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 20 毫秒
1.
Manganese binding and oxidation by spores of a marine bacillus.   总被引:15,自引:3,他引:12       下载免费PDF全文
Mature, dormant spores of a marine bacillus, SG-1, bound and oxidized (precipitated) manganese on their surfaces. The binding and oxidation occurred under dormant conditions, with mature spores suspended in natural seawater. These heat-stable spores were formed in the absence of added manganese in the growth medium. The rate and amount of manganese bound by SG-1 spores was a function of spore concentration. Temperatures greater than 45 degrees C, pH values below 6.5, or the addition of EDTA or the metabolic inhibitors sodium azide, potassium cyanide, and mercuric chloride inhibited manganese binding and oxidation. However, SG-1 spores bound and oxidized manganese after treatment with glutaraldehyde, formaldehyde, ethylene oxide gas, or UV light, all of which killed the spores. Manganese oxidation never occurred in the absence of manganese binding to spores. The data suggest that Mn2+ was complexed by a spore component, perhaps an exosporium or a spore coat protein: once bound, the manganese was rapidly oxidized.  相似文献   

2.
Upon start-up of a rapid sand filter fed with groundwater containing Mn2+ and NH4+, the first to be removed was NH4+, which was oxidized to NO2 -. After both NH4+ and NO2 -. were completely oxidized to NO3 -, the removal of Mn2+ commenced. Batch experiments showed that the addition of Nitrosomonas europaea and Nitrobacter winogradskyi stimulated the Mn2+ removal by sandfilter microbial consortia. NO2 -. was found to have a marked inhibitory effect on the removal of Mn2+ and could reduce the removal rate by half. In this respect, NO2 --mediated chemical reduction of manganese oxide was demonstrated at slightly acidic pH values. In pure cultures of Nitrosomonas europaea and Nitrobacter winogradskyi, no Mn2+ oxidation occurred, but reduction of MnO2 to Mn2+ was found when NO2 -. accumulated. These results indicate that the development of NO2/-. oxidizers is critical in the removal of Mn2+ in rapid sand filters. By oxidizing NO2 -. NO2 -. oxidizers eliminate the negative effect of NO2 -. on the biological oxidation of Mn2+.  相似文献   

3.
The oxidation of Mn2+-pyrophosphate to Mn3+ by superoxide (O2?) was quantitative as evidenced from the formation of Mn3+-pyrophosphate and hydrogen peroxide and from the inhibition by superoxide dismutase. Using the competitive relation between Mn2+-pyrophosphate and superoxide dismutase for the O2?, the rate constant of Mn2+ oxidation was estimated to be about 6 × 106m?1 s?1. The oxidation of Mn2+-pyrophosphate by illuminated chloroplasts was also indicated to be stoichiometrically induced by O2?. In the presence of saturating amounts of the Mn2+, a double enhancement of hydrogen peroxide production and triple uptake of oxygen were found, as expected from the oxidation of Mn2+-pyrophosphate by O2?. Anaerobiosis or superoxide dismutase annuled these increments. We propose that the O2? generated as the sole initial step of the Mehler reaction oxidized Mn2+-pyrophosphate, and we discuss the role of free manganese in chloroplasts.  相似文献   

4.
Mn2+ exerted various effects on the growth of Leptothrix discophora strain SS-1 in batch cultures depending on the concentration added to the medium. Concentrations of 0.55 to 5.5 μM Mn2+, comparable to those in the environment from which strain SS-1 was isolated, decreased cell yield and prolonged stationary-phase survival, but did not affect growth rate. Elevated concentrations of 55 to 910 μM Mn2+ also decreased cell yield and prolonged survival, but growth rate was decreased as well. The addition of 1,820 μM Mn2+ caused a decline in cell numbers followed by an exponential rise after 80 h of incubation, indicating the development of a population of cells resistant to Mn2+ toxicity. When 360 μM Mn2+ or less was added to growth flasks, Mn2+ was oxidized to manganese oxide (MnOx, where x is ~2), which appeared as brown particles in the medium. Quantification of Mn oxidation during growth of cultures to which 55 μM Mn2+ was added showed that nearly all of the Mn2+ was oxidized by the beginning of the stationary phase of growth (15 to 25 h). This result suggested that the decrease in cell yield observed at low and moderate concentrations of Mn2+ was related to the formation of MnOx, which may have bound cationic nutrients essential to the growth of SS-1. The addition of excess Fe3+ to cultures containing 55 μM Mn2+ increased cell yield to levels near those found in cultures with no added Mn2+, indicating that iron deprivation by MnOx was at least partly responsible for the decreased cell yield.  相似文献   

