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.
There is good evidence to show that ferric enterochelin is an essential growth factor for a number of Gram-negative pathogenic bacteria exposed to the host iron binding proteins, transferrin and lactoferrin. Tests of nineteen complexes of enterochelin as potential antibacterial agents showed that only those containing either indium (In3+) or scandium (Sc3+) inhibited bacterial growth. In this study, further evidence is presented which demonstrates a competition between the Sc3+ and Fe3+ complexes. The uptake of both complexes is energy dependent and is also repressed in iron-replete cells. The Sc3+ complex accumulates within the cells at 20% of the rate of the Fe3+ complex. The main components of the ferric enterochelin transport system are required for the transport of the Sc3+ complex although some Sc3+ appears to enter the cell by another route. The accumulation, within the cell, of 14C-labelled enterochelin complexes depends on the growth medium. The relationship of the size of the metal ion to the biological activity of the complex is discussed and possible mechanisms of action of the Sc3+ complex are considered.  相似文献   

2.
Enterochelin, the iron chelator produced by a number of pathogenic enterobacteria, appears to be an essential metabolite for multiplication within the host, where it transports iron from the host iron-binding proteins to the bacteria. Previous work showed that complexes of enterochelin containing either scandium (Sc3+) or indium (In3+) exerted a bacteriostatic effect on Klebsiella pneumoniae in serum, whilst the Sc3+ complex exerted a significant therapeutic effect on mice infected with K. pneumoniae. These observations have now been extended to a number of pathogenic serotypes of Escherichia coli including those carrying either the K1 antigen or the ColV plasmid. The Sc3+ and In3+ complexes each exert a bacteriostatic effect on these organisms growing in either whole serum or media containing an iron-binding protein. Evidence is presented that the Sc3+ complex may act as a competitive inhibitor of the Fe3+ complex. In contrast to their effects on K. pneumoniae, sideramines other than enterochelin fail to reverse the bacteriostatic effect of the Sc3+ complex of enterochelin in E. coli, suggesting that the complex produces a more profound derangement of metabolism in this organism. The Sc3+ complex exerts a significant therapeutic effect on E. coli infections in mice although the In3+ complex is less active.  相似文献   

3.
When a strain (arcB-) of Escherichia coli, unable to synthesize the iron transport compound enterochelin, was transduced to tonB-, it became resistant to phage phi80 and simultaneously lost the growth response to enterochelin and the ability to transport its iron complex. However, enterochelin precursors (shikimate and 2,3-dihydroxybenzoate) still supported growth, via the synthesis of enterochelin. Dihydroxybenzoate was a better growth factor at a low concentration than it was at higher levels. The evidence suggests that tonB- strains lack an outer membrane component necessary both for the uptake of ferric-enterochelin and for the adsorption of phage phi80. Thus, although ferric-enterochelin cannot penetrate the cell surface from outside, the complex that is formed within the envelope is transported normally into the cell. The aroB-, tonB- mutant also lacked growth reponses to citrate and various hydroxamate siderochromes, which supported growth in the tonB+ parent strain via inducible transport systems for their ferric complexes. The aroB-, tonB- mutant was unable to transport iron in the presence of citrate, but the low-affinity uptake of uncomplexed iron and the transport of amino acids and phosphate were unimpaired. The tonB locus, thus, affects all the known active transport systems for iron, possibly indicating that they share some common outer membrane component.  相似文献   

4.
The fes mutation in Escherichia coli K12, which inactivates enterochelin esterase, allows the cell to accumulate ferric enterochelin. The ferric complex of enterochelin was released in significant quantities from a fes mutant after osmotic shock. Analysis of the effects of the individual stages of the shock procedure in wild-type cells showed that prior exposure of cells to sucrose and EDTA was not required, careful dilution of cells into a hypo-osmolar medium being sufficient to induce efflux of Fe3+. Prior treatment with EDTA or exposure to shearing forces served either to enhance efflux or to induce efflux in isotonic media. Neither vitamin B12 nor 5'-nucleotidase was released from the periplasm by these procedures. The release observed under mild conditions was stimulated specifically by Co2+, did not occur at 0 degree C, and was inhibited by 2,4-dinitrophenol at 37 degrees C. From these observations, it was concluded that the efflux of Fe3+ represents a physiological response of the cell to exposure to a hypo-osmolar medium. Such changes may enhance survival following physicochemical stressing of the bacterial outer membrane.  相似文献   

