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 234 毫秒
1.
The unfolding-refolding of proteins is a cooperative process and, as judged by equilibrium properties, occurs in one step involving the native,N, and the unfoldedU, conformational states. Kinetic studies have shown that the denatured protein exists as a mixture of slow-(U)Sand fast-(U)Frefolding forms produced by proline peptidecis-trans isomerization. Proline residues inU Fare in the same configuration as in the native protein while they are in non-native configuration inU S. For protein folding to occur quicklyU Smust be converted intoU F. The fact that the equilibrium and kinetic properties of are the same as those found for prolinecis-trans isomerization taken together with the absence of slow phase in the kinetics of refolding of a protein devoid of proline, support this view. However, the absence of a linear correlation between half-time of reactivation of denatured enzymes and their proline-contents, as well as the dissimilarities in the kinetic properties of in unfolding and refolding experiments are not consistent with the model. Conformational energy calculation and experimental results on refolding of proteins suggest that some proline residues are non-essential. They will not block protein folding even in wrong isomeric form. The native-like folded structure with incorrect proline isomers will serve as intermediate state(s) in which these prolines will more readily isomerize to the correct isomeric form. The picture becomes more complex when one considers the consequence ofcis-trans isomerism of non-proline residues on protein folding.  相似文献   

2.
In Bradyrhizobium japonicum, iron uptake from ferric siderophores involves selective outer membrane proteins and non-selective periplasmic and cytoplasmic membrane components that accommodate numerous structurally diverse siderophores. Free iron traverses the cytoplasmic membrane through the ferrous (Fe2+) transporter system FeoAB, but the other non-selective components have not been described. Here, we identify fsrB as an iron-regulated gene required for growth on iron chelates of catecholate- and hydroxymate-type siderophores, but not on inorganic iron. Utilization of the non-physiological iron chelator EDDHA as an iron source was also dependent on fsrB. Uptake activities of 55Fe3+ bound to ferrioxamine B, ferrichrome or enterobactin were severely diminished in the fsrB mutant compared with the wild type. Growth of the fsrB or feoB strains on ferrichrome were rescued with plasmid-borne E. coli fhuCDB ferrichrome transport genes, suggesting that FsrB activity occurs in the periplasm rather than the cytoplasm. Whole cells of an fsrB mutant are defective in ferric reductase activity. Both whole cells and spheroplasts catalyzed the demetallation of ferric siderophores that were defective in an fsrB mutant. Collectively, the data support a model whereby FsrB is required for reduction of iron and its dissociation from the siderophore in the periplasm, followed by transport of the ferrous ion into the cytoplasm by FeoAB.  相似文献   

3.
We report here that Vibrio anguillarum possesses a non-inducible active transport system which can efficiently supply iron to the cell from ferric citrate, independently of the siderophore-based mechanisms. The strains tested were able to grow in CM9 medium in iron-restricted conditions when ferric citrate was present in the medium. Moreover, the presence of ferric citrate inhibited the production of siderophores in the strains tested. V. anguillarum cells and isolated membranes could incorporate 55Fe3+ complexed by citrate, without a difference between cells grown in the presence or absence of ferric citrate. The presence of 2,4-dinitrophenol, ferrozine, ferricyanide, trypsin, as well as low temperature produced a marked decrease or total inhibition of 55Fe3+ uptake by the cells. All these results suggest that iron uptake from ferric citrate in V. anguillarum must be an energy-dependent process not induced by the presence of iron or citrate in the medium, mediated by a membrane protein(s), which may require an iron reduction step to function.  相似文献   

4.
Based on in vivo Mössbauer spectroscopy it has previously been demonstrated that the intracellular iron pool of Escherichia coli, grown in iron deficient media supplemented with siderophores as the sole iron source, is dominated by a single Fe2+ and a single Fe3+ species. We have isolated the ferrous ion species and have purified it employing native column PAGE, chromatography and ultrafiltration. The purified compound displays an M app of 2.2 kDa and an extremely low isoelectric point (pI) of 1.05. It is shown that this ferrous ion binding compound is neither a protein nor a nucleotide, rather it is composed mainly of phosphorylated sugar derivatives. This compound binds approximately 40% of the cytoplasmic iron. Therefore it is proposed that this oligomeric ferrous carbohydrate phosphate represents the long sought after mobile, low molecular mass iron pool.  相似文献   

