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.
Interaction between iron(II) and acetohydroxamic acid (Aha), alpha-alaninehydroxamic acid (alpha-Alaha), beta-alaninehydroxamic acid (beta-Alaha), hexanedioic acid bis(3-hydroxycarbamoyl-methyl)amide (Dha) or desferrioxamine B (DFB) under anaerobic conditions was studied by pH-metric and UV-Visible spectrophotometric methods. The stability constants of complexes formed with Aha, alpha-Alaha, beta-Alaha and Dha were calculated and turned out to be much lower than those of the corresponding iron(II) complexes. Stability constants of the iron(II)-hydroxamate complexes are compared with those of other divalent 3d-block metal ions and the Irving-Williams series of stabilities was found to be observed. Above pH 4, in the reactions between iron(II) and desferrioxamine B, the oxidation of the metal ion to iron(III) by the ligand was found. The overall reaction that resulted in the formation of the tris-hydroxamato complex [Fe(HDFB)]+ and monoamide derivative of DFB at pH 6 is: 2Fe2+ + 3H4DFB+ = 2[Fe(HDFB)]+ + H3DFB-monoamide+ + H2O + 4H+. Based on these results, the conclusion is that desferrioxamine B can uptake iron in iron(III) form under anaerobic conditions.  相似文献   

2.
The yeast Rhodotorula pilimanae produces the dihydroxamate siderophore rhodotorulic acid (RA) in prodigious amounts when starved for iron. Synthetic dihydroxamate analogs of RA have been prepared in which the diketopiperazine ring of RA is replaced by a simple chain of n methylene groups. It is found that R. pilimanae is able to accumulate iron using these achiral complexes, as well as from simple monohydroxamate analogs, at rates comparable to those of RA. While the Fe2RA3 complex does not enter the cell, there is a receptor system whose geometric requirements for siderophore recognition have been probed using analogs. In contrast to mono- or dihydroxamate ligands, the trihydroxamate siderophores such as ferrioxamine B are completely ineffective at delivering iron to R. pilimanae. This is ascribed to the greater stability of these complexes, which blocks release of the Fe(III) in a ligand exchange process that is required for uptake. To explore whether this ligand exchange involves redox catalysis, Ga(III) was substituted for Fe(III). The gallium was taken up at rates near those of iron and were also energy-dependent, as determined by metabolic inhibition with KCN.  相似文献   

3.
Iron reduction and uptake was studied in wild-type and haem-deficient strains of Saccharomyces cerevisiae. Haem-deficient strains lacked inducible ferri-reductase activity and were unable to take up iron from different ferric chelates such as Fe(III)-citrate or rhodoturulic acid. In contrast, ferrioxamine B was taken up actively by the mutants as well as by the wild-type strains. At a low extracellular concentration, uptake was insensitive to ferrozine and competitively inhibited by Ga(III)-desferrioxamine B. Extracellular reductive dissociation of the siderophore occurred at higher extracellular concentrations. Two mechanisms appear to contribute to the uptake of ferrioxamine B by S. cerevisiae: one with high affinity, by which the siderophore is internalized as such and another with lower affinity by which iron is dissociated from the ligand prior to uptake.  相似文献   

