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 15 毫秒
1.
Marinobacter sp. MnI7-9, a deep-sea manganese [Mn(II)]-oxidizing bacterium isolated from the Indian Ocean, showed a high resistance to Mn(II) and other metals or metalloids and high Mn(II) oxidation/removal abilities. This strain was able to grow well when the Mn(II) concentration reached up to 10 mM, and at that concentration, 76.4% of the added Mn(II) was oxidized and 23.4% of the Mn(II) was adsorbed by the generated biogenic Mn oxides (total 99.9% Mn removal). Scanning electron microscope observation and X-ray diffraction analysis showed that the biogenic Mn oxides were in stick shapes, adhered to the cell surface, and contained two typical crystal structures of γ-MnOOH and δ-MnO2. In addition, the biogenic Mn oxides generated by strain MnI7-9 showed abilities to oxidize the highly toxic As(III) to the less toxic As(V), in both co-culture and after-collection systems. In the co-culture system containing 10 mM Mn(II) and 55 μM As(III), the maximum percentage of As(III) oxidation was 83.5%. In the after-collection system using the generated biogenic Mn oxides, 90% of the As(III) was oxidized into As(V), and the concentration of As(III) decreased from 55.02 to 5.55 μM. This study demonstrates the effective bioremediation by a deep-sea Mn(II)-oxidizing bacterium for the treatment of As-containing water and increases the knowledge of deep-sea bacterial Mn(II) oxidation mechanisms. Supplemental materials are available for this article. Go to the publisher's online edition of Geomicrobiology Journal to view the supplemental file.  相似文献   

2.
Mn-rich filamentous particles present in stratified water environments are considered bacteriogenic; however, little is known about their causative agents. This study investigated the production of these particles by an alphaproteobacterium, Bosea sp. strain BIWAKO-01. Particle formation was promoted in static cultures with slightly viscous medium at pH 6.0?6.3 under low-O2 conditions. The Mn(II) oxidation in cultures was slower in higher O2 concentration. These results suggested that pH and O2 concentration are important factors affecting filamentous Mn particle formation in the Mn(II) oxidizer. Lectin staining followed by fluorescence microscopy revealed the presence of specific carbohydrates in the filamentous structures. In addition, transmission electron microscopy, high angle annular dark field scanning transmission electron microscopy, and electron energy loss spectroscopy revealed the structural and spatial associations of Mn with O and C on a nanometer scale in filaments. The results suggested the occurrence of sheet-type Mn oxide likely due to the catalytic activity in exopolymeric substances including acidic polysaccharides.  相似文献   

3.
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   

4.
Understanding the biogeochemical cycle of the highly toxic element mercury (Hg) is necessary to predict its fate and transport. In this study, we determined that biogenic magnetite isolated from Magnetospirillum gryphiswaldense MSR-1 and Magnetospirillum magnetotacticum MS-1 was capable of reducing inorganic mercury [Hg(II)] to elemental mercury [Hg(0)]. These two magnetotactic bacteria (MTB) lacked mercuric resistance operons in the genomes. However, they revealed high resistance to Hg(II) under atmospheric conditions and an even higher resistance under microaerobic conditions (1% O2 and 99% N2). Neither strain reduced Hg(II) to Hg(0) under atmospheric conditions. However, a slow rate (0.05–0.21 µM·d?1) of Hg(II) loss occurred from late log phase to stationary phase in two MTBs' culture media under microaerobic conditions. Increased Hg(II) entered both cells under microaerobic conditions relative to atmospheric conditions. The majority of Hg(II) was still blocked by the cell membrane. Hg(II) reduction was more effective when biogenic magnetite was extracted out, with or without the magnetosome membrane envelope. When magnetosome membrane was present, 8.55–13.53% of 250 nM Hg(II) was reduced to Hg(0) by 250 mg/L biogenic magnetite suspension within 2 hours. This ratio increased to 55.07–64.70% while magnetosome membrane was removed. We concluded that two MTBs contributed to the reduction of Hg(II) to Hg(0) at a slow rate in vivo. Such reduction was more favorable to occur when biogenic magnetite is released from dead cells. It proposed a new biotic pathway for the formation of Hg(0) in aquatic systems.  相似文献   

5.