5.
Under iron limitation, bacteria scavenge ferric (Fe3+) iron bound to siderophores or other chelates from the environment to fulfill their nutritional requirement. In gram‐negative bacteria, the siderophore uptake system prototype consists of an outer membrane transporter, a periplasmic binding protein and a cytoplasmic membrane transporter, each specific for a single ferric siderophore or siderophore family. Here, we show that spontaneous single gain‐of‐function missense mutations in outer membrane transporter genes of Bradyrhizobium japonicum were sufficient to confer on cells the ability to use synthetic or natural iron siderophores, suggesting that selectivity is limited primarily to the outer membrane and can be readily modified. Moreover, growth on natural or synthetic chelators required the cytoplasmic membrane ferrous (Fe2+) iron transporter FeoB, suggesting that iron is both dissociated from the chelate and reduced to the ferrous form within the periplasm prior to cytoplasmic entry. The data suggest rapid adaptation to environmental iron by facile mutation of selective outer membrane transporter genes and by non‐selective uptake components that do not require mutation to accommodate new iron sources.  相似文献   

6.
Kinetics of radioactive iron transport were examined in three strains of Bacillus megaterium. In strain ATCC 19213, which secretes the ferric-chelating secondary hydroxamic acid schizokinen, 59Fe3+ uptake from 59FeCl3 or the ferric hydroxamate Desferal-59Fe3+ was rapid and reached saturation within 3 min. In strain SK11, which does not secrete schizokinen, transport from 59FeCl3 was markedly reduced; the two ferric hydroxamates Desferal-59Fe3+ or schizokinen-59Fe3+ increased both total 59Fe3+ uptake and the 59Fe3+ appearing in a cellular trichloroacetic acid-insoluble fraction, although 10 min was required to reach saturation. Certain characteristics of transport from both ferric hydroxamates and FeCl3 suggest that iron uptake was an active process. The growth-inhibitory effect of aluminum on strain SK11 was probably due to the formation of nonutilizable iron-aluminum complexes which blocked uptake from 59FeCl3. Desferal or schizokinen prevented this blockage. A strain (ARD-1) resistant to the ferric hydroxamate antibiotic A22765 was isolated from strain SK11. Strain ARD-1 failed to grow with Desferal-Fe3+ as an iron source, and it was unable to incorporate 59Fe3+ from this source. Growth and iron uptake in strain ARD-1 were similar to strain SK11 with schizokinen-Fe3+ or the iron salt as sources. It is suggested that the ferric hydroxamates, or the iron they chelate, may be transported by a special system which might be selective for certain ferric hydroxamates. Strain ARD-1 may be unable to recognize both the antibiotic A22765 and the structurally similar chelate Desferal-Fe3+, while retaining its capacity to utilize schizokinen-Fe3+.  相似文献   

7.
8.
Siderophores are iron-chelating molecules that solubilize Fe3+ for microbial utilization and facilitate colonization or infection of eukaryotes by liberating host iron for bacterial uptake. By fluorescently labeling membrane receptors and binding proteins, we created 20 sensors that detect, discriminate, and quantify apo- and ferric siderophores. The sensor proteins originated from TonB-dependent ligand-gated porins (LGPs) of Escherichia coli (Fiu, FepA, Cir, FhuA, IutA, BtuB), Klebsiella pneumoniae (IroN, FepA, FyuA), Acinetobacter baumannii (PiuA, FepA, PirA, BauA), Pseudomonas aeruginosa (FepA, FpvA), and Caulobacter crescentus (HutA) from a periplasmic E. coli binding protein (FepB) and from a human serum binding protein (siderocalin). They detected ferric catecholates (enterobactin, degraded enterobactin, glucosylated enterobactin, dihydroxybenzoate, dihydroxybenzoyl serine, cefidericol, MB-1), ferric hydroxamates (ferrichromes, aerobactin), mixed iron complexes (yersiniabactin, acinetobactin, pyoverdine), and porphyrins (hemin, vitamin B12). The sensors defined the specificities and corresponding affinities of the LGPs and binding proteins and monitored ferric siderophore and porphyrin transport by microbial pathogens. We also quantified, for the first time, broad recognition of diverse ferric complexes by some LGPs, as well as monospecificity for a single metal chelate by others. In addition to their primary ferric siderophore ligands, most LGPs bound the corresponding aposiderophore with ∼100-fold lower affinity. These sensors provide insights into ferric siderophore biosynthesis and uptake pathways in free-living, commensal, and pathogenic Gram-negative bacteria.  相似文献   