5.
The specificity of the outer membrane protein receptor for ferric enterobactin transport in Escherichia coli and the mechanism of enterobactin-mediated transport of ferric ions across the outer membrane have been studied. Transport kinetic and inhibition studies with ferric enterobactin and synthetic structural analogs have mapped the parts of the molecule important for receptor binding. The ferric complex of the synthetic structural analog of enterobactin, 1,3,5-N,N',N'-tris-(2,3-dihydroxybenzoyl)triaminomethylbenzene (MECAM), was transported with the same maximum velocity as was ferric enterobactin. A double-label transport assay with [59Fe, 3H]MECAM showed that the ligand and the metal are transported across the outer membrane at an identical rate. Under the growth conditions used, large fractions of the transported complexes were available for exchange across the outer membrane when a large excess of extracellular complex was added to the cell suspension; at least 60% of the internalized [59Fe]enterobactin exchanged with extracellular [55Fe]enterobactin. Internalized [59Fe, 3H]MECAM was released from the cell as the intact complex when either unlabeled Fe-MECAM or Fe-enterobactin was added extracellularly. The results suggest a mechanism of active transport of unmodified coordination complex across the outer membrane with possible accumulation in the periplasm.  相似文献   

6.
Spheroplasts of Escherichia coli mutants were used to investigate the roles of the inner and outer membranes in the transport of iron. tonA mutants, known to be defective in an outer membrane component of the ferrichrome transport system, regained the ability to transport ferrichrome when converted to spheroplasts. On the other hand, the tonB mutant was unable to transport ferric enterochelin in either whole cells or spheroplasts. This implies that an element of the inner membrane is affected. fep mutants were also unable to transport ferric enterochelin, and fell into two classes, fepA and fepB. Spheroplasts of the former class transported ferric enterochelin, and those of the latter did not. This implies that the fepA mutants are defective in ferric enterochelin transport across the outer membrane, and that fepB mutants probably lack the facility to transport ferric enterochelin across the inner membrane. Colicin B action on fepA mutants was found to differ from that on fepB mutants.  相似文献   

7.
Bacteria solubilize iron (Fe(3+)) with secreted siderophores, which are then taken up as Fe(3+)-siderophore complexes. Some bacteria also use iron in heme, hemoglobin, hemopexin, transferrin and lactoferrin of eukaryotic hosts. Crystal structures of two outer membrane transport proteins, FhuA and FepA, and biochemical data reveal strong long-range conformational changes of the proteins upon binding of Fe(3+)-siderophore complexes and in response to energy transfer from the cytoplasmic membrane into the outer membrane via the TonB-ExbB-ExbD protein complex. The crystal structure of the periplasmic binding protein FhuD strongly deviates from the uniform overall structure of binding proteins hitherto determined. Sideromycins, antibiotics that contain Fe(3+)-siderophore complexes as carriers, are highly effective, as they enter cells via Fe(3+)-siderophore transport systems. In this review, recently published data is discussed to demonstrate the state of understanding of iron transport across the outer membrane and the cytoplasmic membrane.  相似文献   

8.
Cr(3+), similar to Fe(3+), is transported into cells primarily via endocytosis as the metal-transferrin complex. As Cr(3+) ions are not readily reduced under biological conditions, the ion cannot be transported from endosomes by the same mechanism as iron that utilized divalent metal ion transporters. Cr(3+) has been hypothesized to potentially be transported as small ligand complexes with a free carboxylate functionality by monocarboxylate transporters (MCT), in a similar fashion to that proposed for Al(3+). Consequently, mouse C2C12 muscle cells were utilized to determine if Cr(3+) is potentially transported by MCT by examining the effects of MCT inhibitors on Cr and Fe transport and subcellular distribution when the metals are added as their transferrin complexes. The results suggest that Cr is not primarily transported by MCT from the endosomes to the cytosol, and that another mechanism for this transport needs to be identified.  相似文献   