4.
Cucumber, as a strategy I plant, and Maize as a strategy II plant, were cultivated in hydroponic culture in the presence of a ferrated siderophore mixture (1 M) from a culture of Penicillium chrysogenumisolated from soil. The siderophore mixture significantly improved the iron status of these plants as measured by chlorophyll concentration to the same degree as a 100-fold higher FeEDTA supply. Analysis of the siderophore mixture from P. chrysogenum by HPLC and electrospray mass spectrometry revealed that besides the trihydroxamates, coprogen and ferricrocin, large amounts of dimerum acid and fusarinines were present which represent precursor siderophores or breakdown products of coprogen. In order to prove the iron donor properties of dimerum acid and fusarinines for plants, purified coprogen was hydrolyzed with ammonia and the hydrolysis products consisting of dimerum acid and fusarinine were used for iron uptake by cucumber and maize. In short term experiments radioactive iron uptake and translocation rates were determined using ferrioxamine B, coprogen and hydrolysis products of coprogen. While the trihydroxamates revealed negligible or intermediate iron uptake rates by both plant species, the fungal siderophore mixture and the ammoniacal hydrolysis products of coprogen showed high iron uptake, suggesting that dimerum acid and fusarinines are very efficient iron sources for plants. Iron reduction assays using cucumber roots or ascorbic acid also showed that iron bound to hydrolysis products of coprogen was more easily reduced compared to iron bound to trihydroxamates. Ligand exchange studies with epi-hydroxymugineic acid and EDTA showed that iron was easily exchanged between coprogen hydrolysis products and phytosiderophores or EDTA. The results indicate that coprogen hydrolysis products are an excellent source for Fe nutrition of plants.  相似文献   

5.
The utilization of iron and its complexes by mammalian mitochondria   总被引:5,自引:2,他引:3       下载免费PDF全文
Sonicated mitochondria catalyse the reduction of ferric salts, and the subsequent incorporation of Fe(2+) into haem, when provided with a reducing substrate such as succinate or NADH. The rate of haem synthesis was low under aerobic conditions and, after a short lag period, accelerated once anaerobic conditions were achieved; it was insensitive to antimycin A. The lag period was decreased by preincubating the mitochondria with NADH and Fe(3+). Newly formed Fe(2+) was autoxidized rapidly and the consequent O(2) uptake was measured with an oxygen electrode to determine the rate of enzymic formation of Fe(2+) from FeCl(3); this reaction was rapid in sonicated mitochondria provided with NADH or succinate and was insensitive to antimycin A. The reaction was very slow in intact mitochondria, suggesting a permeability barrier to Fe(3+) ions. This system was used to test the permeability of the mitochondrial membrane to various iron complexes of biological importance. Of the compounds tested only ferrioxamine G appeared to penetrate readily and the iron of this complex was reduced when intact mitochondria were supplied with succinate or NADH-linked substrates. The reduction was insensitive to rotenone or antimycin A. Both ferrioxamine G and ferrioxamine B were, however, reduced by particles. The membrane fraction of sonicated mitochondria was necessary for the reduction. The rate of ferrioxamine B reduction by sonicated mitochondria was measured by a dual-wavelength spectrophotometric assay and was found to be stimulated in conditions where the Fe(2+) produced was utilized for haem synthesis. The addition of FeCl(3) to anaerobic particles caused an oxidation of cytochrome b when this region of the respiratory chain was isolated by treatment with rotenone and antimycin A. These results suggest that the reduction of ferric iron and its complexes occurs inside the inner mitochondrial membrane in proximity to ferrochelatase. Possible sites for this reduction are the flavoproteins, succinate and NADH dehydrogenase.  相似文献   

6.
FhuF is a cytoplasmic 2Fe-2S protein of Escherichia coli loosely associated with the cytoplasmic membrane. E. coli fhuF mutants showed reduced growth on plates with ferrioxamine B as the sole iron source, although siderophore uptake was not defective in transport experiments. Removal of iron from coprogen, ferrichrome, and ferrioxamine B was significantly lower in fhuF mutants compared to the corresponding parental strains, which suggested that FhuF is involved in iron removal from these hydroxamate-type siderophores. A redox potential E(1/2) of -310 +/- 25 mV relative to the normal hydrogen electrode was determined for FhuF by EPR redox titration; this redox potential is sufficient to reduce the siderophores coprogen and ferrichrome. M?ssbauer spectra revealed that FhuF in its [Fe(2+)-Fe(3+)] state is also capable of direct reduction of ferrioxamine B-bound ferric iron, thus proving its reductase function. This is the first report on a bacterial siderophore-iron reductase which in vivo seems to be specific for a certain group of hydroxamates.  相似文献   

