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 42 毫秒
1.
In order to assess the importance of nitrate-dependent Fe(II) oxidation and its impact on the growth physiology of dominant Fe oxidizers, we counted these bacteria in freshwater lake sediments and studied their growth physiology. Most probable number counts of nitrate-reducing Fe(II)-oxidizing bacteria in the sediment of Lake Constance, a freshwater lake in Southern Germany, yielded about 105 cells mL−1 of the total heterotrophic nitrate-reducing bacteria, with about 1% (103 cells mL−1) of nitrate-reducing Fe(II) oxidizers. We investigated the growth physiology of Acidovorax sp. strain BoFeN1, a dominant nitrate-reducing mixotrophic Fe(II) oxidizer isolated from this sediment. Strain BoFeN1 uses several organic compounds (but no sugars) as substrates for nitrate reduction. It also reduces nitrite, dinitrogen monoxide, and O2, but cannot reduce Fe(III). Growth experiments with cultures amended either with acetate plus Fe(II) or with acetate alone demonstrated that the simultaneous oxidation of Fe(II) and acetate enhanced growth yields with acetate alone (12.5 g dry mass mol−1 acetate) by about 1.4 g dry mass mol−1 Fe(II). Also, pure cultures of Pseudomonas stutzeri and Paracoccus denitrificans strains can oxidize Fe(II) with nitrate, whereas Pseudomonas fluorescens and Thiobacillus denitrificans strains did not. Our study demonstrates that nitrate-dependent Fe(II) oxidation contributes to the energy metabolism of these bacteria, and that nitrate-dependent Fe(II) oxidation can essentially contribute to anaerobic iron cycling.  相似文献   

2.
Understanding the mechanisms of anaerobic microbial iron cycling is necessary for a full appreciation of present‐day biogeochemical cycling of iron and carbon and for drawing conclusions about these cycles on the ancient Earth. Towards that end, we isolated and characterized an anaerobic nitrate‐dependent Fe(II)‐oxidizing bacterium from a freshwater sediment. The 16SrRNA gene sequence of the isolated bacterium (strain BoFeN1) places it within the β‐Proteobacteria, with Acidovorax sp. strain G8B1 as the closest known relative. During mixotrophic growth with acetate plus Fe(II) and nitrate as electron acceptor, strain BoFeN1 forms Fe(III) mineral crusts around the cells. The amount of the organic cosubstrate acetate present seems to control the rate and extent of Fe(II) oxidation and the viability of the cells. The crystallinity of the mineral products is influenced by nucleation by Fe minerals that are already present in the inoculum.  相似文献   

3.
Nitrate‐reducing, Fe(II)‐oxidizing bacteria were suggested to couple with enzymatic Fe(II) oxidation to nitrate reduction. Denitrification proceeds via intermediates (, NO) that can oxidize Fe(II) abiotically at neutral and particularly at acidic pH. Here, we present a revised Fe(II) quantification protocol preventing artifacts during acidic Fe extraction and evaluate the contribution of abiotic vs. enzymatic Fe(II) oxidation in cultures of the nitrate‐reducing, Fe(II) oxidizer Acidovorax sp. BoFeN1. Sulfamic acid used instead of HCl reacts with nitrite and prevents abiotic Fe(II) oxidation during Fe extraction. Abiotic experiments without sulfamic acid showed that acidification of oxic Fe(II) nitrite samples leads to 5.6‐fold more Fe(II) oxidation than in anoxic samples because the formed NO becomes rapidly reoxidized by O2, therefore leading to abiotic oxidation and underestimation of Fe(II). With our revised protocol using sulfamic acid, we quantified oxidation of approximately 7 mm of Fe(II) by BoFeN1 within 4 days. Without addition of sulfamic acid, the same oxidation was detected within only 2 days. Additionally, abiotic incubation of Fe(II) with nitrite in the presence of goethite as surface catalyst led to similar abiotic Fe(II) oxidation rates as observed in growing BoFeN1 cultures. BoFeN1 growth was observed on acetate with N2O as electron acceptor. When adding Fe(II), no Fe(II) oxidation was observed, suggesting that the absence of reactive N intermediates (, NO) precludes Fe(II) oxidation. The addition of ferrihydrite [Fe(OH)3] to acetate/nitrate BoFeN1 cultures led to growth stimulation equivalent to previously described effects on growth by adding Fe(II). This suggests that elevated iron concentrations might provide a nutritional effect rather than energy‐yielding Fe(II) oxidation. Our findings therefore suggest that although enzymatic Fe(II) oxidation by denitrifiers cannot be fully ruled out, its contribution to the observed Fe(II) oxidation in microbial cultures is probably lower than previously suggested and has to be questioned in general until the enzymatic machinery‐mediating Fe(II) oxidation is identified.  相似文献   

