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.
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   

2.
Streptomyces pilosus is one of several microbes which produce ferrioxamine siderophores. In the accompanying paper (G. Müller and K. Raymond, J. Bacteriol. 160:304-312), the mechanism of iron uptake mediated by the endogenous ferrioxamines B, D1, D2, and E was examined. Here we report iron transport behavior in S. pilosus as mediated by the exogenous siderophores ferrichrome, ferrichrysin, rhodotorulic acid (RA), and synthetic enantio-RA. In each case iron acquisition depended on metabolic energy and had uptake rates comparable to that of [55Fe]ferrioxamine B. However, the synthetic ferric enantio-RA (which has the same preferred chirality at the metal center as ferrichrome) was twice as effective in supplying iron as was the natural ferric RA complex, suggesting that stereospecific recognition at the metal center is involved in the transport process. Iron uptake mediated by ferrichrome and ferric enantio-RA was strongly inhibited by kinetically inert chromic complexes of desferrioxamine B. These inhibition experiments indicate that iron from these exogenous siderophores is transported by the same uptake system as ferrioxamine B. Since the ligands have no structural similarity to ferrioxamine B except for the presence of three hydoxamate groups, we conclude that only the hydroxamate iron center and its direct surroundings are important for recognition and uptake. This hypothesis is supported by the fact that ferrichrome A and ferrirubin, which are both substituted at the hydroxamate carbonyl groups, were not (or were poorly) effective in supplying iron to S. pilosus.  相似文献   

3.
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   

4.
Summary Iron deprivation ofErwinia herbicola (Enterobacter agglomerans) induces the biosynthesis of six high-M r outer-membrane proteins and large amounts of ferrioxamine E. Mutagenesis withN-methyl-N-nitro-N-nitrosoguanidine and selection with ferrimycin A yielded mutants ofE. herbicola K4 (wild type), defective in the expression of a 76-kDa outer-membrane protein, as determined by SDS/polyacrylamide gel electrophoresis. While in bioassays wild-type cells showed growth promotion in the presence of ferrioxamines (B, D1, D2, E, G), enterobactin, citrate, ferrichrome and coprogen, these mutants failed to respond to ferrioxamines. Moreover, experiments with55Fe-labelled siderophores confirmed that iron transport mediated by ferrioxamine E and B in the mutants was completely inhibited, whereas iron transport by other hydroxamate siderophores, such as ferrichrome and coprogen was unaffected. The results are evidence that the 76-kDa protein in the outer membrane represents the receptor protein (FoxA) for ferrioxamines inE. herbicola.  相似文献   

5.
《Phytochemistry》1987,26(5):1317-1320
Under iron-deficient conditions Stemphylium botryosum f. sp. lycopersici produces three major siderophores; dimerum acid, coprogen B and an unidentified monohydroxamate siderophore designated as A. The system of siderophores mediating uptake of iron was characterized. It exhibits active transport, saturation kinetics and an optimum at pH 6 and 30°. The rate of iron uptake via dimerum acid and coprogen B was four times higher than siderophore A. S. botryosum was capable of taking up iron from hydroxamate siderophores produced by other fungi, e.g. ferrichrome, fusigen, rhodotorulic acid but not ferrioxamine B. Double labelling experiments suggest that ferric coprogen B accumulates in mycelial cells as an intact chelate.  相似文献   

6.
Summary Under conditions of iron-deprivationHafnia alvei (Enterobacteriaceae) produces ferrioxamine G as the principal siderophore. Maximum hydroxamate siderophore production occurred at medium iron limitation. The ferrioxamines were extracted, purified by gel filtration and chromatography on silica gel yielding a major and a minor siderophore fraction. The minor siderophore fraction contained three siderophores, among which ferrioxamine E could be identified by HPLC and FAB mass spectrometry. Reductive hydrolysis of the ferrioxamine G fraction yielded succinic acid and a mixture of diaminopentane and diaminobutane, as determined by gas-liquid chromatography and GLC/MS. HPLC and FAB mass spectrometry confirmed that the ferrioxamine G fraction consisted of two different species, G1 and G2, possessing molecular masses of 671 Da and 658 Da respectively.  相似文献   