9.
Mutations Affecting Iron Transport in Escherichia coli   总被引:14,自引:20,他引:14       下载免费PDF全文
A mutant of Escherichia coli K-12 unable to form an essential component of the enterochelin-dependent iron transport system has been isolated. This strain carries a mutation in a gene designated fep, mapping close to two genes, entA and entD, concerned with enterochelin synthesis. Strain AN102, which carries the fep(-) allele, accumulates large quantities of enterochelin and gives a growth response to sodium citrate. The cytochrome b(1) and total iron content, and the measurement of the uptake of (55)Fe(3+), indicate an impairment of the enterochelin-dependent iron transport system. The growth response to sodium citrate is related to the presence, in strain AN102, of an inducible citrate-dependent iron transport system.  相似文献   

10.
Accumulation of iron by yersiniae.   总被引:40,自引:2,他引:38       下载免费PDF全文
Escherichia coli, Bacillus megaterium, and three species of yersiniae grew rapidly without significant production of soluble siderophores in a defined iron-sufficient medium (20 microM Fe3+). In iron-deficient medium (0.1 to 0.3 microM Fe3+) all organisms showed reduced growth, and there was extensive production of siderophores by E. coli and B. megaterium. Release of soluble siderophores by Yersinia pestis, Y. pseudotuberculosis, or Y. enterocolitica in this medium was not detected. Citrate (1 mM) inhibited growth of yersiniae in iron-deficient medium, indicating that the organisms lack an inducible Fe3+-citrate transport mechanism. Uptake of 59Fe3+ by all yersiniae was an energy-dependent saturable process, showing increased accumulation after adaptation to iron-deficient medium. Growth of Y. pseudotuberculosis and Y. enterocolitica but not Y. pestis on iron-limited solid medium was enhanced to varying degrees by exogenous siderophores (desferal, schizokinen, aerobactin, and enterochelin). Only hemin (0.1 pmol) or a combination of inorganic iron plus protoporphyrin IX promoted growth of Y. pestis on agar rendered highly iron deficient with egg white conalbumin (10 microM). Growth of Y. pseudotuberculosis and Y. enterocolitica was stimulated on this medium by Fe3+ or hemin. These results indicate that hemin can serve as a sole source of iron for yersiniae and that the organisms possess an efficient cell-bound transport system for Fe3+.  相似文献   

11.
The fate of the hydroxamic acid-iron transport cofactors during iron uptake from the (59)Fe(3+) chelates of the (3)H-labeled hydroxamates schizokinen and aerobactin was studied by assay of simultaneous incorporation of both (59)Fe(3+) and (3)H. In the schizokinen-producing organism Bacillus megaterium ATCC 19213 transport of (59)Fe(3+) from the (3)H-schizokinen-(59)Fe(3+) chelate at 37 C was accompanied by rapid uptake and release (within 2 min) of (3)H-schizokinen, although (3)H-schizokinen discharge was temperature-dependent and did not occur at 0 C. In the schizokinen-requiring strain B. megaterium SK11 similar release of (3)H-schizokinen occurred only at elevated concentrations of the double-labeled chelate; at lower chelate concentrations, (3)H-schizokinen remained cell-associated. Temperature-dependent uptake of deferri (iron-free) (3)H-schizokinen to levels equivalent to those incorporated from the chelate form was noted in strain SK11, but strain ATCC 19213 showed only temperature-independent binding of low concentrations of deferri (3)H-schizokinen. These results indicate an initial temperature-independent binding of the ferric hydroxamate which is followed rapidly by temperature-dependent transport of the chelate into the cell and an enzyme catalyzed separation of iron from the chelate. The resulting deferri hydroxamate is discharged from the cell only when a characteristic intracellular concentration of the hydroxamate is exceeded, which happens in the schizokinen-requiring strain only at elevated concentrations of the chelate. This strain also appears to draw the deferri hydroxamate into the cell by a temperature-dependent mechanism. The aerobactin-producing organism Aerobacter aerogenes 62-1 also demonstrated rapid initial uptake and temperature-dependent discharge of (3)H-aerobactin during iron transport from (3)H-aerobactin-(59)Fe(3+), suggesting a similar ferric hydroxamate transport system in this organism.  相似文献   

