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.
4-Hydroxy-6-methyl-3-pyridinecarboxylic acid (DQ6) and the new compound 2,6-dimethyl-4-hydroxy-3-pyridinecarboxylic acid (DQ726) were evaluated for possible application for iron (Fe) and aluminium (Al) chelation therapy. Metal/ligand solution chemistry, cytotoxicity, octanol/water partitioning (Do/w), and chelation efficiency were studied. The solution chemistry of the two ligands with Fe(III) and Al(III) was investigated in aqueous 0.6 m (Na)Cl at 25 °C by means of potentiometric titrations, UV-Vis spectrophotometry, and 1H NMR spectroscopy. DQ6 exhibited a high coordination efficiency towards Al(III). Fe(III)/DQ6, Al(III)/DQ726, and Fe(III)/DQ726 complexes were less stable. These results were confirmed by chelation efficiency measurements performed in an octanol/aqueous solution. Accordingly, the effects of the substitution at various ring positions of 4-hydroxy-3-pyridinecarboxylic acid were rationalised. Partitioning experiments at pH 7.4 showed both DQ6 and DQ726, and their Fe(III) and Al(III) complexes, to be hydrophilic. The toxicity of DQ6 and of DQ726 was investigated with human cancer cell lines and normal human primary cells: no cytotoxic effects were observed up to 0.1 mM, following a 3 days exposure. According to our results, DQ6 has the favourable properties to be a chelating agent for Al.  相似文献   

2.
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   

3.
With the aim to design new chelators for the clinical treatment of different diseases involving the trivalent metal ions Fe(III) and Al(III), we present the equilibria of kojic acid and its derivative 6-[5-hydroxy-2-hydroxymethyl-pyran-4-one]-5-hydroxy-2-hydroxymethyl-pyran-4-one with these two metal ions. Potentiometric and spectrophotometric techniques for iron, and potentiometry and 1H NMR for aluminum were used, supported by X-ray, electrospray ionization-mass spectrometry (ESI-MS), calorimetry and quantum chemical calculations. In this work, evidence is given on the formation of MeL, MeL2, and MeL3 complexes of both metal ions with kojic acid, confirmed by the X-ray structure of the FeL3 complex, and of variously protonated Me2L2 and MeL2 complexes of 6-[5-hydroxy-2-hydroxymethyl-pyran-4-one]-5-hydroxy-2-hydroxymethyl-pyran-4-one. The extremely good pFe value for this second ligand gives confidence to, and opens perspectives for, the search of new kojic acid derivatives.  相似文献   

4.
Enzymatic reduction of physiological Fe(III) complexes of the "labile iron pool" has not been studied so far. By use of spectrophotometric assays based on the oxidation of NAD(P)H and formation of [Fe(II) (1,10-phenanthroline)3]2+ as well as by utilizing electron paramagnetic resonance spectrometry, it was demonstrated that the NAD(P)H-dependent flavoenzyme lipoyl dehydrogenase (diaphorase, EC 1.8.1.4) effectively catalyzes the one-electron reduction of Fe(III) complexes of citrate, ATP, and ADP at the expense of the co-enzymes NAD(P)H. Deactivated or inhibited lipoyl dehydrogenase did not reduce the Fe(III) complexes. Likewise, in the absence of NAD(P)H or in the presence of NAD(P)+, Fe(III) reduction could not be detected. The fact that reduction also occurred in the absence of molecular oxygen as well as in the presence of superoxide dismutase proved that the Fe(III) reduction was directly linked to the enzymatic activity of lipoyl dehydrogenase and not mediated by O2. Kinetic studies revealed different affinities of lipoyl dehydrogenase for the reduction of the low molecular weight Fe(III) complexes in the relative order Fe(III)-citrate > Fe(III)-ATP > Fe(III)-ADP (half-maximal velocities at 346-485 microm). These Fe(III) complexes were enzymatically reduced also by other flavoenzymes, namely glutathione reductase (EC 1.6.4.2), cytochrome c reductase (EC 1.6.99.3), and cytochrome P450 reductase (EC 1.6.2.4) with somewhat lower efficacy. The present data suggest a (patho)physiological role for lipoyl dehydrogenase and other flavoenzymes in intracellular iron metabolism.  相似文献   

