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 109 毫秒
1.
Microalgae are extensively used in the remediation of heavy metals like iron. However, factors like toxicity, bioavailability and iron speciation play a major role in its removal by microalgae. Thus, in this study, toxicity of three different iron salts (FeSO4, FeCl3 and Fe(NO3)3) was evaluated towards three soil microalgal isolates, Chlorella sp. MM3, Chlamydomonas sp. MM7 and Chlorococcum sp. MM11. Interestingly, all the three iron salts gave different EC50 concentrations; however, ferric nitrate was found to be significantly more toxic followed by ferrous sulphate and ferric chloride. The EC50 analysis revealed that Chlorella sp. was significantly resistant to iron compared to other microalgae. However, almost 900 μg g?1 iron was accumulated by Chlamydomonas sp. grown with 12 mg L?1 ferric nitrate as an iron source when compared to other algae and iron salts. The time-course bioaccumulation confirmed that all the three microalgae adsorb the ferric salts such as ferric nitrate and ferric chloride more rapidly than ferrous salt, whereas intracellular accumulation was found to be rapid for ferrous salts. However, the amount of iron accumulated or adsorbed by algae, irrespective of species, from ferrous sulphate medium is comparatively lower than ferric chloride and ferric nitrate medium. The Fourier transform infrared spectroscopy (FTIR) analysis shows that the oxygen atom and P?=?O group of polysaccharides present in the cell wall of algae played a major role in the bioaccumulation of iron ions by algae.  相似文献   

2.
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   

3.
Ferric Iron Reduction by Acidophilic Heterotrophic Bacteria   总被引:8,自引:4,他引:4       下载免费PDF全文
Fifty mesophilic and five moderately thermophilic strains of acidophilic heterotrophic bacteria were tested for the ability to reduce ferric iron in liquid and solid media under aerobic conditions; about 40% of the mesophiles (but none of the moderate thermophiles) displayed at least some capacity to reduce iron. Both rates and extents of ferric iron reduction were highly strain dependent. No acidophilic heterotroph reduced nitrate or sulfate, and (limited) reduction of manganese(IV) was noted in only one strain (Acidiphilium facilis), an acidophile which did not reduce iron. Insoluble forms of ferric iron, both amorphous and crystalline, were reduced, as well as soluble iron. There was evidence that, in at least some acidophilic heterotrophs, iron reduction was enzymically mediated and that ferric iron could act as a terminal electron acceptor. In anaerobically incubated cultures, bacterial biomass increased with increasing concentrations of ferric but not ferrous iron. Mixed cultures of Thiobacillus ferrooxidans or Leptospirillum ferrooxidans and an acidophilic heterotroph (SJH) produced sequences of iron cycling in ferrous iron-glucose media.  相似文献   

4.
When tetanus toxin is made by fermentation with Clostridium tetani, the traditional source of iron is an insoluble preparation called reduced iron powder. This material removes oxygen from the system by forming FeO2 (rust). When inoculated in a newly developed medium lacking animal and dairy products and containing glucose, soy-peptone, and inorganic salts, growth and toxin production were poor without reduced iron powder. The optimum concentration of reduced iron powder for toxin production was found to be 0.5 g/l. Growth was further increased by higher concentrations, but toxin production decreased. Inorganic iron sources failed to replace reduced iron powder for growth or toxin formation. The iron source that came closest was ferrous ammonium sulfate. The organic iron sources ferric citrate and ferrous gluconate were more active than the inorganic compounds but could not replace reduced iron powder. Insoluble iron sources, such as iron wire, iron foil, and activated charcoal, were surprisingly active. Combinations of activated charcoal with soluble iron sources such as ferrous sulfate, ferric citrate, and ferrous gluconate showed increased activity, and the ferrous gluconate combination almost replaced reduced iron powder. It thus appears that the traditional iron source, reduced iron powder, plays a double role in supporting tetanus toxin formation, i.e., releasing soluble sources of iron and providing an insoluble surface.  相似文献   