12.
The mechanism by which iron uptake is facilitated by the siderophore rhodotorulic acid (RA) in the yeast Rhodotorula pilimanae was investigated with radioactively labeled Fe and RA and kinetically inert, chromic-substituted RA complexes. The weight of the evidence supports a model in which RA mediates iron transport to the cell but does not actually transport iron into the cell. It is proposed that RA exchanges the ferric ion at the cell surface with a membrane-bound chelating agent that completes the active transport of iron into the cell. Uptake of 55Fe in ferric rhodotorulate was much more rapid than uptake of RA itself. Two exchange-inert chromic complexes of RA showed no uptake.  相似文献   

13.
14.
Uptake studies with [14C]picolinate and 55Fe3+ have provided an explanation for the change in streptonigrin killing on adaptation of Escherichia coli to picolinate, in terms of the available iron within the cell. When picolinic acid is added to a growing culture of E. coli an interval of bacteriostasis ensues; this adaptation period is followed by resumption of exponential growth. Addition of picolinate (4 mM) to a log phase culture of strain W3110 gave protection from the lethal action of streptonigrin (30 microM) when the two agents were added simultaneously. In contrast streptonigrin killed cells that had adapted to picolinate; however, a preincubation of adapted W3110 with phenethyl alcohol protected the cells from streptonigrin lethality. [14C]Picolinate uptake studies showed that initially picolinate entered the cells, but that it was excluded from adapted cells; addition of phenethyl alcohol permitted the entry of picolinate into adapted W3110. The changes in streptonigrin killing parallel the changes in concentration of intracellular picolinate, which can chelate the iron required by streptonigrin for its bactericidal action. 55Fe3+ uptake studies showed that initially picolinate prevented iron accumulation by strain W3110, whereas adapted cells did take up iron in the presence of picolinate. Addition of phenethyl alcohol prevented any observed uptake of iron by adapted W3110. This modulation of iron transport by picolinate also affects streptonigrin lethality. Experiments with iron transport mutants showed that picolinate acted on both the enterochelin and citrate routes of uptake. Therefore picolinate affects the concentration of available iron within the cell both by (a) its intracellular presence resulting in chelation of iron and (b) its action on iron uptake; these effects explain the change in streptonigrin killing on adaptation of E. coli to picolinate.  相似文献   

15.
Iron supply to Escherichia coli by synthetic analogs of enterochelin.   总被引:10,自引:6,他引:4       下载免费PDF全文
Synthetic analogs of enterochelin (enterobactin) were tested for their ability to support the growth of Escherichia coli K-12 under iron-limiting conditions. The cyclic compound MECAM [1,3,5-N.N'; N"-tris-(2,3-dihydroxybenzoyl)-triamino-methylbenzene] and its N-methyl derivative Me3MECAM promoted growth, whereas the 2,3-dihydroxy-5-sulfonyl derivatives MECAMS and Me3MECAMS were inactive. The same results were obtained with TRIMCAM [1,3,5-tris(2,3-dihydroxybenzoylcarbamido)-benzene] and TRIMCAMS (the 2,3-dihydroxy-5-sulfonyl derivative of TRIMCAM). However, the sulfonic acid-containing linear compound LICAMS [1,5,10-N,N', N"-tris(5-sulfo-2,3-dihydroxybenzoyl)-triaza-decane] supported growth. In contrast, LIMCAMC, in which the sulfonyl groups at the five position of LICAMS are replaced by carboxyl groups at the four position, was inactive. The uptake of the active analogs required the functions specified by the fepB, fesB, and tonB genes. Surprisingly, growth promotion of mutants lacking the enterochelin receptor protein in the outer membrane was observed. Only MECAM protected cells against colicin B (which kills cells after entering at the enterochelin uptake sites) and transported Fe3+ at about half the enterochelin rate.  相似文献   