4.
Pathways for production of Fenton's reagent by wood-rotting fungi   总被引:2,自引:0,他引:2  
Abstract: Many forms of Fe(II) react with H202 to generate hydroxyl radicals (Fenton reaction). There is evidence that hydroxyl radicals are important in brown-rot, while they can be formed by secondary reactions during lignin breakdown by white-rot fungi. Their involvement in cellulose breakdown creates a range of oxidized sugars. The two reactants of Fenton's reagent can be generated by Fe(II) autoxidation, or by superoxide in reaction with Fe(III). A rapid autoxidation is not possible for complexes with a high Fe(III)/Fe(II) redox potential. Turning to specific pathways for formation of Fenton's reagent, decomposition of Fe(III)-oxalate is probably solely a photochemical process. Lignin peroxidases can act indirectly as a source of superoxide, either by reactions that lead to a peroxyradical, or by 1-electron oxidation of an aliphatic compound creating a strong reductant. Cellobiose dehydrogenase can provide a direct enzymic source for Fenton's reagent (S.M. Kremer and P.M. Wood (1992) Eur. J. Biochem. 208, 807–814). In the experiments as published, hydroxyl radical production was limited by the slow interaction of cellobiose dehydrogenase with O2. This limitation can be removed by the presence of an iron complex with an autoxidizable Fe(lI) state. The successful use of Fenton's reagent by a living organism requires a spatial separation between initiating enzyme(s) and the site of production of hydroxyl radicals. The mobility of the extra electron on Fe(II) by intermolecular transfer may be important for achieving this separation.  相似文献   

5.
【目的】探究不同菌浓度和亚铁浓度条件下,Acidovorax sp. strain BoFeN1介导的厌氧亚铁氧化耦合硝酸盐还原过程的动力学和次生矿物。【方法】构建包含菌BoFeN1、硝酸盐、亚铁的厌氧培养体系,测试硝酸根、亚硝酸根、乙酸根、亚铁等浓度,并收集次生矿物,采用XRD、SEM进行矿物种类和形貌表征。【结果】在微生物介导硝酸盐还原耦合亚铁氧化的体系中,高菌浓度促进硝酸盐还原,对亚铁氧化也有一定促进作用;高浓度亚铁在低菌浓度下氧化反应速率和程度降低,但是在高菌浓度下无明显影响;亚铁浓度越高次生矿物结晶度越高,但对硝酸盐还原具有一定抑制作用。在微生物介导亚硝酸盐还原耦合亚铁氧化的体系中,高的菌浓度和亚铁浓度都会促进亚硝酸盐还原,但亚铁氧化的次生矿物会对亚硝酸盐的微生物还原产生较强的抑制作用,次生矿物的种类和结晶度主要受亚铁浓度影响。【结论】硝酸盐还原主要是生物反硝化作用,亚硝酸盐还原包含生物反硝化和化学反硝化两部分,在硝酸盐体系中亚铁氧化与次生矿物生成是受生物和化学反硝化作用的共同影响,但亚硝酸盐体系中亚铁氧化与次生矿物生成主要是受化学反硝化作用影响。该研究可为深入理解厌氧微生物介导铁氮耦合反应机制提供基础数据和理论支撑。  相似文献   

