reductase (CPR). To investigate the electron transfer pathway from CPR to HO, we examined the reactions of heme and verdoheme, the second intermediate in the heme degradation, complexed with rat HO-1 (rHO-1) using a rat FMN-depleted CPR; the FMN-depleted CPR was prepared by dialyzing the CPR mutant, Y140A/Y178A, against 2 m KBr. Degradation of heme in complex with rHO-1 did not occur with FMN-depleted CPR, notwithstanding that the FMN-depleted CPR was able to associate with the heme-rHO-1 complex with a binding affinity comparable with that of the wild-type CPR. Thus, the first electron to reduce the ferric iron of heme complexed with rHO-1 must be transferred from FMN. In contrast, verdoheme was converted to the ferric biliverdin-iron chelate with FMN-depleted CPR, and this conversion was inhibited by ferricyanide, indicating that electrons are certainly required for conversion of verdoheme to a ferric biliverdin-iron chelate and that they can be supplied from the FMN-depleted CPR through a pathway not involving FMN, probably via FAD. This conclusion was supported by the observation that verdoheme dimethyl esters were accumulated in the reaction of the ferriprotoporphyrin IX dimethyl ester-rHO-1 complex with the wild-type CPR. Ferric biliverdin-iron chelate, generated with the FMN-depleted CPR, was converted to biliverdin by the addition of the wild-type CPR or desferrioxamine. Thus, the final electron for reducing ferric biliverdin-iron chelate to release ferrous iron and biliverdin is apparently provided by the FMN of CPR.  相似文献   

7.
The in vitro effects of four different species of arsenic (arsenate, arsenite, monomethylarsonic acid, and dimethylarsinic acid) in mobilizing iron from horse spleen ferritin under aerobic and anaerobic conditions were investigated. Dimethylarsinic acid (DMA(V)) and dimethylarsinous acid (DMA(III)) significantly released iron from horse spleen ferritin either with or without the presence of ascorbic acid, a strong synergistic agent. Ascorbic acid-mediated iron release was time-dependent as well as both DMA(III) and ferritin concentration-dependent. Iron release from ferritin by DMA(III)) alone or with ascorbic acid was not significantly inhibited by superoxide dismutase (150 or 300 units/ml). However, the iron release was greater under anaerobic conditions (nitrogen gas), which indicates direct chemical reduction of iron from ferritin by DMA(III), with or without ascorbic acid. Both DMA(V) and DMA(III)) released iron from both horse spleen and human liver ferritin. Further, the release of ferritin iron by DMA(III)) with ascorbic acid catalyzed bleomycin-dependent degradation of calf thymus DNA. These results indicate that exogenous methylated arsenic species and endogenous ascorbic acid can cause (a) the release of iron from ferritin, (b) the iron-dependent formation of reactive oxygen species, and (c) DNA damage. This reactive oxygen species pathway could be a mechanism of action of arsenic carcinogenesis in man.  相似文献   

8.
The iron-binding properties of hen ovotransferrin.   总被引:1,自引:0,他引:1       下载免费PDF全文
1. The distribution of iron between the two iron-binding sites in partially saturated ovotransferrin was studied by labelling with 55Fe and 59Fe and by gel electrophoresis in a urea-containing buffer. 2. When iron is added in the form of chelate complexes at alkaline pH, binding occurs preferentially at the N-terminal binding site. In acid, binding occurs preferentially at the C-terminal site. 3. When simple iron donors (ferric and ferrous salts) are used the metal is distributed at random between the binding sites, as judged by the gel-electrophoresis method. The double-isotope method shows a preference of ferrous salts for the N-terminal site. 4. Quantitative treatment of the results of double-isotope labelling suggests that in the binding of iron to ovotransferrin at alkaline pH co-operative interactions between the sites occur. These interactions are apparently absent in the displacement of copper and in the binding of iron at acid pH.  相似文献   