5.
The interaction of chelators and reducing agents is of particular importance in understanding iron-associated pathology since catalytic iron undergoes cyclic reduction and oxidation in vivo. Therefore, we treated plasmid DNA with free or chelated Fe(III) in the presence of biological reductants, and simultaneously measured the number of single strand breaks (SSBs) and oxidative base modification (8-hydroxy-2'-deoxyguanosine; 8-OHdG) by quantitative gel electrophoresis and HPLC with electrochemical detection, respectively. Production of SSBs and 8-OHdG was linearly correlated suggesting that these two different lesions share a common chemical mechanism. The levels of both lesions were enhanced when Fe(III) was chelated to citrate or nitrilotriacetic acid. Reducing agents showed different potency in inducing DNA damage catalyzed by chelated iron (L-ascorbate > L-cysteine > H2O2). Chelation increased SSB formation by approximately 8-fold and 8-OHdG production by approximately 4-fold. The ratio of SSB/8-OHdG catalyzed by chelated iron, which is twice as high as by unchelated iron, indicates that chelation affects iron-catalyzed oxidative DNA damage in a specific way favoring strand breakage over base modification. Since iron is mostly chelated in biological systems, the production of genomic and mitochondrial DNA damage, particularly strand breaks, in diseases involving iron overload is likely to be higher than previously predicted from studies using unchelated iron.  相似文献   

6.
Deferriferrioxamine B (H3DFB) is a linear trihydroxamic acid siderophore with molecular formula NH2(CH2)5[N(OH)C(O)(CH2)2C(O)NH(CH2)5]2N(OH)C(O)CH3 that forms a kinetically and thermodynamically stable complex with iron(III), ferrioxamine B. Under the conditions of our study (pH = 4.30, 25 degrees C), ferrioxamine B, Fe(HDFB)+, is hexacoordinated and the terminal amine group is protonated. Addition of simple hydroxamic acids, R1C(O)N(OH)R2 (R1 = CH3, R2 = H; R1 = C6H5, R2 = H; R1 = R2 = CH3), to an aqueous solution of ferrioxamine B at pH = 4.30, 25.0 degrees C, I = 2.0, results in the formation of ternary complexes Fe(H2DFB)A+ and Fe(H3DFB)A2+, and tris complexes FeA3, where A- represents the bidendate hydroxamate anion R1C(O)N(O)R2-. The addition of a molar excess of ethylenediaminetetraacetic acid (EDTA) to an aqueous solution of ferrioxamine B at pH 4.30 results in a slow exchange of iron(III) to eventually completely form Fe(EDTA)- and H4DFB+. The addition of a hydroxamic acid, HA, catalyzes the rate of this iron exchange reaction: (formula; see text) A four parallel path mechanism is proposed for reaction (1) in which catalysis occurs via transient formation of the ternary and tris complexes Fe(H2DFB) A+, Fe(H3DFB)A2+, and FeA3. Rate and equilibrium constants for the various reaction paths to products were obtained and the influence of hydroxamic acid structure on catalytic efficiency is discussed. The importance of a low energy pathway for iron dissociation from a siderophore complex in influencing microbial iron bio-availability is discussed. The system represented by reaction (1) is proposed as a possible model for in vivo catalyzed release of iron from its siderophore complex at the cell wall or interior, where EDTA represents the intracellular storage depot or membrane-bound carrier and HA represents a low molecular weight hydroxamate-based metabolite capable of catalyzing interligand iron exchange.  相似文献   