Natural manganese oxide nanoparticles and grain coatings profoundly impact contaminant cycling in the environment through their ability to degrade organic compounds and sequester metal ions. Previous studies of biogenic manganese oxides have shown that the interlayer cation may have an important effect on the resulting oxide structure. The effect of Na and Ca ions was investigated to determine their fundamental roles in the stabilization of the phyllomanganate biooxide structure, its unit cell symmetry, and order/disorder relations. Biogenic oxides were created by incubating Mn(II) with spores of the marine Bacillus sp., strain SG-1 and the resulting oxide structures examined using X-ray absorption spectroscopy and X-ray diffraction to determine the short-range and long-range atomic structure. Phyllomanganates were observed exclusively, with differing degrees of layer stacking disorder, degree of crystallinity, and layer symmetry, depending on the cation present. In general, Ca was found to promote biooxide long-range order. We conclude that the presence of Ca in these oxides will confer greater stability to these bacteriogenic manganese bioxodes.  相似文献   

6.
In this study, we report a bacterium, Achromobacter sp. TY3-4, capable of concurrently removing Mn (II) and Cr (VI) under oxic condition. TY3-4 reduced as much as 2.31?mM of Cr (VI) to Cr (III) in 70?h, and oxidized as much as 20?mM of Mn(II) to Mn oxides in 80?h. When 0.58?mM Cr (VI) and 10?mM Mn(II) were present together, both Cr(VI) and Mn(II) were completely removed by TY3-4 and the generated precipitates are MnIIIOOH, MnIII,IV3O4, MnIVO2 and CrIII(OH)3. Experiments also show that both biosroption and bioreduction of Mn(II) are the driving forces for Mn(II) removal, whereas bioreduction of Cr(VI) is the driving force for Cr(VI) removal. On the basis of these results, a possible reaction was proposed that TY3-4 concurrently reduces Cr(VI) and oxidizes Mn(II). This study is fundamental for Mn and Cr cycles. The strain shows potential for practical application.  相似文献   

7.

Biomineralization in heterogeneous aqueous systems results from a complex association between pre-existing surfaces, bacterial cells, extracellular biomacromolecules, and neoformed precipitates. Fourier transform infrared (FTIR) spectroscopy was used in several complementary sample introduction modes (attenuated total reflectance [ATR], diffuse reflectance [DRIFT], and transmission) to investigate the processes of cell adhesion, biofilm growth, and biological Mn-oxidation by Pseudomonas putida strain GB-1. Distinct differences in the adhesive properties of GB-1 were observed upon Mn oxidation. No adhesion to the ZnSe crystal surface was observed for planktonic GB-1 cells coated with biogenic MnO x , whereas cell adhesion was extensive and a GB-1 biofilm was readily grown on ZnSe, CdTe, and Ge crystals prior to Mn-oxidation. IR peak intensity ratios reveal changes in biomolecular (carbohydrate, phosphate, and protein) composition during biologically catalyzed Mn-oxidation. In situ monitoring via ATR-FTIR of an active GB-1 biofilm and DRIFT data revealed an increase in extracellular protein (amide I and II) during Mn(II) oxidation, whereas transmission mode measurements suggest an overall increase in carbohydrate and phosphate moieties. The FTIR spectrum of biogenic Mn oxide comprises Mn-O stretching vibrations characteristic of various known Mn oxides (e.g., “acid” birnessite, romanechite, todorokite), but it is not identical to known synthetic solids, possibly because of solid-phase incorporation of biomolecular constituents. The results suggest that, when biogenic MnO x accumulates on the surfaces of planktonic cells, adhesion of the bacteria to other negatively charged surfaces is hindered via blocking of surficial proteins.  相似文献   