6.
Toxic Cu (II) effect on cytochrome b 559 under aerobic photoinhibitory conditions was examined in two different photosystem II (PSII) membrane preparations active in oxygen evolution. The preparations differ in the content of cytochrome b 559 redox potential forms. Difference absorption spectra showed that the presence of Cu (II) induced the oxidation of the high-potential form of cytochrome b 559 in the dark. Addition of hydroquinone reduced the total oxidized high-potential form of cytochrome b 559 present in Cu (II)-treated PSII membranes indicating that no conversion to the low-potential form took place. Spectroscopic determinations of cytochrome b 559 during photoinhibitory treatment showed slower kinetics of Cu (II) effect on cytochrome b 559 in comparison with the rapid loss of oxygen evolution activity in the same conditions. This result indicates that cytochrome b 559 is affected after PSII centres are photoinhibited. The high-potential form was more sensitive to toxic Cu (II) action than the low-potential form under illumination at pH 6.0. The content of the high-potential form of cytochrome b 559 was completely lost; however, the low-potential content was unaffected in these conditions. This loss did not involve cytochrome protein degradation. The results are discussed in terms of different binding properties of the heme iron to the protonated or unprotonated histidine ligand in the high-potential and low-potential forms of cytochrome b 559, respectively.  相似文献   

7.
Scanning transmission X-ray microscopy (STXM) at the C 1s, O 1s, Ni 2p, Ca 2p, Mn 2p, Fe 2p, Mg 1s, Al 1s and Si 1s edges was used to study Ni sorption in a complex natural river biofilm. The 10-week grown river biofilm was exposed to 10 mg L−1 Ni2+ (as NiCl2) for 24 h. The region of the biofilm examined was dominated by filamentous structures, which were interpreted as the discarded sheaths of filamentous bacteria, as well as a sparse distribution of rod-shaped bacteria. The region also contained discrete particles with spectra similar to those of muscovite, SiO2 and CaCO3. The Ni(II) ions were selectively adsorbed by the sheaths of the filamentous bacteria. The sheaths were observed to be metal rich with significant amounts of Ca, Fe and Mn, along with the Ni. In addition, the sheaths had a large silicate content but little organic material. The metal content of the rod-shaped bacterial cells was much lower. The Fe on the sheath was mainly in the Fe(III) oxidation state. Mn was found in II, III and IV oxidation states. The Ni was likely sorbed to Mn–Fe minerals on the sheath. These STXM results have probed nano-scale biogeochemistry associated with bacterial species in a complex, natural biofilm community. They have implications for selective Ni contamination of the food chain and for developing bioremediation strategies.  相似文献   

8.
Iron ions in the two iron centers of beef heart mito-chondrial F, ATPase, which we have been recently characterized (FEBS Letters 1996,379, 231-235), exhibit different redox properties. In fact, the ATP-dependent site is able to maintain iron in the redox state of Fe(II) even in the absence of reducing agents, whereas in the nucleotide-independent site iron is oxidized to Fe(III) upon removal of the reductant. Fe(III) ions in the two sites display different reactivity towards H2O2, because only Fe(III) bound in the nucleotide-independent site rapidly reacts with H2O2 thus mediating a 30% enzyme inactivation. Thermophilic bacterium PS3 bears one Fe(III) binding site, which takes up Fe(III) either in the absence or presence of nucleotides and is unable to maintain iron in the redox state of Fe(II) in the absence of ascorbate. Fe(III) bound in thermophilic F1ATPase in a molar ratio 1:1 rapidly reacts with H2O2 mediating a 30% enzyme inactivation. These results support the presence in mitochon-drial and thermophilic F1ATPase of a conserved site involved in iron binding and in oxidative inactivation, in which iron exhibits similar redox properties. On the other hand, at variance with thermophilic F1ATPase, the mitochondrial enzyme has the possibility of maintaining one equivalent of Fe(II) in its peculiar ATP-dependent site, besides one equivalent of Fe(III) in the conserved nucleotide-independent site. In this case mitochondrial F, ATPase undergoes a higher inactivation (75%) upon exposure to H2O2. Under all conditions the inactivation is significantly prevented by PBN and DMSO but not by Cu, Zn superoxide dis-mutase, thus suggesting the formation of OH radicals as mediators of the oxidative damage. No dityrosines, carbonyls or oxidized thiols are formed. In addition, in any cases no protein fragmentation or aggregation is observed upon the treatment with H2O2.  相似文献   