5.
The interaction between saccharose and manganese in different oxidation states was studied in alkaline media by polarographic, potentiometric, ESR spectroscopic and UV-Vis spectrophotometric methods. The results showed that stable manganese(II) and manganese(III) complexes and a complex of manganese(II,III) in a mixed oxidation state were formed with the composition [MnIIL(OH)2], [Mn2IIIL2(OH)8]2− and [MnIIMnIIIL2(OH)6], respectively. The manganese(II)-saccharose complex was shown to dimerize in alkaline media. The stability constants of the Mn(II,III) and Mn(III) complexes were determined. The oxidation of the manganese(II)-saccharose complex by a stoichiometric amount of K3 [FeCN]6 resulted in the formation of the manganese(III) and manganese(IV) complexes. However, oxidation by molecular oxygen only yielded the manganese(III) complex which reduced spontaneously in inert atmosphere to the mixed valence Mn(II,III) complex. The latter was able to be oxidized again by oxygen to the Mn(III) complex. This process proved to be reversible and could be repeated several times.  相似文献   

6.
A new chemolithotrophic bacterial metabolism was discovered in anaerobic marine enrichment cultures. Cultures in defined medium with elemental sulfur (S0) and amorphous ferric hydroxide (FeOOH) as sole substrates showed intense formation of sulfate. Furthermore, precipitation of ferrous sulfide and pyrite was observed. The transformations were accompanied by growth of slightly curved, rod-shaped bacteria. The quantification of the products revealed that S0 was microbially disproportionated to sulfate and sulfide, as follows: 4S0 + 4H2O → SO42- + 3H2S + 2H+. Subsequent chemical reactions between the formed sulfide and the added FeOOH led to the observed precipitation of iron sulfides. Sulfate and iron sulfides were also produced when FeOOH was replaced by FeCO3. Further enrichment with manganese oxide, MnO2, instead of FeOOH yielded stable cultures which formed sulfate during concomitant reduction of MnO2 to Mn2+. Growth of small rod-shaped bacteria was observed. When incubated without MnO2, the culture did not grow but produced small amounts of SO42- and H2S at a ratio of 1:3, indicating again a disproportionation of S0. The observed microbial disproportionation of S0 only proceeds significantly in the presence of sulfide-scavenging agents such as iron and manganese compounds. The population density of bacteria capable of S0 disproportionation in the presence of FeOOH or MnO2 was high, > 104 cm-3 in coastal sediments. The metabolism offers an explanation for recent observations of anaerobic sulfide oxidation to sulfate in anoxic sediments.  相似文献   

7.
This review focuses on our recent efforts in synthetic ruthenium–tyrosine–manganese chemistry mimicking the donor side reactions of Photosystem II. Tyrosine and tryptophan residues were linked to ruthenium photosensitizers, which resulted in model complexes for proton-coupled electron transfer from amino acids. A new mechanistic model was proposed and used to design complexes in which the mechanism could be switched between concerted and step-wise proton-coupled electron transfer. Moreover, a manganese dimer linked to a ruthenium complex could be oxidized in three successive steps, from Mn2II,II to Mn2III,IV by the photo-oxidized ruthenium sensitizer. This was possible thanks to a charge compensating ligand exchange in the manganese complex. Detailed studies of the ligand exchange suggested that at high water concentrations, each oxidation step is coupled to a proton-release of water-derived ligands, analogous to the oxidation steps of the manganese cluster of Photosystem II.  相似文献   