5.
Fenton reaction is thought to play an important role in wood degradation by brown-rot fungi. In this context, the effect of oxalic acid and pH on iron reduction by a biomimetic fungal chelator and on the adsorption/desorption of iron to/from wood was investigated. The results presented in this work indicate that at pH 2.0 and 4.5 and in the presence of oxalic acid, the phenolate chelator 2,3-dihydroxybenzoic acid (2,3-DHBA) is capable of reducing ferric iron only when the iron is complexed with oxalate to form Fe3+-mono-oxalate (Fe(C2O4)+). Within the pH range tested in this work, this complex formation occurs when the oxalate:Fe3+ molar ratio is less than 20 (pH 2.0) or less than 10 (pH 4.5). When aqueous ferric iron was passed through a column packed with milled red spruce (Picea rubens) wood equilibrated at pH 2.0 and 4.5, it was observed that ferric iron binds to wood at pH 4.5 but not at pH 2.0, and the bound iron could then be released by application of oxalic acid at pH 4.5. The release of bound iron was dependent on the amount of oxalic acid applied in the column. When the amount of oxalate was at least 20-fold greater than the amount of iron bound to the wood, all bound iron was released. When Fe–oxalate complexes were applied to the milled wood column equilibrated in the pH range of 2–4.5, iron from Fe–oxalate complexes was bound to the wood only when the pH was 3.6 or higher and the oxalate:Fe3+ molar ratio was less than 10. When 2,3-DHBA was evaluated for its ability to release iron bound to the milled wood, it was found that 2,3-DHBA possessed a greater affinity for ferric iron than the wood as 2,3-DHBA was capable of releasing the ferric iron bound to the wood in the pH range 3.6–5.5. These results further the understanding of the mechanisms employed by brown-rot fungi in wood biodegradation processes.  相似文献   

6.
Five moderately thermophilic iron-oxidizing bacteria, including representative strains of the three classified species (Sulfobacillus thermosulfidooxidans, Sulfobacillus acidophilus, and Acidimicrobium ferrooxidans), were shown to be capable of reducing ferric iron to ferrous iron when they were grown under oxygen limitation conditions. Iron reduction was most readily observed when the isolates were grown as mixotrophs or heterotrophs with glycerol as an electron donor; in addition, some strains were able to couple the oxidation of tetrathionate to the reduction of ferric iron. Cycling of iron between the ferrous and ferric states was observed during batch culture growth in unshaken flasks incubated under aerobic conditions, although the patterns of oxidoreduction of iron varied in different species of iron-oxidizing moderate thermophiles and in strains of a single species (S. acidophilus). All three bacterial species were able to grow anaerobically with ferric iron as a sole electron acceptor; the growth yields correlated with the amount of ferric iron reduced when the isolates were grown in the absence of oxygen. One of the moderate thermophiles (identified as a strain of S. acidophilus) was able to bring about the reductive dissolution of three ferric iron-containing minerals (ferric hydroxide, jarosite, and goethite) when it was grown under restricted aeration conditions with glycerol as a carbon and energy source. The significance of iron reduction by moderately thermophilic iron oxidizers in both environmental and applied contexts is discussed.Moderately thermophilic acidophilic bacteria that catalyze the dissimilatory oxidation of ferrous iron are distinct both phylogenetically and in aspects of their physiology. They differ from the known acidophilic mesophilic iron oxidizers (gram-negative, nonsporulating chemolithotrophic bacteria) and the extremely thermophilic iron oxidizers (certain archaea) in several fundamental ways, including cellular morphology (they are gram-positive rods that often form endospores) and growth temperature optima, which are typically 45 to 55°C (15). In addition, the moderately thermophilic iron-oxidizing acidophiles characteristically have a highly versatile metabolism (18) and may grow as autotrophs (e.g., in media containing ferrous iron or reduced sulfur), heterotrophs (e.g., on yeast extract), mixotrophs (e.g., in media containing both ferrous iron and glucose, in which both CO2 and glucose are used as carbon sources), or chemolithoheterotrophs (e.g., in ferrous iron-yeast extract medium, in which iron acts as the energy source and yeast extract is the carbon source). Isolates have been obtained from a range of thermal acidic environments, such as geothermal areas, self-heating mine waste spoils, and commercial mineral-processing operations (2a, 5, 14). There are currently two recognized genera of these bacteria. All but one Sulfobacillus species are iron- and sulfur-oxidizing, gram-positive, sporulating rods. Two such species have been described, Sulfobacillus thermosulfidooxidans and Sulfobacillus acidophilus, which may be distinguished by their different chromosomal DNA base compositions and by their abilities to grow autotrophically on reduced sulfur (16). The genus Acidimicrobium currently contains a single species, Acidimicrobium ferrooxidans. This organism differs from Sulfobacillus spp. by its greater capacity to fix CO2, by its lower tolerance of ferric iron, by its apparent lack of spore formation (although it is also gram positive), and by its chromosomal DNA base composition (4). Analysis of 16S rRNA sequences has also differentiated this moderate thermophile from Sulfobacillus spp. (9).The small amount of energy associated with the oxidation of ferrous iron (−30 kJ mol−1 at pH 2) can serve as the exclusive source of energy for moderately thermophilic iron-oxidizing acidophiles when they are growing autotrophically with oxygen as the terminal electron acceptor. Under limited aeration conditions, ferric iron, which is often abundant and present in a soluble form in extremely acidic environments, is a thermodynamically attractive alternative electron sink (electrode potential [E′], +780 mV). Ferric iron reduction by mesophilic chemolithotrophic and heterotrophic acidophiles has been observed previously (5, 7, 17). Some moderately thermophilic, acidophilic, heterotrophic bacteria (Alicyclobacillus-like isolates) (5a) and the extremely thermophilic archaeon Sulfolobus acidocaldarius (3) can also reduce iron. While many neutrophilic microorganisms are also able to reduce ferric iron, the ability to conserve energy to support growth by coupling organic matter oxidation exclusively to ferric iron reduction appears to be more restricted among neutrophilic bacteria (11).In this paper, we describe the dissimilatory reduction of ferric iron by representative isolates of different species of iron-oxidizing moderate thermophiles with both an organic electron donor (glycerol) and an inorganic electron donor (tetrathionate), and we also describe the reductive dissolution of ferric iron-containing minerals by a Sulfobacillus isolate.  相似文献   