8.
Abstract

Reductive immobilization of Cr(VI) has been widely explored as a cost-effective approach for Cr-contaminated site remediation. In soils containing manganese oxides, however, the immobilized form of chromium, i.e., Cr(III), could potentially be reoxidized. In this study, batch experiments were conducted to assess whether there were any microbial processes that could accelerate Cr(III) oxidation in aerobic, manganese-containing systems. The results showed that in the presence of at least one species of manganese oxidizers, Pseudomonas putida, Cr(III) oxidation took place at low concentrations of Cr(III). About 30–50% of added Cr(III) (10–200 μ M) was oxidized to Cr(VI) within five days in the systems with P. putida and biogenic Mn oxides. The rate of Cr(III) oxidation was approximately proportional to the initial concentration of Cr(III) up to 100 μ M, but the growth of P. putida was partially inhibited by Cr(III) at 200 μ M and totally stopped when it reached 500 μ M. Cr(III) oxidation was dependent upon the biogenic formation of Mn oxides, though the oxidation rate was not directly proportional to the amount of Mn oxides formed. Chromium(III) oxidation took place through a catalytic pathway, in which the microbes mediated Mn(II) oxidation to form Mn-oxides, and Cr(III) was subsequently oxidized by the biogenic Mn-oxides.  相似文献   

9.
Cobalt(II) oxidation in aquatic environments has been shown to be linked to Mn(II) oxidation, a process primarily mediated by bacteria. This work examines the oxidation of Co(II) by the spore-forming marine Mn(II)-oxidizing bacterium Bacillus sp. strain SG-1, which enzymatically catalyzes the formation of reactive nanoparticulate Mn(IV) oxides. Preparations of these spores were incubated with radiotracers and various amounts of Co(II) and Mn(II), and the rates of Mn(II) and Co(II) oxidation were measured. Inhibition of Mn(II) oxidation by Co(II) and inhibition of Co(II) oxidation by Mn(II) were both found to be competitive. However, from both radiotracer experiments and X-ray spectroscopic measurements, no Co(II) oxidation occurred in the complete absence of Mn(II), suggesting that the Co(II) oxidation observed in these cultures is indirect and that a previous report of enzymatic Co(II) oxidation may have been due to very low levels of contaminating Mn. Our results indicate that the mechanism by which SG-1 oxidizes Co(II) is through the production of the reactive nanoparticulate Mn oxide.  相似文献   

10.
Abstract

Mn oxides have long been considered the primary environmental oxidant of Cr(III), however, since most of the reactive Mn oxides in the environment are believed to be of biological origin, microorganisms may indirectly mediate Cr(III) oxidation and accelerate the rate over that seen in purely abiotic systems. In this study, we examined the ability of the Mn(II)-oxidizing bacterium, Pseudomonas putida strain GB-1, to oxidize Cr(III). Our results show that GB-1 cannot oxidize Cr(III) directly, but that in the presence of Mn(II), Cr(III) can be rapidly and completely oxidized. Growth studies suggest that in growth medium with few organics the resulting Cr(VI) may be less toxic to P. putida GB-1 than Cr(III), which is generally considered less hazardous. In addition, Cr(III) present during the growth of P. putida GB-1 appeared to cause iron stress as determined by the production of the fluorescent siderophore pyoverdine. When stressed by Fe limitation or Cr(III) toxicity, Mn(II) oxidation by GB-1 is inhibited.  相似文献   

11.
Ascomycetes that can deposit Mn(III, IV) oxides are widespread in aquatic and soil environments, yet the mechanism(s) involved in Mn oxide deposition remains unclear. A Mn(II)-oxidizing ascomycete, Acremonium sp. strain KR21-2, produced a Mn oxide phase with filamentous nanostructures. X-ray absorption near-edge structure (XANES) spectroscopy showed that the Mn phase was primarily Mn(IV). We purified to homogeneity a laccase-like enzyme with Mn(II) oxidase activity from cultures of strain KR21-2. The purified enzyme oxidized Mn(II) to yield suspended Mn particles; XANES spectra indicated that Mn(II) had been converted to Mn(IV). The pH optimum for Mn(II) oxidation was 7.0, and the apparent half-saturation constant was 0.20 mM. The enzyme oxidized ABTS [2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonic acid)] (pH optimum, 5.5; Km, 1.2 mM) and contained two copper atoms per molecule. Moreover, the N-terminal amino acid sequence (residues 3 to 25) was 61% identical with the corresponding sequence of an Acremonium polyphenol oxidase and 57% identical with that of a Myrothecium bilirubin oxidase. These results provide the first evidence that a fungal multicopper oxidase can convert Mn(II) to Mn(IV) oxide. The present study reinforces the notion of the contribution of multicopper oxidase to microbially mediated precipitation of Mn oxides and suggests that Acremonium sp. strain KR21-2 is a good model for understanding the oxidation of Mn in diverse ascomycetes.  相似文献   