8.
In this paper a few calcium-manganese oxides and calcium-manganese minerals are studied as catalysts for water oxidation. The natural mineral marokite is also studied as a catalyst for water oxidation for the first time. Marokite is made up of edge-sharing Mn3+ in a distorted octahedral environment and eight-coordinate Ca2+ centered polyhedral layers. The structure is similar to recent models of the oxygen evolving complex in photosystem II. Thus, the oxygen evolving complex in photosystem II does not have an unusual structure and could be synthesized hydrothermally. Also in this paper, oxygen evolution is studied with marokite (CaMn2O4), pyrolusite (MnO2) and compared with hollandite (Ba0.2Ca0.15K0.3Mn6.9Al0.2Si0.3O16), hausmannite (Mn3O4), Mn2O3.H2O, CaMn3O6.H2O, CaMn4O8.H2O, CaMn2O4.H2O and synthetic marokite (CaMn2O4). I propose that the origin of the oxygen evolving complex in photosystem II resulted from absorption of calcium and manganese ions that were precipitated together in the archean oceans by protocyanobacteria because of changing pH from ~5 to ~8-10. As reported in this paper, amorphous calcium-manganese oxides with different ratios of manganese and calcium are effective catalysts for water oxidation. The bond types and lengths of the calcium and manganese ions in the calcium-manganese oxides are directly comparable to those in the OEC. This primitive structure of these amorphous calcium-manganese compounds could be changed and modified by environmental groups (amino acids) to form the oxygen evolving complex in photosystem II.  相似文献   

9.
The strict anaerobe Clostridium difficile is the most common cause of antibiotic-associated diarrhoea. The oxygen-resistant C. difficile spores play a central role in the infectious cycle, contributing to transmission, infection and recurrence. The spore surface layers, the coat and exosporium, enable the spores to resist physical and chemical stress. However, little is known about the mechanisms of their assembly. In this study, we characterized a new spore protein, CotL, which is required for the assembly of the spore coat. The cotL gene was expressed in the mother cell compartment under the dual control of the RNA polymerase sigma factors, σE and σK. CotL was localized in the spore coat, and the spores of the cotL mutant had a major morphologic defect at the level of the coat/exosporium layers. Therefore, the mutant spores contained a reduced amount of several coat/exosporium proteins and a defect in their localization in sporulating cells. Finally, cotL mutant spores were more sensitive to lysozyme and were impaired in germination, a phenotype likely to be associated with the structurally altered coat. Collectively, these results strongly suggest that CotL is a morphogenetic protein essential for the assembly of the spore coat in C. difficile.  相似文献   

10.
It has been postulated that the heat stabilization of the essential macromolecules in the core of the spore may be produced by dehydration at two levels: (i) the spore is drier at high relative humidity than the vegetative cell and (ii) the core of the spore may be less hydrated than the cortex and the coat. The latter postulate was subjected to experimental testing by 1H-NMR studies of the water signal in the five species of spores and coat and (coat + cortex) preparations. The transverse relaxation rate (1T2) was determined in samples equilibrated at constant relative humidity. To allow for the effect of paramagnetic ions on 1T2 a model system (wool keratin) was studied in the presence of known amounts of Ca(II), Mn(II), Cu(II), Ni(II) and Fe(III). Because of the dominant effect of Mn(II) on 1T2, the effect of small amounts of other metal ions in spores was neglected. The relaxation rate of water at a particular relative humidity and manganese concentration was consistently less for intact spores than for coat or coat + cortex, hence the water in the core is more mobile than in the outer integuments. Sorption isotherm studies have shown that at a particular relative humidity there is about as much water in the core as in the cortex and coat. These two results taken together indicate that the hypothesis that the core is drier than the cortex and coat is incorrect, hence the spore is not heat-stabilized in this way. A theory is proposed in which heat stabilization is attributed to immobilization of essential enzymes and nuclei acids by a solid support, calcium dipicolinate, in a similar fashion to the heat stabilization of enzymes in a charged polymer matrix. It is proposed that stabilization is effected by electrostatic and hydrogen bonds between the calcium dipicolinate and the essential macromolecules. Experiments in progress show that enzymes and DNA are heat-stabilized in vitro by calcium dipicolinate.  相似文献   