9.
Iron limitation is one major constraint of microbial life, and a plethora of microbes use siderophores for high affinity iron acquisition. Because specific enzymes for reductive iron release in gram-positives are not known, we searched Firmicute genomes and found a novel association pattern of putative ferric siderophore reductases and uptake genes. The reductase from the schizokinen-producing alkaliphile Bacillus halodurans was found to cluster with a ferric citrate-hydroxamate uptake system and to catalyze iron release efficiently from Fe[III]-dicitrate, Fe[III]-schizokinen, Fe[III]-aerobactin, and ferrichrome. The gene was hence named fchR for ferric citrate and hydroxamate reductase. The tightly bound [2Fe-2S] cofactor of FchR was identified by UV-visible, EPR, CD spectroscopy, and mass spectrometry. Iron release kinetics were determined with several substrates by using ferredoxin as electron donor. Catalytic efficiencies were strongly enhanced in the presence of an iron-sulfur scaffold protein scavenging the released ferrous iron. Competitive inhibition of FchR was observed with Ga(III)-charged siderophores with K(i) values in the micromolar range. The principal catalytic mechanism was found to couple increasing K(m) and K(D) values of substrate binding with increasing k(cat) values, resulting in high catalytic efficiencies over a wide redox range. Physiologically, a chromosomal fchR deletion led to strongly impaired growth during iron limitation even in the presence of ferric siderophores. Inductively coupled plasma-MS analysis of ΔfchR revealed intracellular iron accumulation, indicating that the ferric substrates were not efficiently metabolized. We further show that FchR can be efficiently inhibited by redox-inert siderophore mimics in vivo, suggesting that substrate-specific ferric siderophore reductases may present future targets for microbial pathogen control.  相似文献   

10.
Release of iron from ferritin requires reduction of ferric to ferrous iron. The iron can participate in the diabetogenic action of alloxan. We investigated the ability of ascorbate to catalyze the release of iron from ferritin in the presence of alloxan. Incubation of ferritin with ascorbate alone elicited iron release (33 nmol/10 min) and the generation of ascorbate free radical, suggesting a direct role for ascorbate in iron reduction. Iron release by ascorbate significantly increased in the presence of alloxan, but alloxan alone was unable to release measurable amounts of iron from ferritin. Superoxide dismutase significantly inhibited ascorbate-mediated iron release in the presence of alloxan, whereas catalase did not. The amount of alloxan radical (A·) generated in reaction systems containing both ascorbate and alloxan decreased significantly upon addition of ferritin, suggesting that A· is directly involved in iron reduction. Although release of iron from ferritin and generation of A· were also observed in reactions containing GSH and alloxan, the amount of iron released in these reactions was not totally dependent on the amount of A· present, suggesting that other reductants in addition to A· (such as dialuric acid) may be involved in iron release mediated by GSH and alloxan. These results suggest that A· is the main reductant involved in ascorbate-mediated iron release from ferritin in the presence of alloxan and that both dialuric acid and A· contribute to GSH/alloxan-mediated iron release.  相似文献   

11.
Effect of zinc on translocation of iron in soybean plants   总被引:4,自引:3,他引:1       下载免费PDF全文
Zinc interfered with translocation of iron from roots to above ground parts of Glycine max. (L.) Merrill var. Hawkeye. During periods in which zinc impeded iron translocation, it also suppressed the production of reductant by roots. Addition of iron, as a ferric metal chelate (iron ethylenediaminedihydroxyphenylacetic acid), to the growth medium overcame the interference of zinc. In the root epidermis, potassium ferricyanide formed a precipitate (Prussian blue) with ferrous iron derived from the previously supplied iron ethylenediaminedihydroxyphenylacetic acid. The reduction of ferric iron was suppressed by zinc.  相似文献   

12.
Via the amides, α-amino acids can be converted to analogs of ethylenedinitrilotetraacetic acid, with retention of configuration at the asymmetric carbon. 1-(p-Carboxymethoxybenzyl)-ethylenedinitrilotetraacetic acid has been prepared from l-tyrosine and then coupled as the iron(III) chelate to 1,2-diaminoethane or to human serum albumin. The iron(III) chelate is effective in preventing EDTA carboxyl groups from reacting via carbodiimide coupling with amino groups. After the coupling reaction, iron is readily removed from the chelate upon reduction with ascorbate. These new procedures possess several attractive features for the use of “bifunctional” chelating agents in biophysical studies.  相似文献   