7.
The mammalian ZIP (Zrt-, Irt-like Protein) family of transmembrane transport proteins consists of 14 members that share considerable homology. ZIP proteins have been shown to mediate the cellular uptake of the essential trace elements zinc, iron, and manganese. The aim of the present study was to determine the effect of dietary iron deficiency and overload on the expression of all 14 ZIP transporters in the liver, the main site of iron storage. Weanling male rats (n = 6/group) were fed iron-deficient (FeD), iron-adequate (FeA), or iron-overloaded (FeO) diets in two independent feeding studies. In study 1, diets were based on the TestDiet 5755 formulation and contained iron at 9 ppm (FeD), 215 ppm (FeA), and 27,974 ppm (3% FeO). In study 2, diets were based on the AIN-93G formulation and contained iron at 9 ppm Fe (FeD), 50 ppm Fe (FeA), or 18916 ppm (2% FeO). After 3 weeks, the FeD diets depleted liver non-heme iron stores and induced anemia, whereas FeO diets resulted in hepatic iron overload. Quantitative RT-PCR revealed that ZIP5 mRNA levels were 3- and 8-fold higher in 2% FeO and 3% FeO livers, respectively, compared with FeA controls. In both studies, a consistent downregulation of ZIP6, ZIP7, and ZIP10 was also observed in FeO liver relative to FeA controls. Studies in H4IIE hepatoma cells further documented that iron loading affects the expression of these ZIP transporters. Overall, our data suggest that ZIP5, ZIP6, ZIP7, and ZIP10 are regulated by iron, indicating that they may play a role in hepatic iron/metal homeostasis during iron deficiency and overload.  相似文献   

8.
Anabaena sp. strain 6411, which produces the dihydroxamate siderophore schizokinen to facilitate iron uptake, is also capable of using the related siderophore aerobactin. The two siderophores compete for the same iron transport system, but there is a markedly higher affinity for ferric schizokinen than for ferric aerobactin. The trihydroxamate siderophore ferrioxamine B is far less effective as an iron donor in this organism. Anabaena sp. strain 7120 appears to be closely related to strain 6411. It synthesizes schizokinen as its major siderophore and shows rates of iron uptake from ferric schizokinen, ferric aerobactin, and ferrioxamine B which are similar to those observed with strain 6411. Anabaena cylindrica Lemm. 7122 and 1611, on the other hand, differ from strain 6411. In contrast to schizokinen, the hydroxamate which they produce in response to iron starvation cannot be extracted with water from the organic layer and does not support the growth of the siderophore auxotroph Arthrobacter flavescens JG-9. Strain 7122 can use its endogenous siderophore or schizokinen to promote iron uptake, but at 50-fold-lower rates than are observed with Anabaena sp. strain 6411 or 7120.  相似文献   

9.
10.
Iron is a member of a small group of nutrients that limits aquatic primary production. Mechanisms for utilizing iron have to be efficient and adapted according to the ecological niche. In respect to iron acquisition cyanobacteria, prokaryotic oxygen evolving photosynthetic organisms can be divided into siderophore‐ and non‐siderophore‐producing strains. The results presented in this paper suggest that the situation is far more complex. To understand the bioavailability of different iron substrates and the advantages of various uptake strategies, we examined iron uptake mechanisms in the siderophore‐producing cyanobacterium Anabaena sp. PCC 7120. Comparison of the uptake of iron complexed with exogenous (desferrioxamine B, DFB) or to self‐secreted (schizokinen) siderophores by Anabaena sp. revealed that uptake of the endogenous produced siderophore complexed to iron is more efficient. In addition, Anabaena sp. is able to take up dissolved, ferric iron hydroxide species (Fe′) via a reductive mechanism. Thus, Anabaena sp. exhibits both, siderophore‐ and non‐siderophore‐mediated iron uptake. While assimilation of Fe′ and FeDFB are not induced by iron starvation, FeSchizokinen uptake rates increase with increasing iron starvation. Consequently, we suggest that Fe′ reduction and uptake is advantageous for low‐density cultures, while at higher densities siderophore uptake is preferred.  相似文献   