9.
Degeneration of biogenic superparamagnetic magnetite   总被引:2,自引:0,他引:2  
Magnetite crystals precipitated as a consequence of Fe(III) reduction by Shewanella algae BrY after 265 h incubation and 5-year anaerobic storage were investigated with transmission electron microscopy, Mössbauer spectroscopy and X-ray diffraction. The magnetite crystals were typically superparamagnetic with an approximate size of 13 nm. The lattice constants of the 265 h and 5-year crystals are 8.4164Å and 8.3774Å, respectively. The Mössbauer spectra indicated that the 265 h magnetite had excess Fe(II) in its crystal-chemistry (Fe3+1.990Fe2+1.015O4) but the 5-year magnetite was Fe(II)-deficient in stoichiometry (Fe3+2.388Fe2+0.419O4). Such crystal-chemical changes may be indicative of the degeneration of superparamagnetic magnetite through the aqueous oxidization of Fe(II) anaerobically, and the concomitant oxidation of the organic phases (fatty acid methyl esters) that were present during the initial formation of the magnetite. The observation of a corona structure on the aged magnetite corroborates the anaerobic oxidation of Fe(II) on the outer layers of magnetite crystals. These results suggest that there may be a possible link between the enzymatic activity of the bacteria and the stability of Fe(II)-excess magnetite, which may help explain why stable nano-magnetite grains are seldom preserved in natural environments.  相似文献   

10.
Cyclic voltammetry at a micro electrode of Co(II) salen, Fe(II) salen, electrode generated Fe(II)(acac)2, Fe(II) (salicylaldehyde)2, Fe(II) (salicylaldoxime)2, Fe(II) (bipy)3, Fe(II) (bipy)2, Co(II) (bipy)3, Co(II) (benzacac)2, and electrode generated Co(acac)2 in oxygen saturated aprotic solvents show positive shift of the O2 sigmoidal wave, as well as enhancement of the limiting current in the case of the first five compounds. In the case of Co(II) (bipy)3 the slope of the sigmoidal wave due to O2 becomes more positive, while for the other two Co(II) complexes there is no change except a small decrease in the wave height. The data are used to correlate and predict the O2 binding properties of the chelates in solution. The data for the diketone complexes of Co(II) indicate absence of any direct association, which is in line with the interpretation offered in the literature on the mechanism of their catalytic role in the O2 oxidation of substrates. The mechanism of the autoxidation of dimethylformamide in the presence of Fe(III) (bipy)3 and Cu(II) (bipy)2 is elucidated by the observation that these higher valent compounds are reduced to their next lower oxidation state by DMF.  相似文献   