7.
Polynuclear iron complexes of Fe(III) and phosphate occur in seawater and soils and in cells where the iron core of ferritin, the iron storage protein, contains up to 4500 Fe atoms in a complex with an average composition of (FeO.OH)8FeO.OPO3H2. Although phosphate influences the size of the ferritin core and thus the availability of stored iron, little is known about the nature of the Fe(III)-phosphate interaction. In the present study, Fe-phosphate interactions were analyzed in stable complexes of Fe(III).ATP which, in the polynuclear iron form, had phosphate at interior sites. Such Fe(III).ATP complexes are important not only as models but also because they may play a role in intracellular iron transport and in iron toxicity; the complexes were studied by extended x-ray absorption fine structure, EPR, NMR spectroscopy, and measurement of proton release. Mononuclear iron complexes exhibiting a g' = 4.3 EPR signal were formed at Fe:ATP ratios less than or equal to 1:3, and polynuclear iron complexes (Fe greater than or equal to 250, EPR silent at g' = 4.3) were formed at an Fe:ATP ratio of 4:1. No NMR signals due to ATP were observed when Fe was in excess (Fe:ATP = 4:1). Extended x-ray absorption fine structure analysis of the polynuclear Fe(III).ATP complex was able to distinguish an Fe-P distance at 3.27 A in addition to the octahedral O at 1.95 A and 4-5 Fe atoms at 3.36 A. The Fe-O and Fe-Fe distances are the same as in ferritin, and the Fe-P distance is analogous to that in another metal-ATP complex. An observable Fe-P environment in such a large polynuclear iron cluster as the Fe(III).ATP (4:1) complex indicates that the phosphate is distributed throughout rather than merely on the surface, in contrast to earlier models of chelate-stabilized iron clusters. Complexes of Fe(III) and ATP similar to those described here may form in vivo either as normal components of intracellular iron metabolism or during iron excess where the consequent alteration of free nucleotide triphosphate pools could contribute to the observed toxicity of iron.  相似文献   

8.
The alpha-hydroxyacid 2-hydroxy-4-methylthiobutanoic acid (the so-called methionine hydroxy-analogue, MHA), largely used in animal nutrition as a source of methionine, forms stable metal chelates with divalent metals of formula [{CH(3)SCH(2) CH(2)CH(OH)COO}(2)M].nH(2)O. Protonation and iron(III) and copper(II) complex formation constants have been determined by potentiometry at 25 degrees C. Distribution diagrams show that no free Fe(3+) cations are present in solution at pH>2.5. ESI-MS (Electron-Spray Ionization Mass Spectrometry) investigations carried out both on iron and zinc complexes in solution have evidenced various species with different MHA/metal ratios. In vivo trials were carried out with rats. After receiving a zinc-deficient diet for 3 weeks, animals were fed the same diet added with zinc sulfate or zinc/MHA chelate; the zinc content of faeces was higher (+45%; P<0.05) in sulfate fed rats, whereas zinc retention was higher (+61%; P<0.05) in the Zn/MHA diet. Experiments in vitro with human intestinal Caco-2 cells indicated that the MHA/Fe chelate was taken up by the cells without any apparent toxic effect. The iron uptake was higher than that of iron nitrilotriacetate (Fe(3+)NTA), an effective chelate for delivering iron to milk diets. In conclusion, these data indicate that the use of MHA chelates could be a valuable tool to increase bioavailability of trace minerals and reduce the environmental impact of animal manure.  相似文献   

9.
With the aim to investigate the role of the polyuronic components in the accumulation of iron and phosphate at the soil-root interface, the interactions of Ca-polygalacturonates (PGAs) with Fe(III) and P ions and of Fe(III)-Ca-polygalacturonates (Fe-PGAs) with P ions were studied at pH 4.7. The role of citric, malic and pyruvic acids in the mobilization of Fe(III) and P, in the presence and absence of Ca(II) 2.5mM, was also investigated. The sorption kinetics evidenced that P diffuses freely through the calcium polysaccharidic matrix whereas Fe(III) accumulates as an hydroxypolymer. The sorption kinetics of P by the Fe-PGA indicated that the amount of P sorbed increases with increasing its initial concentration up to a constant value equal to 0.98micromol/3.87micromolmg(-1) of Fe(III)-polymer trapped. The FT-IR spectra of the P-Fe-PGA systems, show bands attributable to P-O(H) stretching vibrations. The study of systems with a constant initial P amount and varying Fe(III) amounts allowed to hypothesize that phosphate settles down inside holes formed by the carboxylate groups of galacturonic units. Citric and malic acids showed to be active in the mobilization of both Fe and P whereas pyruvic acid appeared inactive.  相似文献   