7.
Thiobacillus ferrooxidans is found in many Alaskan and Canadian drainages contaminated by metals dissolved from placer and lode gold mines. We have examined the iron-limited growth and iron oxidation kinetics of a T. ferrooxidans isolate, AK1, by using batch and continuous cultures. Strain AK1 is an arsenic-tolerant isolate obtained from placer gold mine drainage containing large amounts of dissolved arsenic. The steady-state growth kinetics are described with equations modified for threshold ferrous iron concentrations. The maximal specific growth rate (μmax) for isolate AK1 at 22.5°C was 0.070 h−1, and the ferrous iron concentration at which the half-maximal growth rate occurred (Kμ) was 0.78 mM. Cell yields varied inversely with growth rate. The iron oxidation kinetics of this organism were dependent on biomass. We found no evidence of ferric inhibition of ferrous iron oxidation for ferrous iron concentrations between 9.0 and 23.3 mM. A supplement to the ferrous medium of 2.67 mM sodium arsenite did not result in an increased steady-state biomass, nor did it appear to affect the steady-state growth kinetics observed in continuous cultures.  相似文献   

8.
Myoglobin (Mb) is reported in biochemistry and physiology textbooks to act as an O2 reservoir and to facilitate O2 diffusion from capillaries to mitochondria, to sustain cellular respiration. Recently, it has been proposed that Mb is an intracellular scavenger of bioactive nitric oxide (NO), regulating its level in the skeletal and cardiac muscle and thereby protecting mitochondrial respiration, which is impaired by NO. This novel function of Mb is based on the rapid and irreversible reaction of ferrous oxygenated Mb (MbO2) with NO yielding ferric oxidized Mb (metMb) and nitrate (NO3). The efficiency of this process, which is postulated to depend on the superoxide (O2) character acquired by O2 once bound to the heme iron, may be enhanced by intramolecular diffusion of NO trapped momentarily into cavities of the protein matrix. O2 can also react with ferrous nitrosylated Mb (MbNO), albeit very slowly, leading to metMb and NO3. The O2-dependent NO-detoxification process may be considered to be pseudo-enzymatic given that metMb obtained by the primary reaction of MbO2 with NO is reduced back to ferrous Mb by a specific metMb-reductase, and can therefore repeat a cycle of NO conversion to harmless nitrate.  相似文献   