11.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

12.
Fe-limited cells of the green alga Chlamydomonas reinhardtii (Fe-limited growth rate = 0.3 d−1) reduced extracellular Fe(III) to Fe(II) when Fe(III) was supplied as ferricyanide or Fe(III)-EDTA; Fe(III) reduction was stimulated by light. In both darkness and during photosynthesis, ferricyanide reduction was accompanied by a decrease in cellular NADPH levels, with a concomitant increase in NADP+. NADH and NAD+ levels were not measurably altered during ferricyanide reduction. Furthermore, cellular hexose monophosphate levels declined and 6-phosphogluconate levels increased during ferricyanide reduction. Levels of most glycolytic and tricarboxylic acid cycle intermediates were mostly unaltered. Ferricyanide reduction was also associated with a decrease in cellular ATP levels, a concomitant increase in ADP and AMP, and increased extracellular acidification. The acidification was sensitive to inhibition by the H+-ATPase inhibitor N,N' -dicyclohexylcarbodiimide (DCCD). We conclude that the oxidative pentose phosphate pathway provides reducing equivalents for Fe(III) reduction in darkness and also contributes reducing equivalents to Fe(III) reduction during photosynthesis. The decline in ATP was likely due to activation of the plasma membrane H+-ATPase during ferricyanide reduction and was not directly associated with provision of reducing equivalents.  相似文献   

13.
1. Six- to eight-week greenhouse experiments with independent control of pH and dissolved CO2 evaluated the potential for CO2 enrichment to stimulate the accumulation of Al, Fe, P and N in shoots of Vallisneria americana , particularly at pH 5. These minerals were provided only as they occurred in natural lake sediments.
2. The effect of CO2 enrichment at pH 5 v pH 7.3 on growth and tissue N concentration was also determined.
3. CO2 enrichment at pH 5 effected 5.5- and 7-fold increases in total shoot accumulation of Al and Fe, respectively. In a two-way factorial experiment, CO2 enrichment yielded 6- to 11-fold greater total shoot P accumulation in plants grown on less and more fertile sediments, respectively.
4. In a three-way factorial experiment, CO2 enrichment stimulated Vallisneria growth, especially at pH 5, and resulted in a 31–58% reduction in tissue [N] for different pH × sediment combinations. These are greater reductions than previously reported. It also increased total shoot N accumulation up to 6-fold, and there were significant interactions with pH and sediment source: the CO2 enrichment effect on shoot N accumulation was greater at pH 5 than at pH 7.3, and it was greater with the more fertile sediment at pH 5.
5. Water chemistry (pH and/or [CO2]) and sediment fertility thus both indirectly influenced the accumulation of sediment-derived minerals in macrophyte shoots within the water column.  相似文献   

14.
1. Six- to eight-week greenhouse experiments with independent control of pH and dissolved CO2 evaluated the potential for CO2 enrichment to stimulate the accumulation of Al, Fe, P and N in shoots of Vallisneria americana , particularly at pH 5. These minerals were provided only as they occurred in natural lake sediments.
2. The effect of CO2 enrichment at pH 5 v pH 7.3 on growth and tissue N concentration was also determined.
3. CO2 enrichment at pH 5 effected 5.5- and 7-fold increases in total shoot accumulation of Al and Fe, respectively. In a two-way factorial experiment, CO2 enrichment yielded 6- to 11-fold greater total shoot P accumulation in plants grown on less and more fertile sediments, respectively.
4. In a three-way factorial experiment, CO2 enrichment stimulated Vallisneria growth, especially at pH 5, and resulted in a 31–58% reduction in tissue [N] for different pH × sediment combinations. These are greater reductions than previously reported. It also increased total shoot N accumulation up to 6-fold, and there were significant interactions with pH and sediment source: the CO2 enrichment effect on shoot N accumulation was greater at pH 5 than at pH 7.3, and it was greater with the more fertile sediment at pH 5.
5. Water chemistry (pH and/or [CO2]) and sediment fertility thus both indirectly influenced the accumulation of sediment-derived minerals in macrophyte shoots within the water column.  相似文献   