12.
A new bacterial strain isolated from soil consumed nitric oxide (NO) under oxic conditions by oxidation to nitrate. Phenotypic and phylogenetic characterization of the new strain PS88 showed that it represents a previously unknown species of the genus Pseudomonas, closely related to Pseudomonas fluorescens and Pseudomonas putida. The heterotrophic, obligately aerobic strain PS88 was not able to denitrify or nitrify; however, strain PS88 oxidized NO to nitrate. NO was not reduced to nitrous oxide (N2O). Nitrogen dioxide (NO2) and nitrite (NO2 ) as possible intermediates of NO oxidation to nitrate (NO3 ) could not be detected. NO oxidation was inhibited under anoxic conditions and by high osmolarity, but not by nitrite. NO oxidation activity was inhibited by addition of formaldehyde, HgCl2, and antimycin, and by autoclaving or disintegrating the cells, indicating that the process was enzyme-mediated. However, the mechanism remains unclear. A stepwise oxidation at a metalloenzyme and a radical mechanism are discussed. NO oxidation in strain PS88 seems to be a detoxification or a co-oxidation mechanism, rather than an energy-yielding process. Received: 15 November 1995 / Accepted: 24 February 1996  相似文献   

13.
Anaerobic Fe(II) Oxidizing Denitrifiers (AFODN), a type of newly found Fe(II)-oxidizing bacteria, play an important role in iron and nitrogen cycling. In the present study, a novel AFODN strain PXL1 was isolated from anaerobic activated sludge. Phylogenetic analysis of 16S rRNA gene sequence revealed similarity between this strain and Citrobactor freundii. The strain reduced 30% of nitrate and oxidized 85% of Fe(II) over 72 h with an initial Fe(II) concentration of 3.4 mM and nitrate concentration of 9.5 mM. Oxidation of iron was dependent on the reduction of nitrate to nitrite in the absence of other electron donors or acceptors. Nitrate reduction and Fe(II) oxidation followed first-order reaction kinetics. Iron oxides accumulated in the culture were analyzed by Fourier transform infrared (FTIR) spectroscopy, X-ray diffraction (XRD) spectroscopy and scanning electron microscopy-energy dispersive X-ray spectroscopy (SEM-EDS). The strain recovered deposited oxidized Fe in the form of amorphous Fe oxides.  相似文献   

14.
Oxidation of C1 compounds by Pseudomonas sp. MS   总被引:4,自引:2,他引:4       下载免费PDF全文
Pseudomonas sp. MS is capable of growth on a number of compounds containing only C1 groups. They include trimethylsulphonium salts, methylamine, dimethylamine and trimethylamine. Although formaldehyde and formate will not support growth they are rapidly oxidized by intact cells. Methanol neither supports growth nor is oxidized. A particulate fraction of the cell oxidizes methylamine to carbon dioxide in the absence of any external electron acceptor. Formaldehyde and formate are more slowly oxidized to carbon dioxide by the particulate fraction, although they do not appear to be free intermediates in the oxidation of methylamine. Soluble NAD-linked formaldehyde dehydrogenase and formate dehydrogenase are also present. The particulate methylamine oxidase is induced by growth on methylamine, dimethylamine and trimethylamine, whereas the soluble formaldehyde dehydrogenase and formate dehydrogenase are induced by trimethylsulphonium nitrate as well as the aforementioned amines.  相似文献   