10.
Although microbial activity and associated iron (oxy)hydroxides are known in general to affect the environmental dynamics of 4-hydroxy-3-nitrobenzenearsonic acid (roxarsone), the mechanistic understanding of the underlying biophysico-chemical processes remains unclear due to limited experimental information. We studied how Shewanella oneidensis MR-1 –a widely distributed metal-reducing bacterium, in the presence of dissolved Fe(III), affects roxarsone transformations and biogeochemical cycling in a model aqueous system. The results showed that the MR-1 strain was able to anaerobically use roxarsone as a terminal electron acceptor and to convert it to a single product, 3-amino-4-hydroxybenzene arsonic acid (AHBAA). The presence of Fe(III) stimulated roxarsone transformation via MR-1-induced Fe(III) reduction, whereby the resulting Fe(II) acted as an efficient reductant for roxarsone transformation. In addition, the subsequent secondary Fe(III)/Fe(II) mineralization created conditions for adsorption of organoarsenic compounds to the yielded precipitates and thereby led to arsenic immobilization. The study provided direct evidence of Shewanella oneidensis MR-1-induced direct and Fe(II)-associated roxarsone transformation. Quantitative estimations revealed a candidate mechanism for the early-stage environmental dynamics of roxarsone in nature, which is essential for understanding the environmental dynamics of roxarsone and successful risk assessment.  相似文献   

11.
Potentiometric, conductometric and 31P NMR titrations have been applied to study interactions between myo-inositol hexakisphosphate (phytic acid), (±)-myo-inositol 1,2,3,5-tetrakisphosphate and (±)-myo-inositol 1,2,3-trisphosphate with iron(III) ions. Potentiometric and conductometric titrations of myo-inositol phosphates show that addition of iron increases acidity and consumption of hydroxide titrant. By increasing the Fe(III)/InsP(6) ratio (from 0.5 to 4) 3 mol of protons are released per 2 mol of iron(III). At first, phytates coordinate iron octahedrally between P2 and P1,3. The second coordination site represents P5 and neighbouring P4,6 phosphate groups. Complexation is accompanied with the deprotonation of P1,3 and P4,6 phosphate oxygens. At higher concentration of iron(III) intermolecular P-O-Fe-O-P bonds trigger formation of a polymeric network and precipitation of the amorphous Fe(III)-InsP(6) aggregates. (31)P NMR titration data complement the above results and display the largest chemical shift changes at pD values between 5 and 10 in agreement with strong interactions between iron and myo-inositol phosphates. The differences in T(1) relaxation times of phosphorous atoms have shown that phosphate groups at positions 1, 2 and 3 are complexated with iron(III). The interactions between iron(III) ions and inositol phosphates depend significantly on the metal to ligand ratio and an attempt to coordinate more than two irons per InsP(6) molecule results in an unstable heterogeneous system.  相似文献   

12.
1. Micrococcus denitrificans excretes three catechol-containing compounds, which can bind iron, when grown aerobically and anaerobically in media deficient in iron, and anaerobically in medium with a high concentration of Ca2+. 2. One of these compounds was identified as 2,3-dihydroxybenzoic acid (compound I), and the other two were tentatively identified as N1N8-bis-(2,3-dihydroxybenzoyl)spermidine (compound II) and 2-hydroxybenzoyl-N-L-threonyl-N4[N1N8-bis-(2,3-dihydroxybenzoyl)]spermidine (compound III). 3. The equimolar ferric complex of compound III was prepared; compound III also forms complexes with Al3+, Cr3+ and Co2+ ions. 4. Cell-free extracts from iron-deficient organisms catalyse the formation of compound II from 2,3-dihydroxybenzoic acid and spermidine, and of compound III from compound II, L-threonine and 2-hydroxybenzoic acid; both reactions require ATP and dithiothreitol, and Mg2+ stimulates activity. The enzyme system catalysing the formation of compound II has optimum activity at pH 8.8 Fe2+ (35muM), Fe3+ (35muM) and Al3+ (65muM) inhibit the reaction by 50 percent. The enzyme system forming compound III has optimum activity at pH 8.6. Fe2+ (110 muM), Fe3+ (110 muM) and Al3+ (135 muM) inhibit the reaction by 50 percent. 5. At least two proteins are required for the formation of compound II, and another two proteins for its conversion into compound III. 6. The changes in the activities of these two systems were followed after cultures became deficient in iron. 7. Ferrous 1,10-phenanthroline is formed when a cell-free extract from iron-deficient cells is incubated with the ferric complex of compound III, succinate, NADH and 1,10-phenanthroline under N2.  相似文献   