11.
Lindsay S  Brosnahan D  Watt GD 《Biochemistry》2001,40(11):3340-3347
The reaction of Fe2+ with O2 in the presence of horse spleen ferritin (HoSF) results in deposition of FeOH3 into the hollow interior of HoSF. This reaction was examined at low Fe2+/HoSF ratios (5-100) under saturating air at pH 6.5-8.0 to determine if H2O2 is a product of the iron deposition reaction. Three methods specific for H2O2 detection were used to assess H2O2 formation: (1) a fluorometric method with emission at 590 nm, (2) an optical absorbance method based on the reaction H2O2 + 3I- + 2H+ = I3- + 2H2O monitored at 340 nm for I3- formation, and (3) a differential pulsed electrochemical method that measures O2 and H2O2 concentrations simultaneously. Detection limits of 0.25, 2.5, and 5.0 microM H2O2 were determined for the three methods, respectively. Under constant air-saturation conditions (20% O2) and for a 5-100 Fe2+/HoSF ratio, Fe2+ was oxidized and the resulting Fe3+ was deposited within HoSF but no H2O2 was detected as predicted by the reaction 2Fe2+ + O2 + 6H2O = 2Fe(OH)3 + H2O2 + 4H+. Two other sets of conditions were also examined: one with excess but nonsaturating O2 and another with limiting O2. No H2O2 was detected in either case. The absence of H2O2 formation under these same conditions was confirmed by microcoulometric measurements. Taken together, the results show that under low iron loading conditions (5-100 Fe2+/HoSF ratio), H2O2 is not produced during iron deposition into HoSF using O2 as an oxidant. This conclusion is inconsistent with previous, carefully conducted stoichiometric and kinetic measurements [Xu, B., and Chasteen, N. D. (1991) J. Biol. Chem. 266, 19965], predicting that H2O2 is a quantitative product of the iron deposition reaction with O2 as an oxidant, even though it was not directly detected. Possible explanations for these conflicting results are considered.  相似文献   

12.
The ability of a siderophore-catabolizing bacterium to assimilate ferric ion was examined. While the bacterium utilizes the siderophore deferrioxamine B (DFB) as a carbon source, it was incapable of using the ferricion analogue (ferrioxamine B) as an iron source. It did, however, assimilate the ferric ion of the chelator ferric nitrilotriacetic acid and of the siderophore ferrirhodotorulic acid (ferriRA). Neither ferriRA nor its deferrated analog (RA), however, were capable of functioning as carbon sources for the bacterium. The microbe thus employs a nutritional selectivity with respect to these two siderophores. That is, it does not use the siderophore it employs as a carbon source (DFB) as an iron source nor does the siderophore utilized as an iron source, i.e. ferriRA, nor its deferrated analog (RA), serve as carbon sources for the organism.This paper is dedicated to the memory of Professor Thomas Emery. Professor Emery was instrumental in giving support and advice at a time when such mentorship greatly aided the corresponding author in developing a program concerning the catabolism of siderophores by microbes.  相似文献   