9.
A lactic acid bacterium capable of anaerobic respiration was isolated from soil with ferric iron-containing glucose basal medium and identified as L. garvieae by using 16S rDNA sequence homology. The isolate reduced ferric iron, nitrate, and fumarate to ferrous iron, nitrite, and succinate, respectively, under anaerobic N2 atmosphere. Growth of the isolate was increased about 30-39% in glucose basal medium containing nitrate and fumarate, but not in the medium containing ferric iron. Specifically, metabolic reduction of nitrate and fumarate is thought to be controlled by the specific genes fnr, encoding FNR-like protein, and nir, regulating fumarate-nitrate reductase. Reduction activity of ferric iron by the isolate was estimated physiologically, enzymologically, and electrochemically. The results obtained led us to propose that the isolate metabolized nitrate and fumarate as an electron acceptor and has specific enzymes capable of reducing ferric iron in coupling with anaerobic respiration.  相似文献   

10.
Inge Romslo  Torgeir Flatmark 《BBA》1975,387(1):80-94
1. Depending on the metabolic state, the addition of iron(III)-sucrose induces an inhibition or a stimulation of the respiration rate when added to isolated rat liver mitochondria.2. Under conditions identical to those used in the accumulation studies (Romslo, I. and Flatmark, T. (1973) Biochim. Biophys. Acta 305, 29?40), the ferric complex induces a decrease in the oxygen uptake concomitant to an oxidation of cytochromes c (+c1) and a (+a3). These results suggest that ferric iron is reduced to ferrous iron by the respiratory chain prior to or simultaneously with its energy-dependent accumulation.3. On the other hand, the addition of iron(III)-sucrose induces a stimulation of respiration in State 4 and State 3 provided Mg2+ is present in the suspending medium. In contrast to Ca2+, iron stimulates State 4 respiration in a cyclic process only within narrow concentration limits; at concentrations of iron above 100 μM the respiration remains in the activated state until anaerobiosis. The stimulation of State 4 respiration is more pronounced with succinate than with NAD-linked substrates, a difference which partly may be attributed to a stimulation of the succinate dehydrogenase complex.4. The stimulation of respiration by iron is approx. 3 times higher in State 3 than in State 4 and this difference can be attributed to a stimulation of the adenine nucleotide exchange reaction in State 3 with a concomitant increase in the rate of oxidative phosphorylation, although the PO ratio is slightly diminished.  相似文献   

11.
Iron (Fe) is essential for virtually all living organisms. The identification of the chemical forms of iron (the speciation) circulating in and between cells is crucial to further understand the mechanisms of iron delivery to its final targets. Here we analyzed how iron is transported to the seeds by the chemical identification of iron complexes that are delivered to embryos, followed by the biochemical characterization of the transport of these complexes by the embryo, using the pea (Pisum sativum) as a model species. We have found that iron circulates as ferric complexes with citrate and malate (Fe(III)3Cit2Mal2, Fe(III)3Cit3Mal1, Fe(III)Cit2). Because dicotyledonous plants only transport ferrous iron, we checked whether embryos were capable of reducing iron of these complexes. Indeed, embryos did express a constitutively high ferric reduction activity. Surprisingly, iron(III) reduction is not catalyzed by the expected membrane-bound ferric reductase. Instead, embryos efflux high amounts of ascorbate that chemically reduce iron(III) from citrate-malate complexes. In vitro transport experiments on isolated embryos using radiolabeled 55Fe demonstrated that this ascorbate-mediated reduction is an obligatory step for the uptake of iron(II). Moreover, the ascorbate efflux activity was also measured in Arabidopsis embryos, suggesting that this new iron transport system may be generic to dicotyledonous plants. Finally, in embryos of the ascorbate-deficient mutants vtc2-4, vtc5-1, and vtc5-2, the reducing activity and the iron concentration were reduced significantly. Taken together, our results identified a new iron transport mechanism in plants that could play a major role to control iron loading in seeds.  相似文献   