7.
The siderophore production of various isolates of Phialocephala fortinii was assessed quantitatively as well as qualitatively in batch assays under pure culture conditions at different pH values and iron(III) concentrations. We found a distinct effect of both of these parameters on siderophore synthesis and as well as on fungal growth. In comparative analyses of two of the isolates, maximum siderophore production was found at a pH in the range of pH 4.0 to 4.5 while, under the experimental conditions employed, the optimal concentration of ferric iron was determined to be between 20–40 g iron (III) l–1 (0.36–0.72 M, respectively). HPLC analysis of the culture filtrate of most of the isolates of P. fortinii revealed the excretion of ferricrocin as main hydroxamate siderophore, followed by ferrirubin and ferrichrome C. The pattern of release of these three substances proved to be dependent on pH and iron(III) concentration of the culture medium, and to be specific for each isolate under investigation.  相似文献   

8.
After uptake of microbial ferrisiderophores, iron is assumed to be released by reduction. Two ferrisiderophore-reductase activities were identified in Escherichia coli K-12. They differed in cellular location, susceptibility to amytal, and competition between oxygen and ferrichrome-iron(III) reduction. The ferrisiderophore reductase associated with the 40,000×g sediment (membrane-bound enzyme) was inhibited by 10 mM amytal in contrast to the ferrisiderophore reductase present in the 100,000×g supernatant (soluble enzyme). Reduction by the membrane-bound enzyme followed sigmoid kinetics, but was biphasic in the case of the soluble enzyme. The soluble reductase could be assigned to a protein consisting of a single polypeptide of M r 26000. Reduction of iron(III) by the purified enzyme depended on the addition of NADH or NADPH which were equally active reductants. The cofactor FMN and to a lesser degree FAD stimulated the reaction. Substrate specificity of the soluble reductase was low. In addition to the hydroxamate siderophores arthrobactin, schizokinen, fusigen, aerobactin, ferrichrome, ferrioxamine B, coprogen, and ferrichrome A, the iron(III) complexes of synthetic catecholates, dihydroxy benzoic acid, and dicitrate, as well as carrier-free iron(III) were accepted as substrates. Both ferrisiderophore reductases were not controlled by the fur regulatory system and were not suppressed by anaerobic growth.Abbreviations DHB dihydroxybenzoic acid - MECAM 1,3,5-N,N,N-tris-(2,3-dihydroxybenzoyl)-triamino-methylbenzene - MECAMS 2,3-dihydroxy-5-sulfonyl-derivative of MECAM  相似文献   

9.
Summary Several strains of the enterobacterial groupErwinia herbicola (Enterobacter agglomerans) were screened for siderophore production. After 3 days of growth in a low-iron medium, all strains studied produced hydroxamate siderophores. The retention values of the main siderophore during thin-layer chromatography on silica gel plates and on HPLC reversed-phase columns were identical with those of an authentic sample of ferrioxamine E (norcardamine). Gas-chromatographic analysis of the HI hydrolyzate yielded succinic acid and 1,5-diaminopentane in equimolar amounts; fast-atom-bombardment (FAB) mass spectroscopy showed a molecular mass of 653 Da. Iron from55Fe-labelled ferrioxamine E was well taken up by iron-starved cells ofE. herbicola (K m=0.1 M,V max=8 pmol mg–1 min–1). However, besides ferrioxamine E (100%), several exogenous siderophores such as enterobactin (94.5%), ferric citrate (78.5%), coprogen (63.5%) and ferrichrome (17.5%) served as siderophores, suggesting the presence of multiple siderophore receptors in the outer membrane ofE. herbicola.  相似文献   

10.
While the naturally occurring reducing agents glutathione (GSH) and ascorbate (H2A) alone are ineffective at reducing iron(III) sequestered by the siderophore ferrioxamine B, the addition of an iron(II) chelator, sulfonated bathophenanthroline (BPDS), facilitates reduction by either reducing agent. A mechanism is described in which a ternary complex is formed between ferrioxamine B and BPDS in a rapidly established pre-equilibrium step, which is followed by rate limiting reduction of the ternary complex by glutathione or ascorbate. Spectral, thermodynamic, and kinetic evidence are given for ternary complex formation. Ascorbate was found to be slightly more efficient at reducing the ternary complex than glutathione (k4=2.1×10−3 M−1 s−1 and k4=6.3×10−4 M−1 s−1, respectively) at pH 7. Reduction is followed by a rapid ligand exchange step where iron is released from ferrioxamine B to form tris-(BPDS)iron(II). The implications of these results for siderophore mediated iron transport and release are discussed.  相似文献   

