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.
Reduction of Fe(III) to Fe(II) by Fe(III) chelate reductase is thought to be an obligatory step in iron uptake as well as the primary factor in making iron available for absorption by all plants except grasses. Fe(III) chelate reductase has also been suggested to play a more general role in the regulation of cation absorption. In order to experimentally address the importance of Fe(III) chelate reductase activity in the mineral nutrition of plants, three Arabidopsis thaliana mutants (frd1-1, frd1-2 and frd1-3), that do not show induction of Fe(III) chelate reductase activity under iron-deficient growth conditions, have been isolated and characterized. These mutants are still capable of acidifying the rhizosphere under iron-deficiency and accumulate more Zn and Mn in their shoots relative to wild-type plants regardless of iron status. frd1 mutants do not translocate radiolabeled iron to the shoots when roots are presented with a tightly chelated form of Fe(III). These results: (1) confirm that iron must be reduced before it can be transported, (2) show that Fe(III) reduction can be uncoupled from proton release, the other major iron-deficiency response, and (3) demonstrate that Fe(III) chelate reductase activity per se is not necessarily responsible for accumulation of cations previously observed in pea and tomato mutants with constitutively high levels of Fe(III) chelate reductase activity.  相似文献   

2.
In Bacillus subtilis, uracil (Ura), uridine (Urd), and deoxyuridine (dUrd) are metabolized through pathways similar to those of enteric bacteria. Ura is probably converted to uridine 5'-monophosphate by uridine 5'-monophosphate pyrophosphorylase. More than 95% of dUrd added to cultures is converted to Ura and deoxyribose-1-phosphate. Although dUrd kinase activity is detectable in vitro, this enzyme does not seem to play an important role in the metabolism of dUrd. The metabolism of cytosine (Cyt), cytidine (Cyd), and deoxycytidine (dCyd) in B. subtilis appears to be different from that in enteric bacteria. Cytosine cannot be used by Ura-requiring mutants as pyrimidine source. dCyd is deaminated by dCyd-Cyd deaminase or phosphorylated to dCyd nucleotides by dCyd kinase. Cyd is deaminated by dCyd-Cyd deaminase of phosphorylated by Cyd kinase. This Cyd kinase activity has never been reported for B. subtilis.  相似文献   

3.
In a speG-disrupted Escherichia coli mutant, which cannot metabolize spermidine to acetylspermidine, addition of spermidine to the medium caused a decrease in cell viability at the late stationary phase of growth. There were parallel decreases in the levels of ribosome modulation factor (RMF), the sigma(38) subunit of RNA polymerase, and the outer membrane protein C (OmpC). To clarify that these three proteins are strongly involved in cell viability, the rmf, rpoS (encoding sigma(38)), and ompC genes were disrupted. Viability of the triple mutant decreased to less than 1% of normal cells. The triple mutant had a reduced cell viability compared to any combination of double mutants, which also had a reduced cell viability. The single rmf and rpoS, but not ompC, mutant only slightly reduced cell viability. The results indicate that cooperative functions of these three proteins are necessary for cell viability at the late stationary phase. The triple mutant had a reduced level of ribosomes and of intracellular cations.  相似文献   

4.
AimsThe aim of this study was to investigate the effect of iron or/and zinc supplementation and termination of this treatment on the antioxidant defence of the male reproductive system and sperm viability in rats.MethodsThe study consisted of 3 stages: I) 4-week adaptation to the diets (C-control or D-iron deficient); II) 4-week iron and/or zinc supplementation (10-times more than in the C diet of iron: CSFe, DSFe; zinc: CSZn, DSZn; or iron and zinc: CSFeZn, DSFeZn; and III) 2-week post-supplementation period (the same diets as during stage I). Parameters of antioxidant status (total antioxidant capacity and SOD, GPx, and CAT activiy), oxidative damage (lipid and protein peroxidation), and sperm viability were measured.ResultsSimultaneous iron and zinc supplementation compared to iron supplementation (CSFeZn vs CSFe) increased SOD activity in the testes and decreased the level of malondialdehyde in the epididymis after stage II, and increased the percentage of live sperm after stage III. After discontinuation of the iron and zinc supplementation and a return to the control diet, the following was observed a decrease of SOD activity in the testes and GPx activity in the epididymis, and a increase malondialdehyde concentration in prostates. After stage III, in DSFeZn vs DSFe rats, an increase of SOD and CAT activity in the epididymis was found.ConclusionZinc supplementation simultaneous with iron may protect the male reproductive system against oxidative damage induced by high doses of iron and may have a beneficial effect on sperm viability. The effect of this supplementation was observed even two weeks after the termination of the intervention.  相似文献   

