首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A new method to determine microbial (bacterial and fungal) activity in various freshwater habitats is described. Based on microbial reduction of dimethyl sulfoxide (DMSO) to dimethyl sulfide (DMS), our DMSO reduction method allows measurement of the respiratory activity in interstitial water, as well as in the water column. DMSO is added to water samples at a concentration (0.75% [vol/vol] or 106 mM) high enough to compete with other naturally occurring electron acceptors, as determined with oxygen and nitrate, without stimulating or inhibiting microbial activity. Addition of NaN(3), KCN, and formaldehyde, as well as autoclaving, inhibited the production of DMS, which proves that the reduction of DMSO is a biotic process. DMSO reduction is readily detectable via the formation of DMS even at low microbial activities. All water samples showed significant DMSO reduction over several hours. Microbially reduced DMSO is recovered in the form of DMS from water samples by a purge and trap system and is quantified by gas chromatography and detection with a flame photometric detector. The DMSO reduction method was compared with other methods commonly used for assessment of microbial activity. DMSO reduction activity correlated well with bacterial production in predator-free batch cultures. Cell-production-specific DMSO reduction rates did not differ significantly in batch cultures with different nutrient regimes but were different in different growth phases. Overall, a cell-production-specific DMSO reduction rate of 1.26 x 10(-17) +/- 0. 12 x 10(-17) mol of DMS per produced cell (mean +/- standard error; R(2) = 0.78) was calculated. We suggest that the relationship of DMSO reduction rates to thymidine and leucine incorporation is linear (the R(2) values ranged from 0.783 to 0.944), whereas there is an exponential relationship between DMSO reduction rates and glucose uptake, as well as incorporation (the R(2) values ranged from 0.821 to 0.931). Based on our results, we conclude that the DMSO reduction method is a nonradioactive alternative to other methods commonly used to assess microbial activity.  相似文献   

2.
Dimethyl sulfide (DMS) is a sulfur compound of importance for the organoleptic properties of beer, especially some lager beers. Synthesis of DMS during beer production occurs partly during wort production and partly during fermentation. Methionine sulfoxide reductases are the enzymes responsible for reduction of oxidized cellular methionines. These enzymes have been suggested to be able to reduce dimethyl sulfoxide (DMSO) as well, with DMS as the product. A gene for an enzymatic activity leading to methionine sulfoxide reduction in Saccharomyces yeast was recently identified. We confirmed that the Saccharomyces cerevisiae open reading frame YER042w appears to encode a methionine sulfoxide reductase, and propose the name MXR1 for the gene. We found that Mxr1p catalyzes reduction of DMSO to DMS and that an mxr1 disruption mutant cannot reduce DMSO to DMS. Mutant strains appear to have unchanged fitness under several laboratory conditions, and in this paper I hypothesize that disruption of MXR1 in brewing yeasts would neutralize the contribution of the yeast to the DMS content in beer.  相似文献   

3.
Summary Pseudomonas acidovorans DMR-11, capable of oxidizing dimethyl sulfide (DMS), was isolated from peat biofilter. DMS as a sole carbon or energy source was not degraded, but it was co-degraded in the medium containing organic carbon sources. The removal rate of DMS in heat-treated glucose medium was 1.12×10–17 mole/h cell at 30 °C. Dimethyl sulfoxide (DMSO) was the only product of DMS oxidation and was formed stoichiometrically. DMS was reversibly evolved in excess of DMSO. The cell free extract of strain DMR-11 oxidized DMS in presence of NADPH.  相似文献   