15.
Anaeromyxobacter dehalogenans strain 2CP-C reduces U(VI) and Tc(VII) to U(IV)O2(s) (uraninite) and Tc(IV)O2(S) respectively. Kinetic studies with resting cells revealed that U(VI) or Tc(VII) reduction rates using H2 as electron donor exceeded those observed in acetate-amended incubations. The reduction of U(VI) by A. dehalogenans 2CP-C resulted in extracellular accumulation of ∼5 nm uraninite nanoparticles in association with a lectin-binding extracellular polymeric substance (EPS). The electron donor did not affect UO2(S) nanoparticle size or association with EPS, but the utilization of acetate as the source of reducing equivalents resulted in distinct UO2(S) nanoparticle aggregates that were ∼50 nm in diameter. In contrast, reduction of Tc(VII) by A. dehalogenans 2CP-C cell suspensions produced dense clusters of TcO2 particles, which were localized within the cell periplasm and on the outside of the outer membrane. In addition to direct reduction, A. dehalogenans 2CP-C cell suspensions reduced Tc(VII) indirectly via an Fe(II)-mediated mechanism. Fe(II) produced by strain 2CP-C from either ferrihydrite or Hanford Site sediment rapidly removed 99Tc(VII)O4 from solution. These findings expand our knowledge of the radionuclide reduction processes catalysed by Anaeromyxobacter spp. that may influence the fate and transport of radionuclide contaminants in the subsurface.  相似文献   

16.
Abstract: Mitochondrial complexes I, II, and III were studied in isolated brain mitochondrial preparations with the goal of determining their relative abilities to reduce O2 to hydrogen peroxide (H2O2) or to reduce the alternative electron acceptors nitroblue tetrazolium (NBT) and diphenyliodonium (DPI). Complex I and II stimulation caused H2O2 formation and reduced NBT and DPI as indicated by dichlorodihydrofluorescein oxidation, nitroformazan precipitation, and DPI-mediated enzyme inactivation. The O2 consumption rate was more rapid under complex II (succinate) stimulation than under complex I (NADH) stimulation. In contrast, H2O2 generation and NBT and DPI reduction kinetics were favored by NADH addition but were virtually unobservable during succinate-linked respiration. NADH oxidation was strongly suppressed by rotenone, but NADH-coupled H2O2 flux was accelerated by rotenone. α-Phenyl- N-tert -butyl nitrone (PBN), a compound documented to inhibit oxidative stress in models of stroke, sepsis, and parkinsonism, partially inhibited complex I-stimulated H2O2 flux and NBT reduction and also protected complex I from DPI-mediated inactivation while trapping the phenyl radical product of DPI reduction. The results suggest that complex I may be the principal source of brain mitochondrial H2O2 synthesis, possessing an "electron leak" site upstream from the rotenone binding site (i.e., on the NADH side of the enzyme). The inhibition of H2O2 production by PBN suggests a novel explanation for the broad-spectrum antioxidant and antiinflammatory activity of this nitrone spin trap.  相似文献   

17.
In an effort to improve the understanding of electron transfer mechanisms at the microbe–mineral interface, Shewanella oneidensis MR-1 mutants with in-frame deletions of outer-membrane cytochromes (OMCs), MtrC and OmcA, were characterized for the ability to reduce ferrihydrite (FH) using a suite of microscopic, spectroscopic, and biochemical techniques. Analysis of purified recombinant proteins demonstrated that both cytochromes undergo rapid electron exchange with FH in vitro with MtrC displaying faster transfer rates than OmcA. Immunomicroscopy with cytochrome-specific antibodies revealed that MtrC co-localizes with iron solids on the cell surface while OmcA exhibits a more diffuse distribution over the cell surface. After 3-day incubation of MR-1 with FH, pronounced reductive transformation mineral products were visible by electron microscopy. Upon further incubation, the predominant phases identified were ferrous phosphates including vivianite [Fe3(PO4)2·8H2O] and a switzerite-like phase [Mn3,Fe3(PO4)2·7H2O] that were heavily colonized by MR-1 cells with surface-exposed outer-membrane cytochromes. In the absence of both MtrC and OmcA, the cells ability to reduce FH was significantly hindered and no mineral transformation products were detected. Collectively, these results highlight the importance of the outer-membrane cytochromes in the reductive transformation of FH and support a role for direct electron transfer from the OMCs at the cell surface to the mineral.  相似文献   