11.
The iron overload drug desferal (desferrioxamine B) forms the stable iron complex ferrioxamine B. The reduction potential of ferrioxamine B (Eo = −482 mV versus NHE pH 7) prohibits its reduction by biological reducing agents such as ascorbate, but it was found that the iron(II) chelator 2,2′-bipyridine (bipy) facilitates this reduction. Evidence is given to support the formation of a ternary complex between iron, bipy, and desferrioxamine B as the key step in facilitating the reduction. The equilibrium constant for the formation of the ternary complex was found to be 8.9 × 107 and ternary complex formation is explained in terms of a three step mechanism. The mechanism for the reduction of ferrioxamine B is discussed in terms of rapidly established pre-equilibria which include ternary complex formation, ascorbic acid deprotonation, and encounter complex formation between ascorbate and the ternary complex. These equilibria are followed by rate limiting reduction of the ternary complex. Bipy was found to be a similar facilitator to sulfonated bathophenanthroline for the reduction of ferrioxamine B by ascorbate.  相似文献   

12.
Desferrioxamine B is widely used as therapeutic agent for removal of excess body iron and, more recently, for removal of aluminium. A HPLC-based method for direct sensitive and reliable determination of ferrioxamine, desferrioxamine, aluminoxamine and related metabolites has been developed for use in pharmacokinetic studies. The method consists of complete separation of the analytes by an optimized mobile phase avoiding conversion of desferrioxamine to ferrioxamine by the analytical system and overcoming problems due to peak tailing properties of desferrioxamine. A post-column derivatization reaction with colourless fluoro-complexed iron converts all iron free species to ferrioxamine and allows quantification at 430 nm avoiding interference of UV-absorbing coelutes. This reaction might be of interest for other analytical procedures concerning iron chelators. The influence of the post-column reaction system on the column plate number is characterized. As the reaction rate of desferrioxamine and aluminoxamine with iron(III) is of second-order kinetics, a quadratic calibration function is observed, resulting from a compromise between residence time and peak broadening. A solid-phase extraction procedure is presented for extraction of the analytes from plasma. Limits of detection (S/N ratio of 3) were determined as 1.95 ng for ferrioxamine, 3.9 ng for aluminoxamine and 15.7 ng for desferrioxamine, expressed as on-column load. A new iron-free metabolite was identified with fast atom bombardment-mass spectrometry as N-hydroxylated desferrioxamine.  相似文献   

13.
The trihydroxamate siderophore, hydroxamate K, has been purified from culture filtrates of iron-deficient Rhizobium leguminosarum biovar viciae MNF710. The iron complex has a molecular weight of 828 and an absorption maximum at 443 nm (M=1510). 55Fe complexed to purified hydroxamate K was taken up by MNF710, its hydroxamate-negative mutant MNF7102 and Rhizobium leguminosarum biovar trifolii WU95 via an iron-regulated transport system, but Rhizobium meliloti U45 failed to take up the iron-siderophore complex under any conditions. A similar pattern of iron uptake was observed with ferrioxamine B. MNF710, MNF7102, U45 and WU95 all transported 55Fe-ferrichrome but only the first three strains took up 55Fe-ferrichrome A. All these 55Fe-trihydroxamate uptake systems were ironregulated in MNF710, MNF7102 and WU95. In contrast, uptake of 55Fe-rhodotorulate, a dihydroxamate, was essentially constitutive in all four organisms. Similarly, uptake of 55Fe-citrate and 55Fe-nitrilotriacetic acid was constitutive. None of the strains took up 55Fe complexed with enterobactin or with pyoverdins from Pseudomonas aeruginosa ATCC15692 (PAO1) and Pseudomonas fluorescens ATCC17400.  相似文献   