4.
The kinetics for the reduction of sulfate alone and for concurrent uranium [U(VI)] and sulfate reduction, by mixed and pure cultures of sulfate-reducing bacteria (SRB) at 21 ± 3°C were studied. The mixed culture contained the SRB Desulfovibrio vulgaris along with a Clostridium sp. determined via 16S ribosomal DNA analysis. The pure culture was Desulfovibrio desulfuricans (ATCC 7757). A zero-order model best fit the data for the reduction of sulfate from 0.1 to 10 mM. A lag time occurred below cell concentrations of 0.1 mg (dry weight) of cells/ml. For the mixed culture, average values for the maximum specific reaction rate, Vmax, ranged from 2.4 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1) at 0.25 mM sulfate to 5.0 ± 1.1 μmol of sulfate/mg (dry weight) of SRB · h−1 at 10 mM sulfate (average cell concentration, 0.52 mg [dry weight]/ml). For the pure culture, Vmax was 1.6 ± 0.2 μmol of sulfate/mg (dry weight) of SRB · h−1 at 1 mM sulfate (0.29 mg [dry weight] of cells/ml). When both electron acceptors were present, sulfate reduction remained zero order for both cultures, while uranium reduction was first order, with rate constants of 0.071 ± 0.003 mg (dry weight) of cells/ml · min−1 for the mixed culture and 0.137 ± 0.016 mg (dry weight) of cells/ml · min−1 (U0 = 1 mM) for the D. desulfuricans culture. Both cultures exhibited a faster rate of uranium reduction in the presence of sulfate and no lag time until the onset of U reduction in contrast to U alone. This kinetics information can be used to design an SRB-dominated biotreatment scheme for the removal of U(VI) from an aqueous source.  相似文献   

5.
The molecular weight of the large RNA of Sendai virus has been determined by sedimentation analysis in sucrose gradients containing 99% dimethyl sulfoxide (DMSO) to be 2.3 × 106. Sendai RNA recovered from 99% DMSO was found to cosediment with nondenatured Sendai RNA at 46 to 48s in ordinary sucrose gradients. The molecular weight value of 2.3 × 106 is considerably smaller than the estimates of 6 × 106 to 7 × 106 determined under nondenaturing conditions, suggesting a unique structure for Sendai RNA.  相似文献   

6.
Enrichment cultures of phototrophic purple bacteria rapidly oxidized up to 10 mM dimethyl sulfide (DMS) to dimethyl sulfoxide (DMSO). DMSO was qualitatively identified by proton nuclear magnetic resonance. By using a biological assay, DMSO was always quantitatively recovered from the culture media. DMS oxidation was not detected in cultures incubated in the dark, and it was slow in cultures exposed to full daylight. Under optimal conditions, the second-order rate constant for DMS oxidation was 6 day−1 mg of protein−1 ml−1. The rate constant was reduced in the presence of high concentration of sulfide (>1 mM), but was not affected by the addition of acetate. DMS was also oxidized to DMSO by a pure strain (tentatively identified as a Thiocystis sp.) isolated from the enrichment cultures. DMS supported growth of the enrichment cultures and of the pure strain by serving as an electron source for photosynthesis. A determination of the amount of protein produced in the cultures and an estimation of the electron balance suggested that the two electrons liberated during the oxidation of DMS to DMSO were quantitatively used to reduce carbon dioxide to biomass. The oxidation of DMS by phototrophic purple bacteria may be an important source of DMSO detected in anaerobic ponds and marshes.  相似文献   