9.
Iron uptake from two Fe3+-hydroxamate siderophores, ferrioxamine B and Fe3+-rhodotorulate, by iron-stressed Chlorella vulgaris (ATCC strain 11468) was evaluated with some comparison to iron uptake from synthetic and organic acid ferric chelates. Iron-stress induced iron uptake from ferrioxamine B. Dissipation of the electrochemical gradient, via uncouplers, inhibited iron uptake. Respiratory inhibitors gave variable results, an indication that a direct link to respiration was not apparent. Vanadate inhibition of iron uptake indicated that an ATPase or phosphate intermediate could be involved in the uptake mechanism. Divalent cations manifested variable effects dependent on the cation and chelator used. These data confirm that C. vulgaris has an inducible iron-uptake system for Fe3+-hydroxamic acid siderophores which may involve a different mechanism than that observed for other chelates.  相似文献   

10.
Iron has a central role in bioleaching and biooxidation processes. Fe2+ produced in the dissolution of sulfidic minerals is re-oxidized to Fe3+ mostly by biological action in acid bioleaching processes. To control the concentration of iron in solution, it is important to precipitate the excess as part of the process circuit. In this study, a bioprocess was developed based on a fluidized-bed reactor (FBR) for Fe2+ oxidation coupled with a gravity settler for precipitative removal of ferric iron. Biological iron oxidation and partial removal of iron by precipitation from a barren heap leaching solution was optimized in relation to the performance and retention time (τFBR) of the FBR. The biofilm in the FBR was dominated by Leptospirillum ferriphilum and “Ferromicrobium acidiphilum.” The FBR was operated at pH 2.0 ± 0.2 and at 37 °C. The feed was a barren leach solution following metal recovery, with all iron in the ferrous form. 98–99% of the Fe2+ in the barren heap leaching solution was oxidized in the FBR at loading rates below 10 g Fe2+/L h (τFBR of 1 h). The optimal performance with the oxidation rate of 8.2 g Fe2+/L h was achieved at τFBR of 1 h. Below the τFBR of 1 h the oxygen mass transfer from air to liquid limited the iron oxidation rate. The precipitation of ferric iron ranged from 5% to 40%. The concurrent Fe2+ oxidation and partial precipitative iron removal was maximized at τFBR of 1.5 h, with Fe2+ oxidation rate of 5.1 g Fe2+/L h and Fe3+ precipitation rate of 25 mg Fe3+/L h, which corresponded to 37% iron removal. The precipitates had good settling properties as indicated by the sludge volume indices of 3–15 mL/g but this step needs additional characterization of the properties of the solids and optimization to maximize the precipitation and to manage sludge disposal.  相似文献   

11.
Iron acquisition from various ferric chelates and colloids was studied using iron‐limited cells of Anabaena flos‐aquae (Lyng.) Brèb UTEX 1444, a cyanobacterial strain that produces high levels of siderophores under iron limitation. Various chelators of greatly varying affinity for Fe3+ (HEDTA, EDDHA, desferrioxamine mesylate, HBED, 8‐hydroxyquinoline) were assayed for the degree of iron acquisition by iron‐limited cyanobacterial cells. Iron uptake rates (measured by graphite furnace atomic absorption spectrometry) varied approximately inversely with calculated [Fe3+] (calculated as pFe) and decreased with increasing chelator‐to‐iron ratio. No iron uptake was observed when Fe3+ was chelated with HBED, the strongest of the tested chelators. Iron‐limited Anabaena cells were able to take up iron from 8‐hydroxyquinoline (oxine or 8HQ), a compound sometimes used to quantify aquatic iron bioavailability. Iron bound to purified humic acid was poorly available but did support some growth at high humic acid concentrations. These results suggest that for cyanobacteria, even tightly bound iron is biologically available, including to a limited extent iron bound to humic acids. However, iron bound to some extremely strong chelators (e.g. HBED) is likely to be biologically unavailable.  相似文献   

12.
Zusammenfassung Es wird die Eiablage weiselloser vonLeptothorax tuberum unifasciatus und einiger anderer Arten untersucht. Die weisellosen legen mehr Eier als die Königin in einer gleichstarken-Gruppe. Entweiselte brauchen einen gewissen Zeitraum (Inkubationszeit), bis sie fertil werden. Aus diesen Befunden geht hervor, dass die in natürlich zusammengesetzten Völkern steril sind. Der -Duft allein kann die-Fertilitä nicht hemmen. Von der abgetrennte legen keine Eier, wenn ihnen täglich aus der Umgebung der zugesetzt werden. Zur Erklärung dieser Vorgänge wird die Hypothese der profertilen Stoffe entwickelt und dem -Duft eine wesentliche Rolle als Auslöser für ihre Weitergabe zugeschrieben.Herrn Prof. Dr.K. Gösswald danke ich für Anregung und Förderung der Untersuchungen.  相似文献   