14.
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   

15.
The siderophore production of the facultative anaerobe Pseudomonas stutzeri, strain CCUG 36651, grown under both aerobic and anaerobic conditions, was investigated by liquid chromatography and mass spectrometry. The bacterial strain has been isolated at a 626-m depth at the Äspö Hard Rock Laboratory, where experiments concerning the geological disposal of nuclear waste are performed. In bacterial culture extracts, the iron in the siderophore complexes was replaced by gallium to facilitate siderophore identification by mass spectrometry. P. stutzeri was shown to produce ferrioxamine E (nocardamine) as the main siderophore together with ferrioxamine G and two cyclic ferrioxamines having molecular masses 14 and 28 atomic mass units lower than that of ferrioxamine E, suggested to be ferrioxamine D2 and ferrioxamine X1, respectively. In contrast, no siderophores were observed from anaerobically grown P. stutzeri. None of the siderophores produced by aerobically grown P. stutzeri were found in anaerobic natural water samples from the Äspö Hard Rock Laboratory.In order to facilitate iron(III) acquisition, plants and microorganisms, such as fungi and bacteria, produce and excrete strong iron(III) chelators, i.e., siderophores (18, 22, 23, 33, 34). While fungal siderophores bind to iron(III) by hydroxamate ligands, bacterial siderophores are more structurally diverse, and common ligands are catecholates, hydroxamates, and carboxylates (21). The iron(III) stability constants for bacterial siderophores vary in the range of 1020 to 1052 (6). In addition to iron(III), other metals can be complexed by siderophores. For the trihydroxamate siderophore desferrioxamine B, sometimes called proferrioxamine B (10), some actinides have been shown to have stability constants in the same range as the ferric stability constant (1030.6), e.g., 1026.6 with thorium(IV) and 1030.8 with plutonium(VI) (32), while the stability constant for uranium(VI) was lower, i.e., 1018 (2).Concerning bacteria, there are several reports on siderophore production by Pseudomonas spp. (1, 3, 4, 19). More than 50 structurally related siderophores, i.e., pyoverdins, produced by the fluorescent Pseudomonas spp., especially Pseudomonas fluorescens and Pseudomonas aeruginosa, have been characterized (3). All pyoverdins emit yellow fluorescent light due to the presence of a 5-amino-2,3-dihydro-8,9-dihydroxy-1-H-pyrimido-quinoline-carboxylic chromophore, to which a peptide chain and a carboxyl chain are attached (1, 3). Nonfluorescent Pseudomonas has also been shown to produce siderophores, such as ferrioxamine E, also called nocardamine (Fig. (Fig.1),1), which was produced by one strain of Pseudomonas stutzeri (19). In addition to ferrioxamines, the P. stutzeri strain KC produced a smaller siderophore, i.e., pyridine-2,6-bis(thiocarboxylic acid) (35). Conversely, a catecholate-type siderophore was shown to be produced by another strain of P. stutzeri, which did not produce any hydroxamate siderophores (4).Open in a separate windowFIG. 1.Structures, molecular masses (mw), and stability constants (Ks) of ferric complexes of the three ferrioxamines: ferrioxamine B (B), ferrioxamine E (E), and ferrioxamine G (G) (5, 18).Most of the studies on bacterial siderophore production have been conducted on microorganisms growing under aerobic conditions. One field-based report, however, indicates the occurrence of putative siderophores in anaerobic environments also (29). In the present study, siderophore production has been studied with both aerobic and anaerobic cultures of P. stutzeri. This species is a