5.
The kinetics and mechanism of Fe(III) reduction to Fe(II) were studied in pure batch cultures of Pseudomonas sp. 200. The rate of iron reduction has been mechanistically related to aqueous phase iron speciation. In the absence of microbial activity the iron reduction rate was negligible. Initial rates of microbial iron reduction were accelerated more than 20-fold by the addition of equimolar quantities of nitrilotriacetic acid (NTA) to media initially containing 1.86 x 10(-3)M total Fe(III). Numerical techniques were utilized to quantify relationships between the observed rate of Fe(II) production and the calculated (equilibrium) aqueous phase speciation. These results indicate that soluble ferric iron species are not equivalent in terms of their susceptibility to bacterial (dissimilative) iron reduction. The concentration of Fe(NTA)(OH)(2) (2-) correlated strongly with observed iron reduction rates. Ferrous iron species appeared to inhibit the reduction process.  相似文献   

6.
Cucumber plants grown in hydroponics containing 10 μM Cd(II), Ni(II) and Pb(II), and iron supplied as Fe(III) EDTA or Fe(III) citrate in identical concentrations, were investigated by total-reflection X-ray fluorescence spectrometry with special emphasis on the determination of iron accumulation and distribution within the different plant compartments (root, stem, cotyledon and leaves). The extent of Cd, Ni and Pb accumulation and distribution were also determined. Generally, iron and heavy-metal contaminant accumulation was higher when Fe(III) citrate was used. The accumulation of nickel and lead was higher by about 20% and 100%, respectively, if the iron supply was Fe(III) citrate. The accumulation of Cd was similar. In the case of Fe(III) citrate, the total amounts of Fe taken up were similar in the control and heavy-metal-treated plants (27-31 μmol/plant). Further, the amounts of iron transported from the root towards the shoot of the control, lead- and nickel-contaminated plants were independent of the iron(III) form. Although Fe mobility could be characterized as being low, its distribution within the shoot was not significantly affected by the heavy metals investigated.  相似文献   

7.
Stenchuk NN  Kapustiak KE 《Genetika》2003,39(8):1026-1032
A method of positive selection of mutants with impaired regulation of flavinogenesis and metal homeostasis in yeast Pichia guilliermondii was developed. This positive selection system was based on the isolation of pseudo-wild-type revertants (the Rib+ phenotype) in riboflavin-dependent rib1-86 mutant (the Rib- phenotype) of yeast P. guilliermondii. Mutation rib1-86 blocks activity of the GTP cyclohydrolase II catalyzing the first step in riboflavin (RF) biosynthesis. Study of a collection of spontaneous Rib+ revertants allowed the identification of a considerably large number of genetic loci responsible for the suppression of rib1-86, which include both previously identified three loci (rib80, rib81, and hit1) and six new loci designated red1-red6 (reduction). A comparative analysis of the wild-type strain and red mutants revealed that these mutants had higher activity levels of GTP cyclohydrolase and RF-synthase, elevated levels of RF biosynthesis, enhanced Fe/Cu reductase activity and higher total iron content in cells and that they are characterized by enhanced sensitivity to transition metals (Fe(III), Cu(II), Cd(II), Co(II), Zn(II), Ag(I), and to H2O2. The metal hypersensitivity of mutant cells can be prevented by an increased amount of extracellular iron ions. Mutations red1 and red6 synergistically interact with the locus rib81 in the course of RF biosynthesis. Obviously, each RED gene plays an important role in the regulation of both flavinogenesis and metal homeostasis in P. guilliermondii cells.  相似文献   

8.
The Escherichia coli mutant (ppk) lacking the enzyme polyphosphate kinase, which makes long chains of inorganic polyphosphate (poly P), is deficient in functions expressed in the stationary phase of growth. After 2 days of growth in a medium limited in carbon sources, only 7% of the mutants survived compared with nearly 100% of the wild type; the loss in viability of the mutant was even more pronounced in a rich medium. The mutant showed a greater sensitivity to heat, to an oxidant (H2O2), to a redox-cycling agent (menadione), and to an osmotic challenge with 2.5 M NaCl. After a week or so in the stationary phase, mutant survivors were far fewer in number and were replaced by an outgrowth of a small-colony-size variant with a stable genotype and with improved viability and resistance to heat and H2O2; neither polyphosphate kinase nor long-chain poly P was restored. Suppression of the ppk feature of heat sensitivity by extra copies of rpoS, the gene encoding the RNA polymerase sigma factor that regulates some 50 stationary-phase genes, further implicates poly P in promoting survival in the stationary phase.  相似文献   