13.
14.
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   

15.
l-Lactate-driven ferric and nitrate reduction was studied in Escherichia coli E4. Ferric iron reduction activity in E. coli E4 was found to be constitutive. Contrary to nitrate, ferric iron could not be used as electron acceptor for growth. Ferric iron reductase activity of 9 nmol Fe2+ mg-1 protein min-1 could not be inhibited by inhibitors for the respiratory chain, like Rotenone, quinacrine, Actinomycin A, or potassium cyanide. Active cells and l-lactate-driven nitrate respiration in E. coli E4 leading to the production of nitrite, was reduced to about 20% of its maximum activity with 5 mM ferric iron, or to about 50% in presence of 5 mM ferrous iron. The inhibition was caused by nitric oxide formed by a purely chemical reduction of nitrite by ferrous iron. Nitric oxide was further chemically reduced by ferrous iron to nitrous oxide. With electron paramagnetic resonance spectroscopy, the presence of a free [Fe2+-NO] complex was shown. In presence of ferrous or ferric iron and l-lactate, nitrate was anaerobically converted to nitric oxide and nitrous oxide by the combined action of E. coli E4 and chemical reduction reactions (chemodenitrification).  相似文献   

16.
The oxidation of ferrous iron (Fe2+) to ferric iron (Fe3+) with dioxygen (O2) by various strains of Thiobacillus ferrooxidans was studied by measuring the rate of O2 consumption at various Fe2+ concentrations and cell concentrations. The apparent Km values for Fe2+ remained constant at different cell concentrations of laboratory strains ATCC 13661 and ATCC 19859 but increased with increasing cell concentrations of mine isolates SM-4 and SM-5. The latter results are explained by the competitive inhibition of the Fe2+-binding site of a cell by other cells in the reaction mixture. Possible mechanisms involving cell surface properties are discussed.  相似文献   

17.
Abstract

Listeria monocytogenes, the causative agent of listeriosis, is a virulent foodborne Gram-positive bacterial pathogen, with 20–30% mortality. It has a broad ability to transport iron, either in the form of ferric siderophores, or by extracting it from mammalian iron binding proteins. In this review we focus on the mechanisms of ferric siderophore and haem transport into the listerial cell. Despite the fact that it does not synthesize siderophores, L. monocytogenes transports ferric siderophores in the wild environment by the actions of cytoplasmic membrane ABC-transporter systems. The bacterium acquires haem, on the other hand, by two mechanisms. At low (nanomolar) concentrations, sortase B-dependent, peptidoglycan-anchored proteins scavenge the iron porphyrin in human or animal tissues, and transfer it to the underlying ABC-transporters in the cytoplasmic membrane for uptake. At concentrations at or above 50 nM, however, haem transport becomes sortase-independent, and occurs by direct interactions of the iron porphyrin with the same ABC-transporter complexes. The architecture of the Gram-positive cell envelope plays a fundamental role in these mechanisms, and the haem acquisition abilities of L. monocytogenes are an element of its ability to cause infectious disease.  相似文献   

18.
The properties of NADH-dependent Fe3+-EDTA reductase in plasma membranes (PM) from roots of iron-deficient and -sufficient tomato plants [Lycopersicon esculentum L. (Mill.) cv. Abunda] were examined. Iron deficiency resulted in a 3-fold increase of in vivo root iron-chelate reductase activity with a Km (Fe3+-EDTA) of 230 μM. In purified root PM, average specific activities of ferric chelate reductase of 410 and 254 nmol Fe (mg protein)?1 min?1 were obtained for iron-deficient and -sufficient plants, respectively. In both cases, the PM-bound activity showed a pH optimum at pH 6.8. Activity depended on NADH and not on NADPH and on the presence of detergent. The activity was inhibited 40-50% by superoxide dismutase (EC 1.15.1.1) and ca 30% by oxygen. Kinetic analysis of the membrane-bound enzyme revealed a Km (Fe3+-EDTA) of ca 200 μM for both iron-stressed and -sufficient plants. For NADH, Km values around 230 μM were obtained. The ferric chelate reductase could be solubilised from salt-washed PM with Triton X-100 at a protein:detergent ratio of 1:2.8 (w/w). The Triton-soluble fraction revealed one enzyme-stained band in native polyacrylamide electrophoresis. Although the membranes showed no nitrate reductase (NR; EC 1.6.6.1) activity, anti-spinach NR immunoglobulin G (IgG) recognized a 54 kDa band both in the PM and the Triton-soluble fraction, but not in the enzymatically active material obtained from the native gel. No evidence could be found for the synthesis of a new, biochemically distinct PM-bound ferric chelate reductase under iron deficiency, which might be identified as the so-called Turbo reductase. It is concluded that iron deficiency in tomato induces increased expression of a ferric chelate reductase in root PM, which is already present in iron-sufficient plants and probably also in plants, which do not contain the Turbo reductase, like the grasses. The iron reductase is not identical with the recently reported PM-associated nitrate reductase.  相似文献   