facultative aerobe, able to grow with oxygen or nitrate as the electron acceptor, meaning that it can be active under both anaerobic and aerobic conditions. The P. stutzeri strain CCUG 36651, studied here, has been isolated from a depth of 626 m below ground at the Äspö Hard Rock Laboratory (16), where research concerning the geological disposal of nuclear waste is performed. The possibility of mobilizing radionuclides by complexing compounds from bacteria is an important research area in the context of nuclear waste disposal research. It is unknown if such compounds are produced in aquifers under conditions relevant to a disposal site, which would be approximately 500 m underground in granitic rock (27).A study from 2004 shows that P. stutzeri growing aerobically in the presence of uranium-containing shale leached Fe, Mo, V, and Cr from the shale material (17). More recently it was shown that the supernatant of aerobically and anaerobically cultured P. stutzeri was able to increase the partitioning of added Fe, Pm, Am, and Th into the aqueous phase in samples where quartz sand was used as a solid surface (16). Aerobic supernatants maintained 60% or more of the added metals in solution, while anaerobic supernatants were best at maintaining Am in solution, reaching a value of 40% in solution. The increased partitioning to the aqueous phase in the presence of the supernatants was ascribed to the production of organic ligands. Supernatants of both aerobically and anaerobically grown P. stutzeri strain CCUG 36651 yielded a positive response on the universal siderophore assay, the CAS assay (16). This assay is based on ligand competition for iron bound to the colored chrome azurol complex (25, 30).In this study, siderophore production by P. stutzeri strain CCUG 36651 was investigated using mass spectrometry (MS) and liquid chromatography (LC) followed by mass spectrometric detection. Electrospray ionization mass spectrometry (ESI-MS) and electrospray ionization tandem mass spectrometry (ESI-MS/MS) are useful tools in characterizing siderophores such as ferrioxamines (10, 13, 14, 28, 31). In order to detect iron(III)-chelating compounds, the ferric iron can be replaced by gallium(III) through ascorbate-mediated reduction of iron(III) (8, 20). In mass spectra, gallium-bound substances are easily recognized due to the characteristic isotope pattern of gallium, where the intensity of the 71Ga signal is about 66% of that of the 69Ga signal. The use of ESI provides so-called soft ionization; thus, information about the molecular weight is obtained. However, by employing MS/MS, fragmentation is achieved, providing more information about the compound structure.In order to verify the chemical difference between the siderophores found by ESI-MS, chromatographic separation was performed. In this case, one reversed-phase C18 column and one column containing a porous graphitic carbon (PGC) stationary phase were used. The separation mechanism of PGC is a combination of hydrophobic interactions, as in C18, and electrostatic interactions between π-electrons. In order to detect substances at low concentrations, column-switched capillary chromatography with MS detection was used. The detection limits of the combined LC-MS/MS system used in this study are in the range of 1 to 5 nM for hydroxamate siderophores of the ferrichrome and ferrioxamine families (9). In order to facilitate analysis of lower concentrations of ferrioxamines, natural water samples were preconcentrated by solid-phase extraction (SPE), resulting in minimum detectable concentrations in the range of 0.02 to 0.1 nM, depending on the initial sample volume.  相似文献   