11.
The growth capability of Trichoderma harzianum Rifaii Tl was tested on Malt Extract and Czapeks Dox agar containing different concentrations of Cu2+, Zn2+, Mn2+, Fe2+ and Ca2+. The T. harzianum Tl isolate was observed to produce mycelia and spores in various mineral-containing media. It showed the lowest tolerance to Ca2+ and the highest tolerance to Fe2+. Solubilization capability of T. harzianum Tl for some insoluble minerals via acidification of medium has been tested on MnO2, CuO, Fe2O3 and metallic Zn. T. harzianum Tl was able to solubilize MnO2 and metallic Zn in a liquid medium.  相似文献   

12.
Spores of Bacillus subtilis possess a thick protein coat that consists of an electron-dense outer coat layer and a lamellalike inner coat layer. The spore coat has been shown to confer resistance to lysozyme and other sporicidal substances. In this study, spore coat-defective mutants of B. subtilis (containing the gerE36 and/or cotE::cat mutation) were used to study the relative contributions of spore coat layers to spore resistance to hydrogen peroxide (H2O2) and various artificial and solar UV treatments. Spores of strains carrying mutations in gerE and/or cotE were very sensitive to lysozyme and to 5% H2O2, as were chemically decoated spores of the wild-type parental strain. Spores of all coat-defective strains were as resistant to 254-nm UV-C radiation as wild-type spores were. Spores possessing the gerE36 mutation were significantly more sensitive to artificial UV-B and solar UV radiation than wild-type spores were. In contrast, spores of strains possessing the cotE::cat mutation were significantly more resistant to all of the UV treatments used than wild-type spores were. Spores of strains carrying both the gerE36 and cotE::cat mutations behaved like gerE36 mutant spores. Our results indicate that the spore coat, particularly the inner coat layer, plays a role in spore resistance to environmentally relevant UV wavelengths.  相似文献   

13.
It was experimentally demonstrated that two strains of Arthrobacter 37, one growing at 25 C and the other at 5 C, could catalyze MnII oxidation at hydrostatic pressures well in excess of the pressure encountered by the parent culture in its original habitat in the ocean (80 atm). The strain grown at 5 C showed an increase in temperature optimum for manganese oxidation with increase in pressure. It was like-wise experimentally shown that induced Bacillus 29 without added ferricyanide and uninduced Bacillus 29 with added ferricyanide could catalyze MnO2 reduction at hydrostatic pressures in excess of the pressure encountered by this organism in its original habitat (187 atm). The uninduced Bacillus 29, in the presence of ferricyanide, was active over a wider range of pressures (1 to 1,000 atm) than the induced Bacillus 29 in the absence of ferricyanide (1 to 467 atm). At corresponding pressures, the uninduced culture was also considerably more active than the induced culture. Special techniques were developed for measuring MnII-oxidizing and MnO2-reducing activity under pressure.  相似文献   