12.
Summary A study of transformations of added nitrate nitrogen in an acid sulfate soil under waterlogged conditions indicated that chemical reduction of NO3–N to NH4–N is effected by high concentrations of ferrous iron, released in the soil following flooding and reduction of ferric to ferrous iron.  相似文献   

13.
Two free flavin-independent enzymes were purified by detecting the NAD(P)H oxidation in the presence of Fe(III)-EDTA and t-butyl hydroperoxide from E. coli. The enzyme that requires NADH or NADPH as an electron donor was a 28 kDa protein, and N-terminal sequencing revealed it to be oxygen-insensitive nitroreductase (NfnB). The second enzyme that requires NADPH as an electron donor was a 30 kDa protein, and N-terminal sequencing revealed it to be ferredoxin-NADP+ reductase (Fpr). The chemical stoichiometry of the Fenton activities of both NfnB and Fpr in the presence of Fe(III)-EDTA, NAD(P)H and hydrogen peroxide was investigated. Both enzymes showed a one-electron reduction in the reaction forming hydroxyl radical from hydrogen peroxide. Also, the observed Fenton activities of both enzymes in the presence of synthetic chelate iron compounds were higher than their activities in the presence of natural chelate iron compounds. When the Fenton reaction occurs, the ferric iron must be reduced to ferrous iron. The ferric reductase activities of both NfnB and Fpr occurred with synthetic chelate iron compounds. Unlike NfnB, Fpr also showed the ferric reductase activity on an iron storage protein, ferritin, and various natural iron chelate compounds including siderophore. The Fenton and ferric reductase reactions of both NfnB and Fpr occurred in the absence of free flavin. Although the k cat/K m value of NfnB for Fe(III)-EDTA was not affected by free flavin, the k cat/K m value of Fpr for Fe(III)-EDTA was 12-times greater in the presence of free FAD than in the absence of free FAD.  相似文献   

14.
Using Mössbauer resonance spectroscopy of 57Fe, we have determined the nature and distribution of major iron compounds in the magnetotactic bacterium Aquaspirillum magnetotacticum. In addition to magnetite (Fe3O4), cells contained a low-density hydrous ferric oxide, a high-density hydrous ferric oxide (ferrihydrite), and ferrous iron. Analysis at different temperatures of whole cells harvested early and late in growth, of mutant cells unable to synthesize magnetite, and of cell fractions enriched in 57Fe indicated that Fe3O4 precipitation resulted from partial reduction of the high-density hydrous ferric oxide precursor.  相似文献   

15.
The reduced forms of cytochrome P-450cam and chloroperoxidase were examined by proton NMR spectroscopy. The pH and temperature dependences of the proton NMR spectra of both ferrous enzymes are reported. A series of alkyl mercaptide complexes of both synthetic and natural-derivative iron(II) porphyrins was also examined. The proton NMR spectra of these complexes facilitated the assignment of resonances due to the axial ligand in the model compounds on the basis of their isotropic shifts and multiplicities. Comparison of model compound data with that for the reduced enzymes supports assignment of the methylene protons for the axial cysteinate of ferrous cytochrome P-450cam and ferrous chloroperoxidase to proton NMR resonances at 279 and 200 ppm (pH 7.0, 298K), respectively. Differences in the active site structure of the two enzymes are further demonstrated by 15N-NMR spectroscopy of the cyanide complexes of the ferric forms.  相似文献   

16.
Anaerobic Growth of Thiobacillus ferrooxidans   总被引:4,自引:0,他引:4       下载免费PDF全文
The obligately autotrophic acidophile Thiobacillus ferrooxidans was grown on elemental sulfur in anaerobic batch cultures, using ferric iron as an electron acceptor. During anaerobic growth, ferric iron present in the growth media was quantitatively reduced to ferrous iron. The doubling time in anaerobic cultures was approximately 24 h. Anaerobic growth did not occur in the absence of elemental sulfur or ferric iron. During growth, a linear relationship existed between the concentration of ferrous iron accumulated in the cultures and the cell density. The results suggest that ferric iron may be an important electron acceptor for the oxidation of sulfur compounds in acidic environments.  相似文献   

