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.

Background

Hydrogen production by fermenting bacteria such as Escherichia coli offers a potential source of hydrogen biofuel. Because H2 production involves consumption of 2H+, hydrogenase expression is likely to involve pH response and regulation. Hydrogenase consumption of protons in E. coli has been implicated in acid resistance, the ability to survive exposure to acid levels (pH 2–2.5) that are three pH units lower than the pH limit of growth (pH 5–6). Enhanced survival in acid enables a larger infective inoculum to pass through the stomach and colonize the intestine. Most acid resistance mechanisms have been defined using aerobic cultures, but the use of anaerobic cultures will reveal novel acid resistance mechanisms.

Methods and Principal Findings

We analyzed the pH regulation of bacterial hydrogenases in live cultures of E. coli K-12 W3110. During anaerobic growth in the range of pH 5 to 6.5, E. coli expresses three hydrogenase isoenzymes that reversibly oxidize H2 to 2H+. Anoxic conditions were used to determine which of the hydrogenase complexes contribute to acid resistance, measured as the survival of cultures grown at pH 5.5 without aeration and exposed for 2 hours at pH 2 or at pH 2.5. Survival of all strains in extreme acid was significantly lower in low oxygen than for aerated cultures. Deletion of hyc (Hyd-3) decreased anoxic acid survival 3-fold at pH 2.5, and 20-fold at pH 2, but had no effect on acid survival with aeration. Deletion of hyb (Hyd-2) did not significantly affect acid survival. The pH-dependence of H2 production and consumption was tested using a H2-specific Clark-type electrode. Hyd-3-dependent H2 production was increased 70-fold from pH 6.5 to 5.5, whereas Hyd-2-dependent H2 consumption was maximal at alkaline pH. H2 production, was unaffected by a shift in external or internal pH. H2 production was associated with hycE expression levels as a function of external pH.

Conclusions

Anaerobic growing cultures of E. coli generate H2 via Hyd-3 at low external pH, and consume H2 via Hyd-2 at high external pH. Hyd-3 proton conversion to H2 is required for acid resistance in anaerobic cultures of E. coli.  相似文献   

2.
The redox potential of the gut of nine species of termites was investigated by feeding the insects with redox dyes. The fore- and mid-gut of all species was aerobic with an E′o probably in excess of + 100 mV. whereas the paunch and colon were anaerobic with an E′o of about ?230 to ?270 mV, except in Coptotermes lacteus and Nasutitermes exitiosus whose colons were at a E′o of ?50 to ?125 mV. In four species (Incisitermes barretti, Glyptotermes brevicornis, Stolotermes victoriensis, Coptotermes lacteus) the rectum was aerobic (E′o about +60 mV), whereas the rectum of the other species was anaerobic (E′o from about ?125 to ?270 mV).  相似文献   

3.
《Insect Biochemistry》1990,20(5):443-449
The fine structure of the mid-gut of Poekilocerus bufonius has been examined and three types of epithelial cells were identified; normal epithelial cells with their apical part possessing well developed microvilli, goblet-like cells containing myelin-like figures and the small basal cells with small and round nuclei, nidi. The regulation of 6-phosphofructo-1-kinase (PFK-1) prepared from the mid-gut of the grasshopper, Poekilocerus bufonius, was studied. Mid-gut PFK-1 displayed cooperativity with respect to fructose-6-phosphate at pH 7.0, and the enzyme was inhibited by high concentrations of ATP. The affinity of the enzyme for fructose-6-phosphate was increased by fru-2,6-P2 whereas the inhibition of the enzyme by high concentrations of ATP was relieved by fru-2,6-P2. The activity of mid-gut PFK-1 was highly stimulated in a simultaneous presence of low concentrations of fru-2,6-P2 and AMP. ADP, AMP and c-AMP were all shown to be activators of the mid-gut PFK-1 with AMP being the greatest effector. The enzyme was not inhibited by citrate either in the presence of low or high concentrations of ATP. These results suggest that the PFK-1 of the mid-gut of the grasshopper is highly regulated with positive stimulators, specially fru-2,6-P2, whereas the enzyme is not regulated by citrate or glucose-1,6-bisphosphate.  相似文献   

4.
Monoaromatic hydrocarbons such as benzene, toluene, ethylbenzene, and xylene (BTEX) are widespread contaminants in groundwater. We examined the anaerobic degradation of BTEX compounds with amorphous ferric oxide as electron acceptor. Successful enrichment cultures were obtained for all BTEX substrates both in the presence and absence of AQDS (9,10-anthraquinone-2,6-disulfonic acid). The electron balances showed a complete anaerobic oxidation of the aromatic compounds to CO2. This is the first report on the anaerobic degradation of o-xylene and ethylbenzene in sediment-free iron-reducing enrichment cultures.  相似文献   