14.
The geochemical cycling of cobalt (Co) has often been considered to be controlled by the scavenging and oxidation of Co(II) on the surface of manganese [Mn(III,IV)] oxides or manganates. Because Mn(II) oxidation in the environment is often catalyzed by bacteria, we have investigated the ability of Mn(II)-oxidizing bacteria to bind and oxidize Co(II) in the absence of Mn(II) to determine whether some Mn(II)-oxidizing bacteria also oxidize Co(II) independently of Mn oxidation. We used the marine Bacillus sp. strain SG-1, which produces mature spores that oxidize Mn(II), apparently due to a protein in their spore coats (R.A. Rosson and K. H. Nealson, J. Bacteriol. 151:1027-1034, 1982; J. P. M. de Vrind et al., Appl. Environ. Microbiol. 52:1096-1100, 1986). A method to measure Co(II) oxidation using radioactive 57Co as a tracer and treatments with nonradioactive (cold) Co(II) and ascorbate to discriminate bound Co from oxidized Co was developed. SG-1 spores were found to oxidize Co(II) over a wide range of pH, temperature, and Co(II) concentration. Leucoberbelin blue, a reagent that reacts with Mn(III,IV) oxides forming a blue color, was found to also react with Co(III) oxides and was used to verify the presence of oxidized Co in the absence of added Mn(II). Co(II) oxidation occurred optimally around pH 8 and between 55 and 65°C. SG-1 spores oxidized Co(II) at all Co(II) concentrations tested from the trace levels found in seawater to 100 mM. Co(II) oxidation was found to follow Michaelis-Menten kinetics. An Eadie-Hofstee plot of the data suggests that SG-1 spores have two oxidation systems, a high-affinity-low-rate system (Km, 3.3 × 10-8 M; Vmax, 1.7 × 10-15 M · spore-1 · h-1) and a low-affinity-high-rate system (Km, 5.2 × 10-6 M; Vmax, 8.9 × 10-15 M · spore-1 · h-1). SG-1 spores did not oxidize Co(II) in the absence of oxygen, also indicating that oxidation was not due to abiological Co(II) oxidation on the surface of preformed Mn(III,IV) oxides. These results suggest that some microorganisms may directly oxidize Co(II) and such biological activities may exert some control on the behavior of Co in nature. SG-1 spores may also have useful applications in metal removal, recovery, and immobilization processes.  相似文献   

15.
Killing of bacterial spores by H2O2 at elevated but sublethal temperatures and neutral pH occurred without lysis. However, with prolonged exposure or higher concentrations of the agent, secondary lytic processes caused major damage successively to the coat, cortex, and protoplast, as evidenced by electron and phase contrast microscopy. These processes were also reflected in changes in differential scanning calorimetric profiles for H2O2-treated spores. Endothermic transitions in the profiles occurred at lower temperatures than usual as a result of H2O2 damage. Thus, H2O2 sensitized the cells to heat damage. Longer exposure to H2O2 resulted in total disappearance of the transitions, indicative of major disruptions of cell structure. Spores but not vegetative cells were protected against the lethal action of H2O2 by the transition metal cations Cu+, Cu2+, Co2+, Co3+, Fe2+, Fe3+, Mn2+, Ti3+, and Ti4+. The metal chelator EDTA was also somewhat protective, while o-phenanthroline, citrate, deferoxamine, and ethanehydroxydiphosphonate were only marginally so. Superoxide dismutase and a variety of other free-radical scavengers were not protective. In contrast, reducing agents such as sulfhydryl compounds and ascorbate at concentrations of 20 to 50 mM were highly protective. Decoating or demineralization of the spores had only minor effects. The marked dependence of H2O2 sporicidal activity on moderately elevated temperature and the known low reactivity of H2O2 itself suggest that radicals are involved in its killing action. However, the protective effects of a variety of oxidized or reduced transition metal ions indicate that H2O2 killing of spores is markedly different from that of vegetative cells.  相似文献   

16.
Protein phosphatase 1γ, a serine/threonine phosphatase, is a metalloprotein that coordinates two Mn2+ in the active site when expressed in Escherichia coli in a buffer containing MnCl2. Herein, we report on the oxidatively induced copper for manganese exchange in protein phosphatase 1γ, thus enabling firm confirmation of the four histidine (His) amino acid residues (His66, His125, His173, and His248) involved in metal coordination. By exchanging manganese with copper the oxidation yields for the peptides increased dramatically, thus simplifying detection of the oxidized peptides and analysis of the oxidation sites within the oxidized peptides. We also found that when copper was added during the oxidation process a new metal coordination center was formed at cysteine 39, 105, 140, and 155.  相似文献   