13.
The oxalate catalyzed iron(III) transfer from a trihydroxamate siderophore ferrioxamine B, [Fe(Hdfb)+], to ethylenediaminetetraacetic acid (H4edta) has been studied spectro-photometrically in weakly acidic aqueous solutions at 298 K and a constant 2.0 M ionic strength maintained by NaClO4. The results reveal that oxalate is a more efficient catalyst than the so far studied synthetic monohydroxamic acids. Any role of reduction of Fe(Hdfb)+ by oxalate in the catalysis has been rejected by the experimentally observed preservation of the oxalate concentration during the reaction time. Therefore, catalysis has been proposed to be a substitution based process. Under our experimental conditions Fe(Hdfb)+ is hexacoordinated and addition of oxalate results in the formation of Fe(H2dfb)(C2O4), Fe(H3dfb)(C2O4)2 and Fe(C2O4)3−3. Therefore, catalysis was proposed to be accomplished by the intermediate formation of the ternary and tris(oxalato) complexes. All three complexes react with H2edta2− to form thermodynamically stable Fe(edta) as a final reaction product. Whereas the formation of the ternary complexes is fast enough to feature a pre-equilibrium process to the iron exchange reaction, the formation of Fe(C2O4)3−3 is slow and is directly involved in the rate determining step of the Fe(edta) formation. Nonlinear dependencies of the rate constant on the oxalate and the proton concentrations have been observed and a four parallel path mechanism is proposed for the exchange reaction. The rate and equilibrium constants for the various reaction paths were determined from the kinetic and equilibrium study involving the desferrioxamine B- (H4dfb+), oxalate- and proton-concentration variations. The observed proton catalysis was attributed to the fast monoprotonation of ferrioxamine B as well as of the oxalate ligand. The observed catalysis of iron dissociation from the siderophore has been discussed in view of its significance with respect to in vivo microbial iron transport.  相似文献   

14.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

15.
The synthesis of two new synthetic analogues of lecithin, two of phosphatidyl ethanolamine ("cephalin"), and one new phosphatidic acid analogue is described. They comprise one of each of the following types: the "isosteric" diether lecithin and cephalin analogues ROCH(2)CH(OR)- CH(2)CH(2)P(O) (O(-))OCH(2)CH(2)N(+)R'(3) (R = C(18)H(37); R' = H or CH(3)); and the "hydrocarbon" analogues of phosphatidic acid, lecithin, and cephalin, C(17)H(35)CH(2)CH(C(18)H(37))CH(2)P(O)(R) = (R'); [R = R' = OH; R = O(-), R' = OCH(2)CH(2)N(+)(CH(3))(3); and R = O(-), R' = OCH(2)CH(2)N(+)H(3)]. Infrared spectra and other properties of these compounds are described.  相似文献   

16.
Intense lipid peroxidation of brain synaptosomes initiated with Fenton's reagent (H2O2 + Fe2+) began instantly upon addition of Fe2+ and preceded detectable OH. formation. Although mannitol or Tris partially blocked peroxidation, concentrations required were 10(3)-fold in excess of OH. actually formed, and inhibition by Tris was pH dependent. Lipid peroxidation also was initiated by either Fe2+ or Fe3+ alone, although significant lag phases (minutes) and slowed reaction rates were observed. Lag phases were dramatically reduced or nearly eliminated, and reaction rates were increased by a combination of Fe3+ and Fe2+. In this instance, lipid peroxidation initiated by optimal concentrations of H2O2 and Fe2+ could be mimicked or even surpassed by providing optimal ratios of Fe3+ to Fe2+. Peroxidation observed with Fe3+ alone was dependent upon trace amounts of contaminating Fe2+ in Fe3+ preparations. Optimal ratios of Fe3+:Fe2+ for the rapid initiation of lipid peroxidation were on order of 1:1 to 7:1. No OH. formation could be detected with this system. Although low concentrations of H2O2 or ascorbate increased lipid peroxidation by Fe2+ or Fe3+, respectively, high concentrations of H2O2 or ascorbate (in excess of iron) inhibited lipid peroxidation due to oxidative or reductive maintenance of iron exclusively in Fe2+ or Fe3+ form. Stimulation of lipid peroxidation by low concentrations of H2O2 or ascorbate was due to the oxidative or reductive creation of Fe3+:Fe2+ ratios. The data suggest that the absolute ratio of Fe3+ to Fe2+ was the primary determining factor for the initiation of lipid peroxidation reactions.  相似文献   