7.
(R,S)-[1-14C]3-Hydroxy eicosanoyl-coenzyme A (CoA) has been chemically synthesized to study the 3-hydroxy acyl-CoA dehydratase involved in the acyl-CoA elongase of etiolated leek (Allium porrum L.) seedling microsomes. 3-Hydroxy eicosanoyl-CoA (3-OH C20:0-CoA) dehydration led to the formation of (E)-2,3 eicosanoyl-CoA, which has been characterized. Our kinetic studies have determined the optimal conditions of the dehydration and also resolved the stereospecificity requirement of the dehydratase for (R)-3-OH C20:0-CoA. Isotopic dilution experiments showed that 3-hydroxy acyl-CoA dehydratase had a marked preference for (R)-3-OH C20:0-CoA. Moreover, the very-long-chain synthesis using (R)-3-OH C20:0-CoA isomer and [2-14C]malonyl-CoA was higher than that using the (S) isomer, whatever the malonyl-CoA and the 3-OH C20:0-CoA concentrations. We have also used [1-14C]3-OH C20:0-CoA to investigate the reductant requirement of the enoyl-CoA reductase of the acyl-CoA elongase complex. In the presence of NADPH, [1-14C]3-OH C20:0-CoA conversion was stimulated. Aside from the product of dehydration, i.e. (E)-2,3 eicosanoyl-CoA, we detected eicosanoyl-CoA resulting from the reduction of (E)-2,3 eicosanoyl-CoA. When we replaced NADPH with NADH, the eicosanoyl-CoA was 8- to 10-fold less abundant. Finally, in the presence of malonyl-CoA and NADPH or NADH, [1-14C]3-OH C20:0-CoA led to the synthesis of very-long-chain fatty acids. This synthesis was measured using [1-14C]3-OH C20:0-CoA and malonyl-CoA or (E)-2,3 eicosanoyl-CoA and [2-14C]malonyl-CoA. In both conditions and in the presence of NADPH, the acyl-CoA elongation activity was about 60 nmol mg−1 h−1, which is the highest ever reported for a plant system.  相似文献   

8.
We investigated the feasibility of bioremediation as a treatment option for a chronically diesel-oil-polluted soil in an alpine glacier area at an altitude of 2,875 m above sea level. To examine the efficiencies of natural attenuation and biostimulation, we used field-incubated lysimeters (mesocosms) with unfertilized and fertilized (N-P-K) soil. For three summer seasons (July 1997 to September 1999), we monitored changes in hydrocarbon concentrations in soil and soil leachate and the accompanying changes in soil microbial counts and activity. A significant reduction in the diesel oil level could be achieved. At the end of the third summer season (after 780 days), the initial level of contamination (2,612 ± 70 μg of hydrocarbons g [dry weight] of soil−1) was reduced by (50 ± 4)% and (70 ± 2)% in the unfertilized and fertilized soil, respectively. Nonetheless, the residual levels of contamination (1,296 ± 110 and 774 ± 52 μg of hydrocarbons g [dry weight] of soil−1 in the unfertilized and fertilized soil, respectively) were still high. Most of the hydrocarbon loss occurred during the first summer season ([42 ± 6]% loss) in the fertilized soil and during the second summer season ([41 ± 4]% loss) in the unfertilized soil. In the fertilized soil, all biological parameters (microbial numbers, soil respiration, catalase and lipase activities) were significantly enhanced and correlated significantly with each other, as well as with the residual hydrocarbon concentration, pointing to the importance of biodegradation. The effect of biostimulation of the indigenous soil microorganisms declined with time. The microbial activities in the unfertilized soil fluctuated around background levels during the whole study.  相似文献   

9.
The photosynthetic bacterium, Rhodopseudomonas capsulata, could be cultured anaerobically in the absence of light on a synthetic medium with glucose as the carbon source only when dimethyl sulfoxide (DMSO) was added. The extent of growth was proportional to both DMSO and glucose concentrations. Optimal growth was achieved with 20 mm DMSO and 0.25% glucose. Under the best conditions, cells divided with a doubling time of 12 h. Pyruvate also supported the anaerobic dark growth of R. capsulata when DMSO was present. R. capsulata, R. sphaeroides, and R. palustris strains were all able to grow under anaerobic dark conditions with DMSO. Experiments using [14C]DMSO showed that more than 95% of the 14C was converted by cultures of R. capsulata to a volatile compound, identified as dimethyl sulfide (DMS) by gas chromatography, thus demonstrating that DMSO was being reduced to DMS during growth. These results indicate that R. capsulata requires a terminal electron acceptor for anaerobic dark growth and that DMSO can serve that function.  相似文献   