9.
Role of phosphate in initial iron deposition in apoferritin   总被引:1,自引:0,他引:1  
Y G Cheng  N D Chasteen 《Biochemistry》1991,30(11):2947-2953
Ferritins from microorganisms to man are known to contain varying amounts of phosphate which has a pronounced effect on the structural and magnetic properties of their iron mineral cores. The present study was undertaken to gain insight into the role of phosphate in the early stages of iron accumulation by ferritin. The influence of phosphate on the initial deposition of iron in apoferritin (12 Fe/protein) was investigated by EPR, 57Fe M?ssbauer spectroscopy, and equilibrium dialysis. The results indicate that phosphate has a significant influence on iron deposition. The presence of 1 mM phosphate during reconstitution of ferritin from apoferritin, Fe(II), and O2 accelerates the rate of oxidation of the iron 2-fold at pH 7.5. In the presence or absence of phosphate, the rate of oxidation at 0 degrees C follows simple first-order kinetics with respect to Fe(II) with half-lives of 1.5 +/- 0.3 or 2.8 +/- 0.2 min, respectively, consistent with a single pathway for iron oxidation when low levels of iron are added to the apoprotein. This pathway may involve a protein ferroxidase site where phosphate may bind iron(II), shifting its redox potential to a more negative value and thus facilitating its oxidation. Following oxidation, an intermediate mononuclear Fe(III)-protein complex is formed which exhibits a transient EPR signal at g' = 4.3. Phosphate accelerates the rate of decay of the signal by a factor of 3-4, producing EPR-silent oligonuclear or polynuclear Fe(III) clusters. In 0.5 mM Pi, the signal decays according to a single phase first-order process with a half-life near 1 min.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
P. McGraw  S. A. Henry 《Genetics》1989,122(2):317-330
We report the isolation of two new opi3 mutants by EMS mutagenesis, and construction of an insertion allele in vitro using the cloned gene. We have demonstrated that the opi3 mutations cause a deficiency in the two terminal phospholipid N-methyltransferase (PLMT) activities required for the de novo synthesis of PC (phosphatidylcholine). The opi3 mutants, under certain growth conditions, produce membrane virtually devoid of PC although, surprisingly, none of the mutants displays a strict auxotrophic requirement for choline. Although the opi3 mutants grow without supplements, we have shown that the atypical membrane affects the ability of the mutant strains to initiate log phase growth and to sustain viability at stationary phase. The commencement of log phase growth is enhanced by addition of choline or to a lesser extent DME (dimethylethanolamine), and retarded by addition of MME (monomethylethanolamine). The mutant cells lose viability at the stationary phase of the cell cycle in the absence of DME or choline, and are also temperature sensitive for growth at 37 degrees especially in media containing MME. These growth defects have been correlated to the presence of specific phospholipids in the membrane. The opi3 growth defects are suppressed by an unusual mutation in the phospholipid methylation pathway that perturbs the N-methyltransferase (PEMT) activity immediately preceding the reactions affected by the opi3 lesion. We believe this mutation, cho2-S, alters the substrate specificity of the PEMT. A secondary effect of opi3 mutations is disruption of the cross pathway regulation of the synthesis of the PI (phosphatidylinositol) precursor inositol. Synthesis of inositol is controlled through regulation of the INO1 gene which encodes inositol-1-phosphate synthase. This highly regulated gene is expressed constitutively in opi3 mutants. We have used the opi3 strains to demonstrate that synthesis of either PC or PD (phosphatidyldimethylethanolamine) will restore normal regulation of the INO1 gene.  相似文献   

11.
Iron is required for the growth of Saccharomyces cerevisiae. High concentrations of iron, however, are toxic, forcing this yeast to tightly regulate its concentration of intracellular free iron. We demonstrate that S. cerevisiae accumulates iron through the combined action of a plasma membrane ferrireductase and an Fe(II) transporter. This transporter is highly selective for Fe(II). Several other transition metals did not inhibit iron uptake when these metals were present at a concentration 100-fold higher than the Km (0.15 microM) for iron transport. Pt(II) inhibited ferrireductase activity but not the ability of cells to transport iron that was chemically reduced to Fe(II). Incubation of cells in a synthetic iron-limited media resulted in the induction of both ferrireductase and Fe(II) transporter activities. In complex media, Fe(II) transport activity was regulated in response to media iron concentration, while the activity of the ferrireductase was not. When stationary phase cells were inoculated into fresh media, ferrireductase activity increased independent of the iron content of the media; in contrast, transporter activity varied inversely with iron levels. These results demonstrate that the ferrireductase and Fe(II) transporter are separately regulated and that iron accumulation may be limited by changes in either activity.  相似文献   