13.
Nine iron chelators were tested in five systems for their effects on radical-generation and conversion at chelator: iron molar ratios from 0.1 to 10. Stimulatory actions might distinguish toxic from safer chelators. Radical-generating reactions which represent different aspects of iron (ferrous and ferric) availability were studied: a) the reaction with hydrogen peroxide to hydroxylate benzoate; b) the oxidation of ascorbate; c) the reaction with hydrogen peroxide to fragment proteins; d) the reaction with hydrogen peroxide to permit amplified chemiluminescence; and e) the induction of peroxidation of mitochondrial membrane lipids. The compounds used were HBED, CP130, Desferal, EDTA, pyridine-hydrazone (CGP 43'902B), Ferrozine, CP 94 (CGP 46'700), LI (CGP 37 391) and rhodotorulic acid (CGP 45 274). Only the hexadentate compounds HBED, CP130 and Desferal were uniformly inhibitory (“protective”). The protective compounds were also apparently more stable during radical fluxes than the other chelators.  相似文献   

14.
Ascorbic acid improves endothelial barrier function by decreasing the permeability of endothelial cells cultured on semi-porous membrane filters. This decrease was not due to enhanced collagen synthesis and was mimicked by the collagen synthesis inhibitor ethyl-3,4-dihydroxybenzoic acid (EDHB). Since EDHB is known to chelate intracellular free iron, the effects of two membrane-permeant iron chelators were tested on endothelial permeability. Both 2,2′-dipyridyl and desferrioxamine decreased trans-endothelial permeability in a concentration-dependent manner. Increasing intracellular iron with a chelate of 8-hydroxyquinoline and ferric iron prevented effects of both EDHB and intracellular ascorbate. That EDHB and ascorbate did in fact chelate intracellular iron was supported by finding that they both decreased the cellular fluorescence quenching of the iron-sensitive dye Phen green SK. These results show that chelation of intracellular iron decreases endothelial barrier permeability and implicate this mechanism in the ability of EDHB and possibly intracellular ascorbate to tighten the endothelial barrier.  相似文献   

15.
《Free radical research》2013,47(6):349-353
Chelates can inhibit the iron- and copper-catalyzed autoxidation of ascorbate at pH 7.0. Diethylenetri-aminepentaacetic acid (DTPA or DETAPAC) and Desferal (deferoximane mesylate) slow the iron-catalyzed oxidation of ascorbate as effectively as reducing the trace levels of contaminating iron in buffers with Chelex resin. DETAPAC, EDTA and HEDTA (N-(2-hydroxyethyl)-ethylenediaminetriacetic acid) are effective at slowing the copper-catalyzed autoxidation of ascorbate while Desferal is ineffective. The ability to inhibit ascorbate autoxidation appears to parallel the rate of the reaction of superoxide with the iron chelate.  相似文献   

16.
Two free flavin-independent enzymes were purified by detecting the NAD(P)H oxidation in the presence of Fe(III)-EDTA and t-butyl hydroperoxide from E. coli. The enzyme that requires NADH or NADPH as an electron donor was a 28 kDa protein, and N-terminal sequencing revealed it to be oxygen-insensitive nitroreductase (NfnB). The second enzyme that requires NADPH as an electron donor was a 30 kDa protein, and N-terminal sequencing revealed it to be ferredoxin-NADP+ reductase (Fpr). The chemical stoichiometry of the Fenton activities of both NfnB and Fpr in the presence of Fe(III)-EDTA, NAD(P)H and hydrogen peroxide was investigated. Both enzymes showed a one-electron reduction in the reaction forming hydroxyl radical from hydrogen peroxide. Also, the observed Fenton activities of both enzymes in the presence of synthetic chelate iron compounds were higher than their activities in the presence of natural chelate iron compounds. When the Fenton reaction occurs, the ferric iron must be reduced to ferrous iron. The ferric reductase activities of both NfnB and Fpr occurred with synthetic chelate iron compounds. Unlike NfnB, Fpr also showed the ferric reductase activity on an iron storage protein, ferritin, and various natural iron chelate compounds including siderophore. The Fenton and ferric reductase reactions of both NfnB and Fpr occurred in the absence of free flavin. Although the k cat/K m value of NfnB for Fe(III)-EDTA was not affected by free flavin, the k cat/K m value of Fpr for Fe(III)-EDTA was 12-times greater in the presence of free FAD than in the absence of free FAD.  相似文献   