10.
The uptake and degradation of nanomolar levels of [methyl-14C]choline in estuarine water samples and in seawater filtrate cultures composed mainly of natural free-living bacteria was studied. Uptake of [14C]choline exhibited Michaelis-Menten kinetics, with Kt + Sn values of 1.7 to 2.9 nM in filtrate cultures and 1.7 to 4.1 nM in estuarine-water samples. Vmax values ranged from 0.5 to 3.3 nM · h−1. The uptake system for choline in natural microbial assemblages therefore displays very high affinity and appears able to scavenge this compound at the concentrations expected in seawater. Uptake of choline was inhibited by some natural structural analogs and p-chloromercuribenzoate, indicating that the transporter may be multifunctional and may involve a thiol binding site. When 11 nM [14C]choline was added to water samples, a significant fraction (>50%) of the methyl carbon was respired to CO2 in incubations lasting 10 to 53 h. Cells taking up [14C]choline produced [14C]glycine betaine ([14C]GBT), and up to 80% of the radioactivity retained by cells was in the form of GBT, a well-known osmolyte. Alteration of the salinity in filtrate cultures affected the relative proportion of [14C]choline degraded or converted to [14C]GBT, without substantially affecting the total metabolism of choline. Increasing the salinity from 14 to 25 or 35 ppt caused more [14C]GBT to be produced from choline but less 14CO2 to be produced than in the controls. Lowering the salinity to 7 ppt decreased [14C]GBT production and increased 14CO2 production slightly. Intracellular accumulations of [14C]GBT in the salt-stressed cultures were osmotically significant (34 mM). Choline may be used as an energy substrate by estuarine bacteria and may also serve as a precursor of the osmoprotectant GBT, particularly as bacteria are mixed into higher-salinity waters.  相似文献   

11.
Conjugative transfer of the Enterococcus faecalis plasmid pPD1 is activated by cPD1, one of several peptide sex pheromones secreted by plasmid-free recipient cells, and is blocked by a donor-produced peptide inhibitor, iPD1. Using a tritiated pheromone, [3H]cPD1, we investigated how pPD1-harboring donor cells receive these peptide signals. Donor cells rapidly incorporated [3H]cPD1. The cell extract but not the membrane fraction of the donor strain exhibited significant [3H]cPD1-binding activity. On the basis of these data and those of tracer studies, it was demonstrated that cPD1 was internalized, where it bound to a high-molecular-weight compound. The cell extract of a strain carrying the traA-bearing multicopy plasmid (pDLHH21) also exhibited high [3H]cPD1-binding activity. A recombinant TraA exhibited a dissociation constant of 0.49 ± 0.08 nM against [3H]cPD1. iPD1 competitively inhibited [3H]cPD1 binding to TraA, whereas pheromones and inhibitors relating to other plasmid systems did not. These results show that TraA is a specific intracellular receptor for cPD1 and that iPD1 acts as an antagonist for TraA. A strain carrying the traC-bearing multicopy plasmid (pDLES23) exhibited significant [3H]cPD1-binding activity. A strain carrying traC-disrupted pPD1 (pAM351CM) exhibited lower [3H]cPD1-binding activity as well as lower sensitivity to cPD1 than a wild-type donor strain. Some of the other pheromones and inhibitors inhibited [3H]cPD1 binding to the traC transformant like cPD1 and iPD1 did. These results show that TraC, as an extracellular less-specific pheromone-binding protein, supports donor cells to receive cPD1.  相似文献   

12.
The reduction of 12-oxophytodienoic acid (OPDA) to 3-oxo-2(2′[Z]-pentenyl)-cyclopentane-1-octanoic acid is catalyzed by 12-oxophytodienoate-10,11-reductase (OPR). Analysis of the isomer preference of OPR has indicated that the activity is composed of two isoenzymes exhibiting different stereoselectivities. The two isoforms of OPR have been separated, using protein extracts of Rock Harlequin (Corydalis sempervirens) as the starting material. OPRI, the enzyme reported earlier from the same species and corresponding to the cloned OPR from Arabidopsis, utilized 9R,13R-OPDA >> 9S,13R-OPDA but not the 13S-configured isomers, whereas the new activity, OPRII, effectively reduced all four OPDA isomers, including the natural 9S,13S-OPDA (cis-[+]-OPDA). OPRII activity is characterized in detail. The enzyme's enzymatic, biochemical, and immunological properties prove that it is a close relative of OPRI. The roles of OPRI and OPRII in octadecanoid biology are discussed.  相似文献   