17.
18.
The purpose of this study was to investigate the possible mechanism by which endotoxin enhances peroxidative damage to membrane lipids. Male B6C3 mice were treated with endotoxin intraperitoneally 0 or 20 mg/kg body weight for 24 h. Freshly prepared liver homogenate was incubated with either 1-5 mM of reduced glutathione (GSH), glucose, H(2)O(2), ascorbic acid (AA), FeSO(4), FeCl(3), EDTA, FeCl(3) plus AA, AA plus EDTA or EDTA plus FeCl(3) in phosphate-buffered saline (PBS), pH 7.0, or PBS, at 37 degrees C for 60 min. The levels of lipid peroxidation products, thiobarbituric acid reactants (TBAR), were significantly higher in the liver of endotoxin-treated mice, and the values were markedly increased following incubation. Compared to PBS, incubation with H(2)O(2), FeCl(3), FeSO(4), and AA, but not glucose, significantly enhanced TBAR formation. The greatest increase of TBAR was found when AA and FeCl(3) were added together. On the other hand, EDTA and GSH inhibited the formation of TBAR during incubation. When added before AA, EDTA completely inhibited the peroxidative effect of AA or FeSO4, and when added subsequent to AA, EDTA partially prevented the adverse effect of AA. The results obtained suggest that ionic iron plays an important role in initiating endotoxin-induced peroxidative damage to membrane lipids, and that AA may be involved in releasing iron from its protein complex and/or maintaining ionic iron in a reduced or catalytic state.  相似文献   

19.
Henia Mor  Isaac Barash 《Biometals》1990,2(4):209-213
Summary Geotrichum candidum is capable of utilizing iron from hydroxamate siderophores of different structural classes. The relative rates of iron transport for ferrichrome, ferrichrysin, ferrioxamine B, fusigen, ferrichrome A, rhodotorulic acid, coprogen B, dimerium acid and ferrirhodin were 100%, 98%, 74%, 59%, 49%, 35%, 24%, 12% and 11% respectively. Ferrichrome, ferrichrysine and ferrichrome A inhibited [59Fe]ferrioxamine-B-mediated iron transport by 71%, 68% and 28% respectively when added at equimolar concentrations to the radioactive complex. The inhibitory mechanism of [59Fe]ferrioxamine B uptake by ferrichrome was non-competitive (K i 2.4 M), suggesting that the two siderophores do not share a common transport system. Uptake of [59Fe]ferrichrome, [59Fe]rhodotorulic acid and [59Fe]fusigen was unaffected by competition with the other two siderophores or with ferrioxamine B. Thus,G. candidum may possess independent transport systems for siderophores of different structural classes. The uptake rates of [14C]ferrioxamine B and67Ga-desferrioxamine B were 30% and 60% respectively, as compared to [59Fe]ferrioxamine B. The specific ferrous chelates, dipyridyl and ferrozine at 6 mM, caused 65% and 35% inhibition of [59Fe]ferrioxamine uptake. From these results we conclude that, although about 70% of the iron is apparently removed from the complex by reduction prior to being transported across the cellular membrane, a significant portion of the chelated ligand may enter the cell intact. The former and latter mechanisms seem not to be mutually exclusive.  相似文献   

20.
The extracellular enzymatic reduction of iron by microorganisms has not been appropriately considered. In this study the reduction and release of iron from ferrioxamine were examined using extracellular microbial iron reductases and compared to iron mobilization by chemical reductants, and to chelation by EDTA and desferrioxamine. A flavin semiquinone was formed during the enzymatic reduction of ferrioxamine, which was consistent with the 1 e(-) reduction of iron by an enzyme. The rates for the enzymatic reactions were substantially faster than both the 2 e(-) chemical reductions and the chelation reactions. The rapid rates of the enzymatic reduction reactions demonstrated that these enzymes are capable of accomplishing the extracellular mobilization of iron required by microorganisms. The data suggest that mechanistically there are two phases for the mobilization and transport of iron by those microorganisms that produce both extracellular iron reductases and siderophores, with reduction being the principle pathway.  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