16.
Bacterioferritins are members of a class of spherical shell-like iron storage proteins that catalyze the oxidation and hydrolysis of iron at specific sites inside the protein shell, resulting in formation of a mineral core of hydrated ferric oxide within the protein cavity. Electrode oximetry/pH stat was used to study iron oxidation and hydrolysis chemistry in E. coli bacterioferritin. Consistent with previous UV-visible absorbance measurements, three distinct kinetic phases were detected, and the stoichiometric equations corresponding to each have been determined. The rapid phase 1 reaction corresponds to pairwise binding of 2 Fe(2+) ions at a dinuclear site, called the ferroxidase site, located within each of the 24 subunits, viz., 2Fe(2+) + P(Z) --> [Fe(2)-P](Z) + 4H(+), where P(Z) is the apoprotein of net charge Z and [Fe(2)-P](Z) represents a diferrous ferroxidase complex. The slower phase 2 reaction corresponds to the oxidation of this complex by molecular oxygen according to the net equation: [Fe(2)-P](Z) + (1)/(2)O(2) --> [Fe(2)O-P](Z) where [Fe(2)O-P](Z) represents an oxidized diferric ferroxidase complex, probably a mu-oxo-bridged species as suggested by UV-visible and EPR spectrometric titration data. The third phase corresponds to mineral core formation according to the net reaction: 4Fe(2+) + O(2) + 6H(2)O --> 4FeO(OH)((core)) + 8H(+). Iron oxidation is inhibited by the presence of Zn(2+) ions. The patterns of phase 2 and phase 3 inhibition are different, though inhibition of both phases is complete at 48 Zn(2+)per 24mer, i.e., 2 Zn(2+) per ferroxidase center.  相似文献   

17.
Complementation of insertion mutants showed that the polypeptides FatD, FatC, FatB, and FatA are essential for the iron-transport process encoded by pJM1. Sequence analysis followed by homology studies indicated that transport of ferric anguibactin into Vibrio anguillarum 775 follows the same mechanism as reported for transport of Fe(3+)-hydroxamates, Fe(3+)-catecholates, ferric dicitrate, and vitamin B12 into Escherichia coli. Homology of FatA, part of the receptor complex, to seven E. coli receptor proteins involved in uptake of siderophores and vitamin B12 supports the idea of a common ancestral gene. A "TonB-Box" was found in FatA suggesting the existence of a TonB-like protein function in V. anguillarum. A high homology in the primary structure of FatB to FhuD, FecB, FepB, and BtuE suggests that FatB is the anguibactin-binding protein located in the periplasmic space. FatD and FatC are polytopic integral membrane proteins. According to their homologies to other proteins from other transport systems, they may be involved in the translocation of ferric anguibactin across the cytoplasmic membrane.  相似文献   

18.
Maes EM  Roberts SA  Weichsel A  Montfort WR 《Biochemistry》2005,44(38):12690-12699
Nitrophorin 4 (NP4), a nitric oxide (NO)-transport protein from the blood-sucking insect Rhodnius prolixus, uses a ferric (Fe3+) heme to deliver NO to its victims. NO binding to NP4 induces a large conformational change and complete desolvation of the distal pocket. The heme is markedly nonplanar, displaying a ruffling distortion postulated to contribute to stabilization of the ferric iron. Here, we report the ferrous (Fe2+) complexes of NP4 with NO, CO, and H2O formed after chemical reduction of the protein and the characterization of these complexes by absorption spectroscopy, flash photolysis, and ultrahigh-resolution crystallography (resolutions vary from 0.9 to 1.08 A). The absorption spectra, both in solution and in the crystal, are typical for six-coordinated ferrous complexes. Closure and desolvation of the distal pocket occurs upon binding CO or NO to the iron regardless of the heme oxidation state, confirming that the conformational change is driven by distal ligand polarity. The degree of heme ruffling is coupled to the nature of the ligand and the iron oxidation state in the following order: (Fe3+)-NO > (Fe2+)-NO > (Fe2+)-CO > (Fe3+)-H2O > (Fe2+)-H2O. The ferrous coordination geometry is as expected, except for the proximal histidine bond, which is shorter than typically found in model compounds. These data are consistent with heme ruffling and coordination geometry serving to stabilize the ferric state of the nitrophorins, a requirement for their physiological function. Possible roles for heme distortion and NO bending in heme protein function are discussed.  相似文献   