13.
High-performance liquid chromatography (HPLC) with UV absorption detection was employed to measure the amounts of 8-hydroxy-2'-deoxyguanosine (8-OH-dG) produced from the nucleoside 2'-deoxyguanosine (dG) under varying reaction conditions using iron and H(2)O(2). The results indicate that 8-OH-dG produced from the reaction of iron and H(2)O(2) with dG can undergo reaction with free (i.e., unchelated) Fe(III) and that adding the chelating agent ethylenediaminetetraacetic acid (EDTA) after the reaction prevents this from occurring. It also appears that the free radical species generated by iron-EDTA chelates in pH 7.4 N-(2-hydroxyethyl)piperazine-N'-(2-ethanesulfonic acid) (Hepes) buffer is either not formed or unstable in unbuffered aqueous solution. Finally, 8-OH-dG levels are significantly larger when Fe(II) is allowed to bind to the nucleoside dG prior to addition of H(2)O(2). However, production of 8-OH-dG from unbound Fe(II) is also relevant. The results of this work show that differing reaction conditions in vivo, especially at the cellular level, will affect significantly the measured yields of 8-OH-dG. These results also have implications for studies involving DNA and the ability to distinguish between 8-OH-dG produced from free iron and iron bound to both phosphate groups and the DNA base guanine.  相似文献   

14.
Novel proline 3-hydroxylase (P3H) and proline 4-hydroxylase (P4H) activities that convert free l-proline to both trans 3- and trans 4-hydroxy- l-proline were detected in protein extracts of the anamorphic fungus Glarea lozoyensis. The enzymatic conversion of l-proline to trans 3- and trans 4-hydroxy- l-proline was strictly dependent on alpha-ketoglutarate, ascorbate, and dithiothreitol. Ferrous iron was required for optimal P3H and P4H activity. These substrate and co-factor requirements indicate these enzyme activities belong to the class of 2-oxoglutarate-dependent dioxygenases. Both P3H and P4H were inhibited by zinc and other trace metals. The addition of proline to the fermentation medium resulted in an increase in the specific activity of P4H and a decrease in the synthesis of pneumocandin C(0). Additionally, the synthesis of trans 3- and trans 4-hydroxy- l-proline in vivo was affected differently by the proline concentration in the medium. This result suggested that two enzymes may be responsible for the regio- and stereospecific hydroxylation of l-proline.  相似文献   