17.
Chelates can inhibit the iron- and copper-catalyzed autoxidation of ascorbate at pH 7.0. Diethylenetri-aminepentaacetic acid (DTPA or DETAPAC) and Desferal (deferoximane mesylate) slow the iron-catalyzed oxidation of ascorbate as effectively as reducing the trace levels of contaminating iron in buffers with Chelex resin. DETAPAC, EDTA and HEDTA (N-(2-hydroxyethyl)-ethylenediaminetriacetic acid) are effective at slowing the copper-catalyzed autoxidation of ascorbate while Desferal is ineffective. The ability to inhibit ascorbate autoxidation appears to parallel the rate of the reaction of superoxide with the iron chelate.  相似文献   

18.
O2-dependent reactions of the ferric and ferrous forms of alpha-hydroxyheme complexed with water-soluble rat heme oxygenase-1 were examined by rapid-scan stopped-flow measurements. Ferric alpha-hydroxyheme reacted with O2 to form ferric verdoheme with an O2-dependent rate constant of 4x10(5) M(-1) s(-1) at pH 7.4 and 9.0. A decrease of the rate constant to 2.8x10(5) M(-1) s(-1) at pH 6.5 indicates that the reaction proceeds by direct attack of O2 on the pi-neutral radical form of alpha-hydroxyheme, which is generated by deprotonation of the alpha-hydroxy group. The reaction of ferrous alpha-hydroxyheme with O2 yielded ferrous verdoheme in a biphasic fashion involving a new intermediate having absorption maxima at 415 and 815 nm. The rate constants for this two-step reaction were 68 and 145 s(-1). These results show that conversion of alpha-hydroxyheme to verdoheme is much faster than the reduction of coordinated iron (<1 s(-1)) under physiological conditions [Y. Liu, P.R. Ortiz de Montellano, Reaction intermediates and single turnover rate constants for the oxidation of heme by human heme oxygenase-1, J. Biol. Chem. 275 (2000) 5297-5307], suggesting that, in vivo, the conversion of ferric alpha-hydroxyheme to ferric verdoheme precedes the reduction of ferric alpha-hydroxyheme.  相似文献   

19.
Graminaceous plant species acquire soil iron by the release of phytosiderophores and subsequent uptake of iron(III)-phytosiderophore complexes. As plant species differ in their ability for phytosiderophore hydroxylation prior to release, an electrophoretic method was set up to determine whether hydroxylation affects the net charge of iron(III)-phytosiderophore complexes, and thus chelate stability. At pH 7.0, non-hydroxylated (deoxymugineic acid) and hydroxylated (mugineic acid; epi-hydroxymugineic acid) phytosiderophores form single negatively charged iron(III) complexes, in contrast to iron(III)-nicotianamine. As the degree of phytosiderophore hydroxylation increases, the corresponding iron(III) complex was found to be less readily protonated. Measured pKa values of the amino groups and calculated free iron(III) concentrations in presence of a 10-fold chelator excess were also found to decrease with increasing degree of hydroxylation, confirming that phytosiderophore hydroxylation protects against acid-induced protonation of the iron(III)-phytosiderophore complex. These effects are almost certainly associated with intramolecular hydrogen bonding between the hydroxyl and amino functions. We conclude that introduction of hydroxyl groups into the phytosiderophore skeleton increases iron(III)-chelate stability in acid environments such as those found in the rhizosphere or the root apoplasm and may contribute to an enhanced iron acquisition.  相似文献   

20.
Myeloperoxidase (MPO), which is involved in host defence and inflammation, is a unique peroxidase in having a globin-like standard reduction potential of the ferric/ferrous couple. Intravacuolar and exogenous MPO released from stimulated neutrophils has been shown to exist in the oxyferrous form, called compound III. To investigate the reactivity of ferrous MPO with molecular oxygen, a stopped-flow kinetic analysis was performed. In the absence of dioxygen, ferrous MPO decays to ferric MPO (0.04 s(-1) at pH 8 versus 1.4 s(-1) at pH 5). At pH 7.0 and 25 degrees C, compound III formation (i.e., binding of dioxygen to ferrous MPO) occurs with a rate constant of (1.1+/-0.1) x 10(4)M(-1)s(-1). The rate doubles at pH 5.0 and oxygen binding is reversible. At pH 7.0, the dissociation equilibrium constant of the oxyferrous form is (173+/-12)microM. The rate constant of dioxygen dissociation from compound III is much higher than conversion of compound III to ferric MPO (which is not affected by the oxygen concentration). This allows an efficient transition of compound III to redox intermediates which actually participate in the peroxidase or halogenation cycle of MPO.  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