13.
A total of 215 Streptomyces strains were screened for their capacity to regio- and stereoselectively hydroxylate β- and/or α-ionone to the respective 3-hydroxy derivatives. With β-ionone as the substrate, 15 strains showed little conversion to 4-hydroxy- and none showed conversion to the 3-hydroxy product as desired. Among these 15 Streptomyces strains, S. fradiae Tü 27, S. arenae Tü 495, S. griseus ATCC 13273, S. violaceoniger Tü 38, and S. antibioticus Tü 4 and Tü 46 converted α-ionone to 3-hydroxy-α-ionone with significantly higher hydroxylation activity compared to that of β-ionone. Hydroxylation of racemic α-ionone [(6R)-(−)/(6S)-(+)] resulted in the exclusive formation of only the two enantiomers (3R,6R)- and (3S,6S)-hydroxy-α-ionone. Thus, the enzymatic hydroxylation of α-ionone by the Streptomyces strains tested proceeds with both high regio- and stereoselectivity.  相似文献   

14.
In extracts of immature soybean (Glycine max [L.] Merr.) seeds inositol tetrakisphosphate was formed from [3H]inositol 1,3,4-trisphosphate but not from [3H]inositol 1,4,5-trisphosphate. Inositol 1,3,4-trisphosphate kinase was purified to a specific activity of 3.55 min−1 mg−1 by polyethylenimine clarification and anion-exchange chromatography. The partially purified enzyme converted [3H]inositol 1,3,4-trisphosphate to inositol 1,3,4,5-tetrakisphosphate as the major product and inositol 1,3,4,6- and/or 1,2,3,4-tetrakisphosphate as the minor product. Subsequent experiments revealed a separate inositol 1,3,4,5-tetrakisphosphate 6-kinase activity, which could link these enzymes to inositol hexakisphosphate synthesis via the previously reported inositol 1,3,4,5,6-pentakisphosphate 2-kinase. The apparent Km values for inositol 1,3,4-trisphosphate kinase were 200 ± 0 nm for inositol 1,3,4-trisphosphate and 171 ± 4 μm for ATP, and the reaction was not reversible. The kinetics were such that no activity could be detected using unlabeled inositol 1,3,4-trisphosphate and [γ-32P]ATP, which suggested that other kinases may have been observed when less purified fractions were incubated with radiolabeled ATP. Inositol 1,3,4-trisphosphate kinase was nonspecifically inhibited more than 80% by various inositol polyphosphates at a concentration of 100 μm.  相似文献   

15.
A key issue in the nucleotide excision repair (NER) of bulky carcinogen–DNA adducts is the ability of the NER machinery to recognize and repair certain adducts while failing to repair others. Unrepaired adducts can survive to cause mutations that initiate the carcinogenic process. Benzo[c]phenanthrene (B[c]Ph), a representative fjord region polycyclic aromatic hydrocarbon, can be metabolically activated to the enantiomeric benzo[c]phenanthrene diol epoxides (B[c]PhDEs), (+)-(1S,2R,3R,4S)-3,4- dihydroxy-1,2-epoxy-1,2,3,4-tetrahydrobenzo[c]phe nanthrene and the corresponding (–)-(1R,2S,3S,4R) isomer. These react predominantly with adenine residues in DNA to produce the stereoisomeric 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts. Duplexes containing the 1R (+) or 1S (–) B[c]Ph-dA adduct in codon 61 of the human N-ras mutational hotspot sequence CA*A, with B[c]Ph modification at A*, are not repaired by the human NER system. However, the analogous stereoisomeric DNA adducts of the bay region benzo[a]pyrene diol epoxide (B[a]PDE), 10S (+)- and 10R (–)-trans-anti-B[a]P-N6-dA, are repaired in the same base sequence. In order to elucidate structural and thermodynamic origins of this phenomenon, we have carried out a 2 ns molecular dynamics simulation for the 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts in an 11mer duplex containing the human N-ras codon 61 sequence, and compared these results with our previous study of the B[a]P-dA adducts in the same sequence. The molecular mechanics Poisson– Boltzmann surface area (MM-PBSA) method was applied to calculate the free energies of the pair of stereoisomeric B[c]Ph-dA adducts, and a detailed structural analysis was carried out. The different repair susceptibilities of the B[a]P-dA adducts and the B[c]Ph-dA adducts can be attributed to different degrees of distortion, stemming from combined effects of differences in the quality of Watson–Crick hydrogen bonding, unwinding, stretching and helix backbone perturbations. These differences are due to the different intrinsic topologies of the rigid, planar bay region adducts versus the twisted, sterically hindered fjord region adducts.  相似文献   