16.
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   

17.
18.
The biochemical and molecular mechanisms used by alkaliphilic bacteria to acquire iron are unknown. We demonstrate that alkaliphilic (pH > 9) Bacillus species are sensitive to artificial iron (Fe3+) chelators and produce iron-chelating molecules. These alkaliphilic siderophores contain catechol and hydroxamate moieties, and their synthesis is stimulated by manganese(II) salts and suppressed by FeCl3 addition. Purification and mass spectrometric characterization of the siderophore produced by Caldalkalibacillus thermarum failed to identify any matches to previously observed fragmentation spectra of known siderophores, suggesting a novel structure.Iron is an abundant element in nature; however, in most aqueous aerobic environments iron forms insoluble ferric hydroxide, Fe(OH)3. This poses a major problem for most aerobic bacteria, as ferric hydroxide has a solubility constant of 10−39 M, therefore limiting the concentration of ferric ions to 10−18 M at pH 7.0. For example, bacteria living in seawater (approximate pH 8.0) require iron, yet dissolved iron is only present at 0.02 to 2.0 nM (5). Despite this apparent lack of bioavailability, iron has been repeatedly demonstrated to be an essential element for aerobic bacterial growth (1).With the lack of readily accessible iron at physiological pH, most bacteria have evolved systems to deal with the incumbent problem of iron acquisition. Under iron-rich conditions, Fe2+ uptake receptors, such as FeoAB, are synthesized in bacteria, which passively import iron in the immediate vicinity of the cell (1, 23). No equivalent system has been identified for Fe3+ transport. To acquire Fe3+ under aqueous aerobic conditions, bacteria commonly have import systems involving the synthesis, secretion, and regathering of a group of secondary metabolites known as siderophores (1, 11). Siderophores are low-molecular-weight chemical moieties that chelate Fe3+ and typically have complex formation (Kf) constants in the range of 1023 to 1052 (11). Siderophores, like other chelators, are known to increase the solubility of iron by hindering the formation of Fe-oxyhydroxides at high pH, at which the Fe-oxyhydroxides are the dominating inorganic species (27). Siderophores are also known to facilitate the dissolution of Fe from minerals (3). Siderophore-iron complexes can either be transported through cellular membranes using dedicated transport systems or if the Fe(III) central atom is reduced, making the iron bioavailable for cellular processes (10, 14). Three major groups of siderophores have been described in bacteria: hydroxamates, catecholates, and carboxylates. Hydroxamates and catechols are commonly produced by aerobic bacteria living at neutral to alkaline pH, whereas carboxylates are significantly more common in bacteria living in mildly acidic pH (11-13). In the genus Bacillus, Bacillus megaterium and Bacillus subtilis are producers of schizokinen and bacillibactin, respectively (6, 20). Bacillus anthracis produces both a catechol and a hydroxamate siderophore (7, 34), and B. licheniformis strain VK21 is the only known example of a thermoresistant catecholate-producing Gram-positive bacterium (32).Although there is extensive literature on iron capture mechanisms in bacteria that thrive at neutral pH, there is little information at a biochemical or molecular level on how aerobic bacteria growing at extreme alkaline pHs (i.e., pH 9 to 11) acquire iron. At alkaline pH, the solubility constant for iron decreases far below the requirement for living cells, and the concentration of bioavailable iron is estimated to be approximately 10−23 M at pH 10 (11). Taking this extreme lack of iron into account, the sequestering mechanisms of alkaliphilic bacteria must be powerful, yet there has been little analysis of the types of iron-chelating molecules these bacteria produce.  相似文献   

19.
Summary Geotrichum candidum (isolate 1–9) pathogenic on citrus fruits, appears to lack siderophore production. Iron uptake byG. candidum is mediated by two distinct iron-regulated, energy-and temperature-dependent transport systems that require sulfhydryl groups. One system exhibits specificity for either ferric or ferrous iron, whereas the other exhibits specificity for ferrioxamine-B-mediated iron uptake and presumably other hydroxamate siderophores. Radioactive iron uptake from59FeCl3 showed an optimum at pH 6 and 35° C, and Michaelis-Menten kinetics (apparentK m = 3 m,V max = 0.054 nmol · mg–1 · min–1). The maximal rate of Fe2+ uptake was higher than Fe3+ (V max = 0.25 nmol · mg–1 · min–1) but theK m was identical. Reduction of ferric to ferrous iron prior to transport could not be detected. The ferrioxamine B system exhibits an optimum at pH 6 and 40° C and saturation kinetics (K m = 2 M,V max = 0.22 nmol · mg–1 · min–1). The two systems were distinguished as two separate entities by negative reciprocal competition, and on the basis of differential response to temperature and phenazine methosulfate. Mössbauer studies revealed that cells fed with either57FeCl3 or57FeCl2 accumulated unknown ferric and ferrous binding metabolites.  相似文献   

20.

Aims

Reclamation following oil sands mining in northeastern Alberta (Canada) creates adverse reforestation soil conditions, including extreme pH values. We elucidated pH tolerance limits of boreal plant species and how pH affects nutrient uptake in these plants.

Methods

We measured growth, gas exchange, and foliar nutrient concentration of 15 common northern boreal forest plants after eight weeks exposure to root zone pH ranging from 5.0 to 9.0. Cluster analyses were used to group these species based on their pH responses.

Results

Based on their growth and gas exchange responses to pH, the 15 plant species could be divided into five groups, each of which contained species that commonly co-occur in particular boreal forest site types. For the foliar nutrient responses to pH, the 15 species could be grouped into only two categories; both showed decreases in foliar N, P, Fe and Zn concentration with increasing pH, with a more pronounced effect on the group that included trembling aspen, paper birch and chokecherry.

Conclusions

The evidence of differential adaptation to pH by habitat type suggests the importance of soil pH as a factor affecting boreal plant species distribution and could be helpful for selection of species suitable for reclamation of sites with altered soil pH.
  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