15.
Ferritin molecules contain 24 polypeptide chains folded as four-helix bundles and arranged as a hollow shell capable of storing up to 4500 Fe(III) atoms. H chains contain ferroxidase centres which lie within the bundle, about 12?Å (1.2?nm) from the outside surface and 8?Å from the inner surface of the protein shell. Catalysis of Fe(II) oxidation precedes storage of Fe(III) as ferrihydrite, with the formation of μ-oxo-bridged Fe(III) dimers as intermediates. Factors influencing the movement of μ-oxo-bridged Fe(III) from the ferroxidase centre to the ferritin cavity are uncertain. Assistance by small chelators is one possibility. The aim of this investigation was to determine whether iron at the dinuclear centres of three ferritins (human H chain homopolymer, HuHF, the non-haem ferritin of Escherichia coli, EcFTN, and horse spleen ferritin, HoSF) is accessible to chelators. Forty-eight Fe(II) atoms/molecule were added to the apoferritins followed, 2?min later, by the addition of chelator (1,10-phenanthroline, 2,2-bipyridine, desferrioxamine or 3,4-dihydroxybenzaldehyde). Iron species were analysed by Mössbauer spectroscopy or visible absorbance. Competition between chelators and apoferritin for Fe(II) was also investigated. The main conclusions of the study are that: (1) dinuclear iron and iron in small iron-cores in HuHF and EcFTN is mobilisable by all four chelators; (2) the chelators penetrate the shell; (3) 3,4-dihydroxybenzaldehyde is the most efficient in mobilising Fe(III) but the least successful in competing for Fe(II); (4) Fe(III) is more readily released from EcFTN than from HuHF; (5) 2,2′-bipyridine aids the movement of Fe(III) from ferroxidase centre to core.  相似文献   

16.
FeCl3 or Fe(III) that attached to chelating ligands such as citrate or nitrilotriacetic acid (NTA) at a molar ratio of 1:1 had a toxic effect on PMN. Uptake of radiolabeled Staphylococcus aureus by PMN, preincubated for 2 hr at 37 degrees C in a medium containing Fe(III)-citrate or Fe(III)-NTA, was significantly lower than that of control PMN preincubated without excess iron (p less than 0.002). However, at a 1:2 molar ratio of Fe(III) to citrate or NTA, the iron was not toxic. In contrast, the iron-liganding molecules transferrin and deferoxamine protected the PMN against the noxious effect of iron at concentrations just high enough to sequester all the iron. Fe(III) increased the generation of luminol chemiluminescence by stimulated PMN, whereas the oxygen consumption of the cells was not altered in the presence of Fe(III); this suggests a catalytic effect of iron on the production by PMN of oxygen metabolites at some step beyond the formation of superoxide. No effect of iron was observed when the incubation was performed at 4 degrees C, nor when an oxygen-radical scavenger such as thiourea, mannitol, or catalase was present in the incubation medium. Also, Fe(III) had much less effect on the phagocytic function of PMN of a patient with chronic granulomatous disease. The results indicate that the Fe(III)-induced defect in the phagocytic capacity of PMN depends on the nature and the concentration of the ligand attached to the iron ion, and also suggest that the noxious effect of iron on the PMN function is a result of its ability to catalyze the generation of toxic oxygen species by these cells.  相似文献   

17.
Potentiometric, conductometric and 31P NMR titrations have been applied to study interactions between myo-inositol hexakisphosphate (phytic acid), (±)-myo-inositol 1,2,3,5-tetrakisphosphate and (±)-myo-inositol 1,2,3-trisphosphate with iron(III) ions. Potentiometric and conductometric titrations of myo-inositol phosphates show that addition of iron increases acidity and consumption of hydroxide titrant. By increasing the Fe(III)/InsP6 ratio (from 0.5 to 4) 3 mol of protons are released per 2 mol of iron(III). At first, phytates coordinate iron octahedrally between P2 and P1,3. The second coordination site represents P5 and neighbouring P4,6 phosphate groups. Complexation is accompanied with the deprotonation of P1,3 and P4,6 phosphate oxygens. At higher concentration of iron(III) intermolecular P–O–Fe–O–P bonds trigger formation of a polymeric network and precipitation of the amorphous Fe(III)–InsP6 aggregates. 31P NMR titration data complement the above results and display the largest chemical shift changes at pD values between 5 and 10 in agreement with strong interactions between iron and myo-inositol phosphates. The differences in T1 relaxation times of phosphorous atoms have shown that phosphate groups at positions 1, 2 and 3 are complexated with iron(III). The interactions between iron(III) ions and inositol phosphates depend significantly on the metal to ligand ratio and an attempt to coordinate more than two irons per InsP6 molecule results in an unstable heterogeneous system.  相似文献   