16.
The dissimilatory metal reducing bacterium Shewanella oneidensis MR-1, known for its capacity of reducing iron and manganese oxides, has great environmental impacts. The iron oxides reducing process is affected by the coexistence of alternative electron acceptors in the environment, while investigation into it is limited so far. In this work, the impact of dimethyl sulphoxide (DMSO), a ubiquitous chemical in marine environment, on the reduction of hydrous ferric oxide (HFO) by S. oneidensis MR-1 was investigated. Results show that DMSO promoted HFO reduction by both wild type and ΔdmsE, but had no effect on the HFO reduction by ΔdmsB, indicating that such a promotion was dependent on the DMSO respiration. With the DMSO dosing, the levels of extracellular flavins and omcA expression were significantly increased in WT and further increased in ΔdmsE. Bioelectrochemical analysis show that DMSO also promoted the extracellular electron transfer of WT and ΔdmsE. These results demonstrate that DMSO could stimulate the HFO reduction through metabolic and genetic regulation in S. oneidensis MR-1, rather than compete for electrons with HFO. This may provide a potential respiratory pathway to enhance the microbial electron flows for environmental and engineering applications.  相似文献   

17.
Escherichia coli can use dimethyl sulfoxide (DMSO) as an electron acceptor during anaerobic growth on the oxidizable substrate, glycerol. During growth, the DMSO is reduced to dimethyl sulfide (DMS). For the reduction of DMSO, NADH, formate, lactate, reduced benzyl viologen, reduced methyl viologen, and dithionite can serve as electron donors. The terminal reductase and the dehydrogenases linking the various electron donors to the electron transport chain were found to be membrane bound. Chlorate-resistant mutants (chl) were unable to grow and reduce DMSO. However, in the case of thechlD mutant, growth and DMSO reduction can be restored by growth in the presence of high concentrations of molybdate. Mutants ofE. coli blocked in menaquinone (vitamin K2) biosynthesis—menB, menC, andmenD—were unable to grow with DMSO as an electron acceptor, even though the terminal reductase is present in these mutants. Both growth and DMSO reduction could be restored in these mutants by growth in the presence of the menaquinone intermediates,o-succinylbenzoate and 1,4-dihydroxy-2-naphthoate, depending on the metabolic block of the mutant. Thus menaquinone is involved in electron transport during DMSO reduction.  相似文献   

18.
High precision ultrasonic and densimetric techniques have been used to study the interaction of Ni2+ ions with right-handed poly[d(G-C)]·poly[d(G-C)], poly-[d(A-C)]·poly[d(G-T)] and poly[d(A-T)]·poly[d(A-T)] in 5 mM CsCl, 0.2 mM HEPES, pH 7.5 at 20°C. From these measurements the changes in the apparent molar volume and the apparent molar adiabatic compressibility due to the interaction have been obtained. The volume effects of the binding, calculated per mole of Ni2+ ions, range from 11.7 to 23.9 cm3 mol–1 and the compressibility effects range from 19.3 × 10–4 to 43.1 × 10–4 cm3 mol–1 bar–1. These data are interpreted in terms of dehydration of the polynucleotides and Ni2+ ions, i.e. the release of water molecules from the hydration shells of the molecules. An increase in G+C content gives an increase in volume and compressibility effects, indicating a rise in the extent of dehydration. The dehydration effects of Ni2+ binding to poly[d(G-C)]·poly[d(G-C)] are approximately twice those of poly[d(A-T)]·poly[d(A-T)]. The volume and compressibility effects of Ni2+–EDTA complex formation have also been measured and used as a model system for quantitative estimation. These values revealed that Ni2+ ions can coordinate two atomic groups of poly[d(G-C)]·poly[d(G-C)], while in the case of the Ni2+–poly[d(A-T)]·poly[d(A-T)] complex volume and compressibility effects correspond to one direct or two indirect (through water) contacts.  相似文献   