12.
Because it can undergo reversible changes in oxidation state, iron is an excellent biocatalyst but also a potentially deleterious metal. Iron-mediated toxicity has been ascribed to Fe(II), which reacts with oxygen to generate free radicals that damage macromolecules and cause cell death. However, we now report that Fe(III) exhibits microbicidal activity towards strains of Salmonella enterica, Escherichia coli and Klebsiella pneumoniae defective in the Fe(III)-responding PmrA/PmrB signal transduction system. Fe(III) bound to a pmrA Salmonella mutant more effectively than to the isogenic wild-type strain and exerted its microbicidal activity even under anaerobic conditions. Moreover, Fe(III) permeabilized the outer membrane of the pmrA mutant, rendering it susceptible to vancomycin, which is normally non-toxic to Gram-negative species. On the other hand, Fe(III) did not affect the viability of a mutant defective in Fur, the major regulator of cytosolic iron homeostasis, which is hypersensitive to Fe(II)-mediated toxicity. A functional pmrA gene was necessary for bacterial survival in soil. Our results indicate that Fe(III) exerts its microbicidal activity by a mechanism that is oxygen independent and different from that mediated by Fe(II).  相似文献   

13.
To acquire iron, many plant species reduce soil Fe(III) to Fe(II) by Fe(III)-chelate reductases embedded in the plasma membrane of root epidermal cells. The reduced product is then taken up by Fe(II) transporter proteins. These activities are induced under Fe deficiency. We describe here the FRO1 gene from pea (Pisum sativum), which encodes an Fe(III)-chelate reductase. Consistent with this proposed role, FRO1 shows similarity to other oxidoreductase proteins, and expression of FRO1 in yeast conferred increased Fe(III)-chelate reductase activity. Furthermore, FRO1 mRNA levels in plants correlated with Fe(III)-chelate reductase activity. Sites of FRO1 expression in roots, leaves, and nodules were determined. FRO1 mRNA was detected throughout the root, but was most abundant in the outer epidermal cells. Expression was detected in mesophyll cells in leaves. In root nodules, mRNA was detected in the infection zone and nitrogen-fixing region. These results indicate that FRO1 acts in root Fe uptake and they suggest a role in Fe distribution throughout the plant. Characterization of FRO1 has also provided new insights into the regulation of Fe uptake. FRO1 expression and reductase activity was detected only in Fe-deficient roots of Sparkle, whereas both were constitutive in brz and dgl, two mutants with incorrectly regulated Fe accumulation. In contrast, FRO1 expression was responsive to Fe status in shoots of all three plant lines. These results indicate differential regulation of FRO1 in roots and shoots, and improper FRO1 regulation in response to a shoot-derived signal of iron status in the roots of the brz and dgl mutants.  相似文献   

14.
15.
Three structurally diverse iron (III) sequestering compounds (siderophores) were isolated from the supernatants of early stationary phase iron-deficient cultures of vegetative mycelia of the cultivated mushroom, Agaricus bisporus (ATCC 36416). The compounds were purified as their ferric chelates to homogeneity by gel permeation, cation exchange, and low-pressure reversed phase C18 chromatographies, and characterized as trihydroxamic acids. The chelates were identified as ferrichrome, ferric fusarinine C, and an unusual compound, des (diserylglycyl) ferrirhodin (DDF) by HPTLC cochromatography and electrophoresis against authentic samples, hydrolysis and amino acid analysis, and FAB-MS and 1H NMR spectroscopy. The iron transport activities of the three compounds (and of some structurally similar exogenous compounds) in young mycelial cells were determined by time- and concentration-dependent kinetic assays and inhibition experiments (CN-, N3-) using 55Fe(3+)-labeled chelates. 55Iron (III) uptake mediated by all three compounds was found to be via high affinity, energy-dependent processes; transport effectiveness was in the order: ferrichrome > DDF > ferric fusarinine C. The relative uptake of iron by lambda-cis ferrichromes was: ferrichrome > ferrirhodin > ferrichrome A; transport activity by the delta-cis fusarinines was: ferric fusarinine C > tris cis-(and trans-) fusarinine iron (III) > ferric N1-triacetylfusarinine C.  相似文献   

16.
M D Mendenhall  C A Jones  S I Reed 《Cell》1987,50(6):927-935
A 40 kd polypeptide that coprecipitates with the CDC28 gene product in immune complexes is specifically phosphorylated by the CDC28 protein kinase. Using this reaction, we detect activity only in extracts from dividing G1 phase cells. Exit from G1 by entry into S phase or the preconjugatory state induced by mating pheromone correlates with loss of p40 phosphorylation activity. Inactive extracts from cdc28 mutants complement extracts from cells arrested in S or M phase, suggesting that non-G1 cells are deficient in an exchangeable activating factor. Stationary and pheromone-treated cultures are rich in this exchangeable factor, but possess an inactive kinase that is not activated by complementation. cAMP-deficient mutants resemble stationary cells.  相似文献   