19.
Previous data indicated that diquat-mediated protein oxidation (protein carbonyl formation) occurs through multiple pathways, one of which is lipid dependent, and the other, lipid independent. Studies reported here investigated potential mechanisms of the lipid-independent pathway in greater detail, using bovine serum albumin as the target protein. One hypothesized mechanism of protein carbonyl formation involved diquat-dependent production of H2O2, which would then react with site-specifically bound ferrous iron as proposed by Stadtman and colleagues. This hypothesis was supported by the inhibitory effect of catalase on diquat-mediated protein carbonyl formation. However, exogenous H2O2 alone did not induce protein carbonyl formation. Hydroxyl radical-generating reactions may result from the H2O2-catalyzed oxidation of ferrous iron, which normally is bound to protein in the ferric state. Therefore, the possible reduction of site-specifically bound Fe3+ to Fe2+ by the diquat cation radical (which could then react with H2O2) was also investigated. The combination of H2O2 and an iron reductant, ascorbate, however, also failed to induce significant protein carbonyl formation. In a phospholipid-containing system, an ADP:Fe2+ complex induced both lipid peroxidation and protein carbonyl formation; both indices were largely inhibitable by antioxidants. There was no substantial ADP:Fe2+-dependent protein carbonyl formation in the absence of phospholipid under otherwise identical conditions. Based on the lipid requirement and antioxidant sensitivity, these data suggest that ADP:Fe2+-dependent protein carbonyl formation occurs through reaction of BSA with aldehydic lipid peroxidation products. The precise mechanism of diquat-mediated protein carbonyl formation remains unclear, but it appears not to be a function of H2O2 generation or diquat cation radical-dependent reduction of bound Fe3+. © 1998 John Wiley & Sons, Inc. J Biochem Toxicol 12: 185–190, 1998  相似文献   

20.
孟令博  赵曼  亢燕  祁智 《西北植物学报》2021,41(10):1681-1690
以羊草幼苗为研究对象,通过调整全营养培养基(CK,0.05 mmol/L Fe2+、0.015 mmol/L Zn2+)中铁或者锌含量设置0、10倍、20倍Fe2+(Zn2+)浓度处理Fe0(Zn0)、Fe10(Zn10)、Fe20(Zn20),以及在高铁培养基中单独添加0.15 mmol/L Zn2+或同时添加10 mmol/L Ca2+、5 mmol/L Mg2+、20 mmol/L K+处理,测定培养6 d后幼苗生长指标和矿质元素含量、以及高铁(Fe20)处理下幼苗根中抗氧化指标和相关基因表达量,探究不同浓度Fe2+、Zn2+对羊草幼苗生长、矿质元素吸收积累及抗氧化指标、基因表达的影响。结果表明:(1)缺锌(Zn0)显著抑制羊草幼苗鲜重的增加和Zn元素的积累,但促进Fe、Mg元素的积累;高浓度锌(Zn10、Zn20)显著促进幼苗叶片生长和Zn元素的积累;缺铁(Fe0)显著抑制幼苗的根长、鲜重和Fe元素的积累,促进Mg、Zn元素的积累;高浓度铁(Fe10、Fe20)显著抑制羊草幼苗根叶生长、根毛发育和Ca、Zn、Mg、K元素的积累。(2)增加Zn2+和Ca2+、Mg2+、K+浓度无法恢复高铁胁迫对幼苗生长的抑制作用。(3)高浓度铁(Fe20)处理羊草幼苗48 h后,根部过氧化物酶、超氧化物歧化酶、过氧化氢酶、抗坏血酸过氧化物酶、谷胱甘肽还原酶活性和丙二醛、抗坏血酸、还原型谷胱甘肽含量显著升高;烟酰胺合成酶基因、过氧化物酶基因表达量显著下调,植物类萌发素蛋白基因表达量显著上调。研究发现,羊草幼苗生长发育和矿质元素积累对环境中Zn2+浓度变化不敏感,却受到环境中高浓度Fe2+的显著抑制,并造成严重的氧化胁迫伤害,这种伤害无法在添加Zn2+或同时添加Ca2+、Mg2+、K+的条件下恢复。  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