19.
There are only a few examples of microbial conversion of picric acid (2,4,6-trinitrophenol). None of the organisms that have been described previously is able to use this compound as a sole source of carbon, nitrogen, and energy at high rates. In this study we isolated and characterized a strain, strain CB 22-2, that was able to use picric acid as a sole source of carbon and energy at concentrations up to 40 mM and at rates of 1.6 mmol · h−1 · g (dry weight) of cells−1 in continuous cultures and 920 μmol · h−1 · g (dry weight) of cells−1 in flasks. In addition, this strain was able to use picric acid as a sole source of nitrogen at comparable rates in a nitrogen-free medium. Biochemical characterization and 16S ribosomal DNA analysis revealed that strain CB 22-2 is a Nocardioides sp. strain. High-pressure liquid chromatography and UV-visible light data, the low residual chemical oxygen demand, and the stoichiometric release of 2.9 ± 0.1 mol of nitrite per mol of picric acid provided strong evidence that complete mineralization of picric acid occurred. During transformation, the metabolites detected in the culture supernatant were the [H]-Meisenheimer complexes of picric acid and 2,4-dinitrophenol (H-DNP), as well as 2,4-dinitrophenol. Experiments performed with crude extracts revealed that H-DNP formation indeed is a physiologically relevant step in picric acid metabolism.  相似文献   

20.
This study investigated the biodegradation of high-molecular-weight polycyclic aromatic hydrocarbons (PAHs) in liquid media and soil by bacteria (Stenotrophomonas maltophilia VUN 10,010 and bacterial consortium VUN 10,009) and a fungus (Penicillium janthinellum VUO 10,201) that were isolated from separate creosote- and manufactured-gas plant-contaminated soils. The bacteria could use pyrene as their sole carbon and energy source in a basal salts medium (BSM) and mineralized significant amounts of benzo[a]pyrene cometabolically when pyrene was also present in BSM. P. janthinellum VUO 10,201 could not utilize any high-molecular-weight PAH as sole carbon and energy source but could partially degrade these if cultured in a nutrient broth. Although small amounts of chrysene, benz[a]anthracene, benzo[a]pyrene, and dibenz[a,h]anthracene were degraded by axenic cultures of these isolates in BSM containing a single PAH, such conditions did not support significant microbial growth or PAH mineralization. However, significant degradation of, and microbial growth on, pyrene, chrysene, benz[a]anthracene, benzo[a]pyrene, and dibenz[a,h]anthracene, each as a single PAH in BSM, occurred when P. janthinellum VUO 10,201 and either bacterial consortium VUN 10,009 or S. maltophilia VUN 10,010 were combined in the one culture, i.e., fungal-bacterial cocultures: 25% of the benzo[a]pyrene was mineralized to CO2 by these cocultures over 49 days, accompanied by transient accumulation and disappearance of intermediates detected by high-pressure liquid chromatography. Inoculation of fungal-bacterial cocultures into PAH-contaminated soil resulted in significantly improved degradation of high-molecular-weight PAHs, benzo[a]pyrene mineralization (53% of added [14C]benzo[a]pyrene was recovered as 14CO2 in 100 days), and reduction in the mutagenicity of organic soil extracts, compared with the indigenous microbes and soil amended with only axenic inocula.  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