15.
Recently, manganese oxides have been considered in the environmental remediation, MRI diagnosis and drug and pharmaceutical industries. Different numbers of physicochemical and biological methods have been reported for the preparation of nanoscale manganese oxides. Although manganese oxide biogenesis by bacterial species has been recognized as the major Mn-oxidizing agent in nature, in this research, for first time, we demonstrated the process which used to produce bixbyite-like Mn(2)O(3) nanoparticles by isolated aerobic bacterium from Persian Gulf water. The 16SRNA sequencing showed that this isolate belong to a gram-negative Acinetobacter which produced nano Mn-oxide crystal particle. Characterization of complement morphology, size and chemical structure of these particles were determined by TEM, SEM, EDAX, XRD and FTIR. The data showed that this bacterium could produce nanosized extracellular bixbyite-like Mn(2)O(3) which depend on enzymatic pathway.  相似文献   

16.
从恒化富集培养物中分离到一株产肌酸酶的菌株WB1,通过对该菌的形态学、生理生化特性、G+C mol%及16S rDNA序列分析,表明该菌为一株副球菌(Paracoccus sp.)。对菌株WB1产酶发酵条件的研究表明,该菌除了产生肌酸酶外还产生肌氨酸脱氢酶,但不产生肌酸酐酶,也不能利用肌酸酐。肌酸酶可以被诱导物,如肌氨酸、肌酸、和氯化胆碱诱导产生。葡萄糖等易用碳源的存在对肌酸酶的合成无代谢产物阻遏作用。该酶的分子量为48kD,最适反应pH为7.0~8.5,pH稳定范围在6.0~9.5之间;其最适反应温度在35℃~40℃之间,在45℃以下是热稳定的; 37℃时以肌酸为底物,酶的Km值为24.6mmol/L;Cu2+、Hg2+和Ag+对酶活性有强烈的抑制作用。  相似文献   

17.
Pseudomonas sp. strain NGK1, a soil bacterium isolated by naphthalene enrichment from biological waste effluent treatment, capable of utilizing 2-methylnaphthalene as sole source of carbon and energy. To deduce the pathway for biodegradation of 2-methylnaphthalene, metabolites were isolated from the spent medium and identified by thin-layer chromatography and high-performance liquid chromatography. The characterization of purified metabolites, oxygen uptake studies, and enzyme activities revealed that the strain degrades 2-methylnaphthalene through more than one pathway. The growth of the bacterium, utilization of 2-methylnaphthalene, and 4-methylsalicylate accumulation by Pseudomonas sp. strain NGK1 were studied at various incubation periods. Received: 20 March 2001 / Accepted: 25 April 2001  相似文献   

18.
A bacterial strain, designated Pseudomonas sp. strain DCA1, was isolated from a 1,2-dichloroethane (DCA)-degrading biofilm. Strain DCA1 utilizes DCA as the sole carbon and energy source and does not require additional organic nutrients, such as vitamins, for optimal growth. The affinity of strain DCA1 for DCA is very high, with a Km value below the detection limit of 0.5 μM. Instead of a hydrolytic dehalogenation, as in other DCA utilizers, the first step in DCA degradation in strain DCA1 is an oxidation reaction. Oxygen and NAD(P)H are required for this initial step. Propene was converted to 1,2-epoxypropane by DCA-grown cells and competitively inhibited DCA degradation. We concluded that a monooxygenase is responsible for the first step in DCA degradation in strain DCA1. Oxidation of DCA probably results in the formation of the unstable intermediate 1,2-dichloroethanol, which spontaneously releases chloride, yielding chloroacetaldehyde. The DCA degradation pathway in strain DCA1 proceeds from chloroacetaldehyde via chloroacetic acid and presumably glycolic acid, which is similar to degradation routes observed in other DCA-utilizing bacteria.  相似文献   

19.
20.
The possibility of using two kinds of sorghum as raw materials in consolidated bioprocessing bioethanol production using Flammulina velutipes was investigated. Enzymatic saccharification of sweet sorghum was not as high as in brown mid-rib (bmr) mutated sorghum, but the amount of ethanol production was higher. Ethanol production from bmr mutated sorghum significantly increased when saccharification enzymes were added to the culture.  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