5.
Anaerobic degradation of naphthalene, 2-methylnaphthalene, and tetralin (1,2,3,4-tetrahydronaphthalene) was investigated with a sulfate-reducing enrichment culture obtained from a contaminated aquifer. Degradation studies with tetralin revealed 5,6,7,8-tetrahydro-2-naphthoic acid as a major metabolite indicating activation by addition of a C1 unit to tetralin, comparable to the formation of 2-naphthoic acid in anaerobic naphthalene degradation. The activation reaction was specific for the aromatic ring of tetralin; 1,2,3,4-tetrahydro-2-naphthoic acid was not detected. The reduced 2-naphthoic acid derivatives tetrahydro-, octahydro-, and decahydro-2-naphthoic acid were identified consistently in supernatants of cultures grown with either naphthalene, 2-methylnaphthalene, or tetralin. In addition, two common ring cleavage products were identified. Gas chromatography-mass spectrometry (GC-MS) and high-resolution GC-MS analyses revealed a compound with a cyclohexane ring and two carboxylic acid side chains as one of the first ring cleavage products. The elemental composition was C11H16O4 (C11H16O4-diacid), indicating that all carbon atoms of the precursor 2-naphthoic acid structure were preserved in this ring cleavage product. According to the mass spectrum, the side chains could be either an acetic acid and a propenic acid, or a carboxy group and a butenic acid side chain. A further ring cleavage product was identified as 2-carboxycyclohexylacetic acid and was assumed to be formed by β-oxidation of one of the side chains of the C11H16O4-diacid. Stable isotope-labeling growth experiments with either 13C-labeled naphthalene, per-deuterated naphthalene-d8, or a 13C-bicarbonate-buffered medium showed that the ring cleavage products derived from the introduced carbon source naphthalene. The series of identified metabolites suggests that anaerobic degradation of naphthalenes proceeds via reduction of the aromatic ring system of 2-naphthoic acid to initiate ring cleavage in analogy to the benzoyl-coenzyme A pathway for monoaromatic hydrocarbons. Our findings provide strong indications that further degradation goes through saturated compounds with a cyclohexane ring structure and not through monoaromatic compounds. A metabolic pathway for anaerobic degradation of bicyclic aromatic hydrocarbons with 2-naphthoic acid as the central intermediate is proposed.  相似文献   

6.
For anaerobic glucose-limited chemostat cultures of Aerobacter aerogenes a values of 14.0 g/mole was found for Y ATP max and a value of 6.8 mmoles ATP/g dry weight/hr for the maintenance coefficient. Both values are much lower than those previously determined for tryptophan-limited anaerobic chemostat cultures. It is concluded that generally the largest part of the maintenance energy is not used for true maintenance processes. For aerobic glucose-limited chemostat cultures two phases could be differentiated. Acetate production started at μ values higher than 0.53. The slopes of the curves relating the specific rates of glucose- and oxygen consumption with μ became higher and lower respectively above the μ value of 0.53. Using the Y ATP values obtained in the anaerobic experiment a P/O ratio of about 1.3 could be calculated for glucose- and tryptophan-limited chemostat cultures. In sulfate-limited chemostat cultures acetate was produced at all growth rates. At high growth rates also pyruvate and α-ketoglutarate were produced. With the Y ATP values obtained in the anaerobic experiment a P/O ratio of about 0.4 was calculated for sulfate-limited chemostat cultures.  相似文献   

7.
The purpose of this study was to characterize precipitates formed in anaerobic, H2S-producing cultures of two Tissierella isolates and Desulfosporosinus strain DB. The cultures were grown in Cu-containing media as part of a larger study of Cu resistance in anaerobic sulfidogens. The Tissierella strains produced H2S from peptone. Desulfosporosinus formed H2S from peptone or through dissimilatory sulfate reduction with lactate. Tissierella cultures precipitated iron phosphate, vivianite, but no crystalline phases or Cu sulfides were detected. Multiple Cu sulfides, including chalcopyrite and covellite, were detected in Desulfosporosinus cultures but vivianite was not formed. Ion microprobe spectra and electron microscopic examination showed major variation in the elemental composition and morphological differences depending on incubation conditions. Extended incubation time for at least 1–2 months increased the crystallinity of the precipitates. The results highlight biogeochemical differences in sulfide and phosphate precipitates between the two major groups of Firmicutes although they may share the same habitat including the human intestinal tract.  相似文献   