18.
Here we describe a novel bacterial community that is embedded in a matrix of carbohydrates and bio/geochemical products of pyrite (FeS2) oxidation. This community grows in stalactite-like structures – snottites – on the ceiling of an abandoned pyrite mine at pH values of 2.2–2.6. The aqueous phase in the matrix contains 200 mM of sulfate and total iron concentrations of 60 mM. Micro-X-ray diffraction analysis showed that jarosite [(K,Na,H3O)Fe3(SO4)2(OH)6] is the major mineral embedded in the snottites. X-ray absorption near-edge structure experiments revealed three different sulfur species. The major signal can be ascribed to sulfate, and the other two features may correspond to thiols and sulfoxides. Arabinose was detected as the major sugar component in the extracellular polymeric substance. Via restriction fragment length polymorphism analysis, a community was found that mainly consists of iron oxidizing Leptospirillum and Ferrovum species but also of bacteria that could be involved in dissimilatory sulfate and dissimilatory iron reduction. Each snottite can be regarded as a complex, self-contained consortium of bacterial species fuelled by the decomposition of pyrite.  相似文献   

19.
Soils contain two different activities for oxidation of hydrogen   总被引:1,自引:0,他引:1  
Abstract Hydrogen oxidation rates were measured in a neutral compost soil and an acidic sandy loam at H2 mixing ratios of 0.01 to 5000 ppmv. The kinetics were biphasic showing two different K m values for H2, one at about 10–40 nM dissolved H2, the other at about 1.2–1.4 μM H2. The low- K m activity was less sensitive to chloroform fumigation than the high- K m activity. If sterile soil was amended with Paracoccus denitrificans or a H2-oxidizing strain isolated from compost soil, it exhibited only a high- K m (0.7–0.9 μM) activity. It also failed to utilize H2 mixing ratios below a threshold of 1.6–3.0 ppmv H2 (160–300 mPa). A similar result was obtained when fresh soil samples were suspended in water, and H2 oxidation was determined from the decrease of dissolved H2. However, H2 was again utilized to mixing ratios lower than 0.05 ppmv, if the supernatant of the soil suspension or the settled soil particles were dried onto sterile soil or purified quarz sand. Obviously, soils contain two different activities for oxidation of H2: (1) a high- K m, high-threshold activity which apparently is due to aerobic H2-oxidizing bacteria, and (2) a low- K m, low-threshold activity whose origin is unknown but presumably is due to soil enzymes.  相似文献   

20.
Isolates of the ericoid mycorrhizal fungus Hymenoscyphus ericae (Read) Korf et Kernan, and the ectomycorrhizal fungi Suillus variegatus (Swartz ex Fr.) and Pisolithus tinctorius (Pers.) Coker & Couch, along with a Cortinarius sp. and the white rot Phanerochaete chrysosporium Burdsall were examined for the ability to oxidize carbohydrates to their corresponding lactones and to excrete the H2O2 produced thereby. All except Phanerochaete chrysosporium were found to express cellobiose oxidase (cellobiose dehydrogenase, EC 1.1.19.88) and glucose oxidase (β- d -glucose:oxygen 1-oxidoreductase, EC 1.1.3.4) when grown on cellobiose and glucose respectively. Production of extracellular H2O2 was visualized during growth on both substrates using ABTS as the chromogen. According to the Fenton reaction, H2O2 will react with hydrated or chelated Fe(II) in the environment to produce hydroxyl (Fenton) radicals, HO·. Mycelial extracts from each of the mycorrhizal fungi produced HO· in the presence of cellobiose and Fe(II), presumably mediated by H2O2 produced by cellobiose oxidase activity in the extracts. Conditions favourable to HO· production were shown to exist in Modified Melin–Norkrans medium, and the data discussed in relation to previously observed lignin degradation by mycorrhizal fungi.  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