19.
There is an increasing interest in the use of lanthanides in medicine. However, the mechanism of their accumulation in cells is not well understood. Lanthanide cations are similar to ferric ions with regard to transferrin binding, suggesting transferrin-receptor mediated transport is possible; however, this has not yet been confirmed. In order to clarify this mechanism, we investigated the binding of Yb3+ to apotransferrin by UV-Vis spectroscopy and stopped-flow spectrophotometry, and found that Yb3+ binds to apotransferrin at the specific iron sites in the presence of bicarbonate. The apparent binding constants of these sites showed that the affinity of Yb3+ is lower than that of Fe3+and binding of Yb3+ in the N-lobe is kinetically favored while the C-lobe is thermodynamically favored. The first Yb3+ bound to the C-lobe quantitatively with a Yb/apotransferrin molar ratio of < 1, whereas the binding to the other site is weaker and approaches completeness by a higher molar ratio only. As demonstrated by 1H NMR spectra, Yb3+ binding disturbed the conformation of apotransferrin in a manner similar to Fe3+. Flow cytometric studies on the uptake of fluorescein isothiocyanate labeled Yb3+-bound transferrin species by K562 cells showed that they bind to the cell receptors. Laser scanning confocal microscopic studies with fluorescein isothiocyanate labeled Yb3+-bound transferrin and propidium iodide labeled DNA and RNA in cells indicated that the Yb3+ entered the cells. The Yb3+-transferrin complex inhibited the uptake of the fluorescein labeled ferric-saturated transferrin (Fe2-transferrin) complex into K562 cells. The results demonstrate that the complex of Yb3+-transferrin complex was recognized by the transferrin receptor and that the transferrin-receptor-mediated mechanism is a possible pathway for Yb3+ accumulation in cells.  相似文献   

20.
Iron deposition within the iron storage protein ferritin involves a complex series of events consisting of Fe(2+) binding, transport, and oxidation at ferroxidase sites and mineralization of a hydrous ferric oxide core, the storage form of iron. In the present study, we have examined the thermodynamic properties of Fe(2+) binding to recombinant human H-chain apoferritin (HuHF) by isothermal titration calorimetry (ITC) in order to determine the location of the primary ferrous ion binding sites on the protein and the principal pathways by which the Fe(2+) travels to the dinuclear ferroxidase center prior to its oxidation to Fe(3+). Calorimetric titrations show that the ferroxidase center is the principal locus for Fe(2+) binding with weaker binding sites elsewhere on the protein and that one site of the ferroxidase center, likely the His65 containing A-site, preferentially binds Fe(2+). That only one site of the ferroxidase center is occupied by Fe(2+) implies that Fe(2+) oxidation to form diFe(III) species might occur in a stepwise fashion. In dilute anaerobic protein solution (3-5 microM), only 12 Fe(2+)/protein bind at pH 6.51 increasing to 24 Fe(2+)/protein at pH 7.04 and 7.5. Mutation of ferroxidase center residues (E62K+H65G) eliminates the binding of Fe(2+) to the center, a result confirming the importance of one or both Glu62 and His65 residues in Fe(2+) binding. The total Fe(2+) binding capacity of the protein is reduced in the 3-fold hydrophilic channel variant S14 (D131I+E134F), indicating that the primary avenue by which Fe(2+) gains access to the interior of ferritin is through these eight channels. The binding stoichiometry of the channel variant is one-third that of the recombinant wild-type H-chain ferritin whereas the enthalpy and association constant for Fe(2+) binding are similar for the two with an average values (DeltaH degrees = 7.82 kJ/mol, binding constant K = 1.48 x 10(5) M(-)(1) at pH 7.04). Since channel mutations do not completely prevent Fe(2+) binding to the ferroxidase center, iron gains access to the center in approximately one-third of the channel variant molecules by other pathways.  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