17.
Young mycelia of the fungus Neurospora crassa contain a soluble NADH-linked sideramine reductase, which may be responsible for liberating iron in vivo from accumulated sideramines during iron-deficient cultivation. The enzymes can be assayed using a soluble supernatant fraction, EDTA, and an atmosphere of pure nitrogen. The enzyme is stable without loss of activity up to 45°C and has an optimum of activity at pH 7.0. Besides coprogen (Km = 100 μM, V = 2.8 nmol/min. per mg protein), some other ferrichrome-type compounds are reduced. However, ferrichrome, ferrirubin, coprogen B and ferrioxamine are poor substrates. When the mucelia were grown in a medium containing 10?5 M ferric iron, the activity of the reductase was found to be only 30% of that found under low iron conditions. The enzyme is inhibited by oxygen, SH-alkylating agents and partly by some detergents. Unlike the reductase of N. crassa, the corresponding enzyme from Aspergillus fumigatus revealed low reduction of coprogen and high reduction of ferrichrome, indicating genus-dependent specificities of sideramine reduction enzymes in fungi. The participation of acids of the citric acid cycle as natural iron acceptors during strong iron deficiency is studied and confirmed by iron uptake measurements on isolated mitochondria.  相似文献   

18.
The endogenous cation in peroxidases may contribute to the type of heme coordination. Here a series of ferric and ferrous derivatives of wild-type Leishmania major peroxidase (LmP) and of engineered K+ site mutants of LmP, lacking potassium cation binding site, has been examined by electronic absorption spectroscopy at 25 °C. Using UV–visible spectrophotometry, we show that the removal of K+ binding site causes substantial changes in spin states of both the ferric and ferrous forms. The spectral changes are interpreted to be, most likely, due to the formation of a bis-histidine coordination structure in both the ferric and ferrous oxidation states at neutral pH 7.0. Stopped flow spectrophotometric techniques revealed that characteristics of Compound I were not observed in the K+ site double mutants in the presence of H2O2. Similarly electron donor oxidation rate was two orders less for the K+ site double mutants compared to the wild type. These data show that K+ functions in preserving the protein structure in the heme surroundings as well as the spin state of the heme iron, in favor of the enzymatically active form of LmP.  相似文献   

19.
A series of ferric and ferrous derivatives of wild-type ascorbate peroxidase (APX) and of an engineered K+-site mutant of APX that has had its potassium cation binding site removed have been examined by electronic absorption and magnetic circular dichroism (MCD) spectroscopy at 4??°C. Wild-type ferric APX has spectroscopic properties that are very similar to those of ferric cytochrome c peroxidase (CCP) and likely exists primarily as a five-coordinate high-spin heme ligated on the proximal side by a histidine at pH 7. There is also evidence for minority contributions from six-coordinate high- and low-spin species (histidine-water, histidine-hydroxide, and bis-histidine). The K+-site mutant of APX varies considerably in the electronic absorption and MCD spectra in both the ferric and ferrous states when compared with spectra of the wild-type APX. The electronic absorption and MCD spectra of the engineered K+-site APX mutant are essentially identical to those of cytochrome b 5, a known bis-imidazole (histidine) ligated heme system. It therefore appears that the K+-site mutant of APX has undergone a conformational change to yield a bis-histidine coordination structure in both the ferric and ferrous oxidation states at neutral pH. This conformational change is the result of mutagenesis of the protein to remove the K+-binding site which is located ~8?Å from the peroxide binding pocket. Thus, mutations of protein residues on the proximal side of the heme cause changes in iron ligation on the distal side.  相似文献   

20.
When ferrous iron and sulfur were supplied, cells of T. ferrooxidans in a well-aerated medium started growth by oxidizing ferrous iron. After ferrous iron depletion a lagphase followed before sulfur oxidation started. During sulfur oxidation at pH-values below 1.3 (±0,2) the ferrous iron concentration increased again, although the oxygen saturation of the medium amounted to more than 95%. The number of viable cells did not increase. Thus resting cells of T. ferrooxidans, which are oxidizing sulfur to maintain their proton balance, reduce ferric to ferrous iron. The ferrous iron-oxidizing system seemed to be inhibited at pH-values below 1.3. At a pH-value of 1.8 the ferrous iron was reoxidized at once. A scheme for the linkage of iron- and sulfur metabolism is discussed.  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