17.
Spores are the infectious form of Bacillus anthracis (BA), causing cutaneous, inhalation and gastrointestinal anthrax. Because of the possible use of BA spores in a bioterrorism attack, there is considerable interest in studying spore biology. In the laboratory, however, it takes a number of days to prepare spores. Standard sporulation protocols, such as the use of ‘PA broth’, allow sporulation of BA to occur in 3 to 5 days. Another method employs growth of BA on plates in the dark for several days until they have efficiently sporulated. In efforts to determine the effect of iron on gene expression in BA, we grew BA Sterne strain 7702 in a minimal defined medium (CDM; Koppisch et al., 2005) with various concentrations of iron and glucose. As part of our initial observations, we monitored BA sporulation in CDM via light microscopy. In glucose-free CDM containing 1.5 mM Fe(NO3)3 (CDM-Fe), > 95% of the BA sporulated by 30 h; a far shorter time period than expected. We pursued this observation and we further characterized spores derived from PA and CDM-Fe media. Purified spores derived from PA or CDM-Fe had similar morphologies when viewed by light or electron microscopy, and were equally resistant to harsh conditions including heat (65 °C), ice and fresh 30% H2O2. Spore viability in long term cold storage in water was similar for the two spore preparations. Extracted spore coat proteins were evaluated by SDS-PAGE and silver staining, which revealed distinct protein profiles for PA and CDM-Fe spore coat extracts. ELISA assays were done to compare the interaction of the two spore preparations with rabbit antiserum raised against UV-killed Sterne strain 7702 spores prepared in PA medium. Spores from both media reacted identically with this antiserum. Finally, the interaction and fate of spores incubated with macrophages in vitro was very similar. In summary, BA spores induced in CDM-Fe or in PA medium are similar by several criteria, but show distinct extractable coat proteins. CDM-Fe liquid medium can be used for rapid production of BA spores, and could save considerable time in spore research studies.  相似文献   

18.
Isolated spore coats of a marine Bacillus species were incubated in 25 mM MnCl2 at pH 7.5. Manganese precipitates, formed on the coat surfaces, were analyzed by transmission electron microscopy, electron diffraction, and energy-dispersive X-ray spectroscopy. Initially, an amorphous manganese oxide was observed on the coats which recrystallized to hausmannite after prolonged incubation in the MnCl2 solution. The spore coats catalyze the oxidation of Mn(II) and have no structural influence on the final mineral phase precipitated.  相似文献   

19.
A study of the kinetics of Mn2+ oxidation catalyzed by cell extracts of two bacterial isolates (E1, Pseudomonas III [new isolate] and E4, Citrobacter freundii) isolated from the core of manganese concretions from Greek soils is presented. The reaction velocity of Mn2+ oxidation was determined from the rate of consumption of Mn2+. The oxidation of Mn2+ was followed by measuring changes in Mn2+ concentration by activation analysis and by atomic absorption spectrophotometry. The reaction velocity was directly proportional to cell extract concentration when the reaction time was 1 h. At longer reaction times, the relationship deviated from linearity because substrate concentration became limiting. The rate of Mn2+ oxidation increased with the Mn2+ concentration. Analysis of the results by application of the integrated Michaelis equation for determining Michaelis constants and maximal velocities either in the presence (Km = 3.33 μmol/ml and Vmax = 1.25 μmol/ml·h) or in the absence of maleate buffer (Km = 2.52 μmol/ml and Vmax = 2.04 μmol/ml·h) indicated a strong affinity between the oxidizing system and manganese. All results in this study are consistent with an enzymatic manganese-oxidizing system and give an indication of the mechanism of biological Mn2+ oxidation in soil which differs from that in the marine environment.  相似文献   

20.
An redox luminescence switch was developed for the sensing of glutathione (GSH), l ‐cysteine (Cys) or l ‐ascorbic acid (AA) based on redox reaction. The Mn2+‐doped NaYF4:Yb,Er upconversion nanorods (UCNRs) with an emission peak located in the red region were synthesized. The luminescence intensity of the UCNRs could be quenched due to the Mn2+ could be oxidized to MnO2 by KMnO4. Subsequently, when the AA, GSH or Cys was added into the MnO2 modified upconversion nanosystem, which could reduced MnO2 to Mn2+ and the luminescence intensity was recovered. The concentration ranges of the nanosystem are 0.500–3.375 mM (R2 = 0.99) for AA, 0.6250–11.88 mM (R2 = 0.99) for GSH and 0.6250–9.375 mM (R2 = 0.99) for Cys, respectively.  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