18.
The effects of some naturally occurring iron ion chelators and their derivatives on the electron transfer from ferrous ions to oxygen molecules were examined by measuring oxygen consumption rates. Of the compounds examined, quinolinic acid, fusaric acid, and 2-pyridinecarboxylic acid repressed the oxygen consumption, whereas chlorogenic acid, caffeic acid, gallic acid, catechol l-β-(3,4-dihydroxyphenyl) alanine, and xanthurenic acid accelerated it. Theoretical calculations showed that the energies of the highest occupied molecular orbitals (HOMOs) of [Fe(II)(ligand)3] complexes were relatively high when the ligands were caffeic acid and its derivatives such as catechol, gallic acid, and l-β-(3,4-dihydroxyphenyl) alanine. On the other hand, the energies of the HOMOs of [Fe(II)(ligand)3] complexes were relatively low when the ligands were quinolinic acid and its derivatives such as 2-pyridinecarboxylic acid and fusaric acid. The energies of the HOMOs appear to be closely related with acceleration or repression of the oxygen consumption; that is to say, when the energy of the HOMO is high, the oxygen consumption is accelerated, and vice versa.  相似文献   

19.
Ferritin stores iron within a hollow protein shell as a polynuclear Fe(III) hydrous oxide core. Although iron uptake into ferritin has been studied previously, the early stages in the creation of the core need to be clarified. These are dealt with in this paper by using M?ssbauer spectroscopy, a technique that enables several types of Fe(II) and Fe(III) to be distinguished. Systematic M?ssbauer studies were performed on samples prepared by adding 57Fe(II) atoms to apoferritin as a function of pH (5.6-7.0), n [the number of Fe/molecule (4-480)], and tf (the time the samples were held at room temperature before freezing). The measurements made at 4.1 and 90 K showed that for samples with n less than or equal to 40 at pH greater than or equal to 6.25 all iron was trivalent at tf = 3 min. Four different Fe(III) species were identified: solitary Fe(III) atoms giving relaxation spectra, which can be identified with the species observed before by EPR and UV difference spectroscopy; oxo-bridged dimers giving doublet spectra with large splitting, observed for the first time in ferritin; small Fe(III) clusters giving doublets of smaller splitting and larger antiferromagnetically coupled Fe(III) clusters, similar to those found previously in larger ferritin iron cores, which, for samples with n greater than or equal to 40, gave magnetically split spectra at 4.1 K. Both solitary Fe(III) and dimers diminished with time, suggesting that they are intermediates in the formation of the iron core. Two kinds of divalent iron were distinguished for n = 480, which may correspond to bound and free Fe(II).  相似文献   

20.
The opportunistic pathogen Pseudomonas aeruginosa forms biofilms, which render it more resistant to antimicrobial agents. Levels of iron in excess of what is required for planktonic growth have been shown to promote biofilm formation, and therapies that interfere with ferric iron [Fe(III)] uptake combined with antibiotics may help treat P. aeruginosa infections. However, use of these therapies presumes that iron is in the Fe(III) state in the context of infection. Here we report the ability of phenazine-1-carboxylic acid (PCA), a common phenazine made by all phenazine-producing pseudomonads, to help P. aeruginosa alleviate Fe(III) limitation by reducing Fe(III) to ferrous iron [Fe(II)]. In the presence of PCA, a P. aeruginosa mutant lacking the ability to produce the siderophores pyoverdine and pyochelin can still develop into a biofilm. As has been previously reported (P. K. Singh, M. R. Parsek, E. P. Greenberg, and M. J. Welsh, Nature 417:552-555, 2002), biofilm formation by the wild type is blocked by subinhibitory concentrations of the Fe(III)-binding innate-immunity protein conalbumin, but here we show that this blockage can be rescued by PCA. FeoB, an Fe(II) uptake protein, is required for PCA to enable this rescue. Unlike PCA, the phenazine pyocyanin (PYO) can facilitate biofilm formation via an iron-independent pathway. While siderophore-mediated Fe(III) uptake is undoubtedly important at early stages of infection, these results suggest that at later stages of infection, PCA present in infected tissues may shift the redox equilibrium between Fe(III) and Fe(II), thereby making iron more bioavailable.  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