17.
The contribution of transmembrane regions I, II, and III of the Rickettsia prowazekii ATP/ADP translocase to the structure of the putative water-filled ATP translocation channel was evaluated from the accessibility of hydrophilic, thiol-reactive, methanethiosulfonate reagents to a library of 68 independent cysteine-substitution mutants heterologously expressed in Escherichia coli. The MTS reagents used were MTSES (negatively charged) and MTSET and MTSEA (both positively charged). Mutants F036C, Y042C, and R046C (TM I), K066C and P072C (TM II), and F101C, F105C, F108C, Y113C, and P114C (TM III) had no assayable transport activity, indicating that cysteine substitution at these positions may not be tolerated. All three MTS reagents inhibit the transport of ATP in mutants of TM I (L039C, S043C, S047C, I048C) and TM II (S061C, S063C, T067C, I069C, V070C, A074C). Further, these residues appear to cluster along a single face of the transmembrane domain. Preexposure of MTS-reactive mutants S047C (TM I) and T067C (TM II) to high levels of ATP resulted in protection from MTS-mediated inhibition. This indicated that both TM I and TM II make major contributions to the structure of an aqueous ATP translocation pathway. Finally, on the basis of the lack of accessibility of charged MTS reagents to the thiol groups in mutants of TM III, it appears that TM III is not exposed to the ATP translocation channel. Cysteine substitution of residues constituting a highly conserved "phenylalanine face" in TM III resulted in ablation of ATP transport activity. Further, substituting these phenylalanine residues for either isoleucine or tyrosine also resulted in much lower transport activity, indicating that some property of phenylalanine at these positions that is not shared by cysteine, isoleucine, or tyrosine is critical to translocase activity.  相似文献   

18.
19.
The function of the reversible oxygen-binding hemoprotein from Vitreoscilla (VHb), which enhances oxygen-limited cell growth and recombinant protein production when functionally expressed in Escherichia coli, was investigated in wild-type E. coli and in E. coli mutants lacking one of the two terminal oxidases, cytochrome o complex (aerobic terminal oxidase, Cyo) or cytochrome d complex (microaerobic terminal oxidase, Cyd). Deconvolution of VHb, cytochrome o, and cytochrome d bands from in vivo absorption spectra revealed a 5-fold enhancement in cytochrome o content and a 1.5-fold increment in cytochrome d by VHb under microaerobic environments (dissolved oxygen less than 2% air saturation). Based upon oxygen uptake kinetics measurements of these mutants, the apparent oxygen affinity of the Cyo(+), Cyd(-) E. coli was increased in the presence of VHb, but no difference in the apparent K(m) was observed for the Cyo(-), Cyd(+) strain. Results suggest that the expression of VHb in E. coli increases the level and activity of terminal oxidases and thereby improves the efficiency of microaerobic respiration and growth.  相似文献   

20.
The shell of the bivalve Montacuta ferruginosa is covered with a rust-colored biofilm. This biofilm includes filamentous bacteria and protozoa encrusted with a mineral, rich in ferric ion and phosphate. The aim of this research was to study two possible microbial iron precipitation pathways in the biofilm, namely, microbial iron oxidation and microbial degradation of organic Fe(III) complexes. The iron-oxidizing activity was assayed spectrophotometrically by monitoring the formation of the dye Wurster blue in biofilm extracts. Iron-oxidizing activity was effectively detected in extracts obtained by oxalic acid treatment of biofilm fragments. Extracts obtained without oxalic acid treatment, heated extracts, or extracts supplemented with HgCl 2 did not show any activity. This suggests that an iron-oxidizing factor (IOF), possibly an enzyme, coprecipitated with the mineral. Additional information gathered by using sodium dodecyl sulfate-polyacrylamide gel electrophoresis, gel-filtration chromatography, and UV spectrophotometry indicate that the IOF would be a small peptide or glycopeptide (1,350 Da). Microbial degradation of organic Fe(III) complexes was assayed with biofilm fragments incubated in a medium containing ferric citrate. Analysis of the supernatants after various intervals revealed that the complex was degraded by living microorganisms much faster than in the heat-killed negative controls. We conclude that ferric iron precipitation in the biofilm may proceed by way of microbial Fe(II) oxidation as well as microbial degradation of organic Fe(III) complexes.  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