8.
Recent studies have demonstrated that fumarate addition and carboxylation are two possible mechanisms of anaerobic alkane degradation. In the present study, we surveyed metabolites formed during growth on hexadecane by the sulfate-reducing isolates AK-01 and Hxd3 and by a mixed sulfate-reducing consortium. The cultures were incubated with either protonated or fully deuterated hexadecane; the sulfate-reducing consortium was also incubated with [1,2-13C2]hexadecane. All cultures were extracted, silylated, and analyzed by gas chromatography-mass spectrometry. We detected a suite of metabolites that support a fumarate addition mechanism for hexadecane degradation by AK-01, including methylpentadecylsuccinic acid, 4-methyloctadecanoic acid, 4-methyloctadec-2,3-enoic acid, 2-methylhexadecanoic acid, and tetradecanoic acid. By using d34-hexadecane, mass spectral evidence strongly supporting a carbon skeleton rearrangement of the first intermediate, methylpentadecylsuccinic acid, was demonstrated for AK-01. Evidence indicating hexadecane carboxylation was not found in AK-01 extracts but was observed in Hxd3 extracts. In the mixed sulfate-reducing culture, however, metabolites consistent with both fumarate addition and carboxylation mechanisms of hexadecane degradation were detected, which demonstrates that multiple alkane degradation pathways can occur simultaneously within distinct anaerobic communities. Collectively, these findings underscore that fumarate addition and carboxylation are important alkane degradation mechanisms that may be widespread among phylogenetically and/or physiologically distinct microorganisms.  相似文献   

9.
Nutritional Requirements of Methanosarcina sp. Strain TM-1   总被引:2,自引:1,他引:1       下载免费PDF全文
Methanosarcina sp. strain TM-1, an acetotrophic, thermophilic methanogen isolated from an anaerobic sludge digestor, was originally reported to require an anaerobic sludge supernatant for growth. It was found that the sludge supernatant could be replaced with yeast extract (1 g/liter), 6 mM bicarbonate-30% CO2, and trace metals, with a doubling time on methanol of 14 h. For growth on either methanol or acetate, yeast extract could be replaced with CaCl2 · 2H2O (13.6 μM minimum) and the vitamin p-aminobenzoic acid (PABA, ca. 3 nM minimum), with a doubling time on methanol of 8 to 9 h. Filter-sterilized folic acid at 0.3 μM could not replace PABA. The antimetabolite sulfanilamide (20 mM) inhibited growth of and methanogenesis by Methanosarcina sp. strain TM-1, and this inhibition was reversed by the addition of 0.3 μM PABA. When a defined medium buffered with 20 mM N,N-bis(2-hydroxyethyl)-2-aminoethanesulfonic acid was used, it was shown that Methanosarcina sp. strain TM-1 required 6 mM bicarbonate-30% CO2 for optimal growth and methanogenesis from methanol. Cells growing on acetate were less dependent on bicarbonate-CO2. When we used a defined medium in which the only organic compounds present were methanol or acetate, nitrilotriacetic acid (0.2 mM), and PABA, it was possible to limit batch cultures of Methanosarcina sp. strain TM-1 for nitrogen at NH4+ concentrations at or below 2.0 mM, in marked contrast with Methanosarcina barkeri 227, which fixes dinitrogen when grown under NH4+ limitation.  相似文献   

10.
Mixed cultures of bacteria, enriched from aquatic sediments, grew anaerobically on all three isomers of phthalic acid. Each culture grew anaerobically on only one isomer and also grew aerobically on the same isomer. Pure cultures were isolated from the phthalic acid (o-phthalic acid) and isophthalic acid (m-phthalic acid) enrichments that grew aerobically on phthalic and isophthalic acids. Cell suspension experiments indicated that protocatechuate is an intermediate of aerobic catabolism. Pure cultures which grew aerobically on terephthalic acid (p-phthalic acid) could not be isolated from the enrichments, and neither could pure cultures that grew anaerobically on any of the isomers. Cell suspension experiments suggested that separate pathways exist for the aerobic and anaerobic oxidation of phthalic acids. Each enrichment culture used only one phthalic acid isomer under anaerobic conditions, but all isomers were simultaneously adapted for the anaerobic catabolism of benzoate. Cells grown anaerobically on a phthalic acid immediately attacked the isomer under anaerobic conditions, whereas there was a lag before aerobic breakdown occurred, and, for phthalic and terephthalic acids, chloramphenicol stopped aerobic adaptation but had no effect on anaerobic catabolism. This work suggests that phthalic acids are biodegradable in anaerobic environments.  相似文献   

11.
Kluyveromyces fragilis (CBS 397) is a nonhalophilic yeast which is capable of lactose utilization from whey permeate and high glycerol production under anaerobic growth conditions. However, the optimum yields of glycerol (11.6 mg/ml of whey permeate medium) obtained in this study occurred only in the presence of 1% Na2SO3 as a steering agent. The use of other concentrations of Na2SO3, as well as 5% NaCl and 1% ascorbic acid, had no or detrimental effects on cell growth, lactose utilization, and glycerol production. Glycerol yields were greater in cultures grown from a light inoculum of K. fragilis than in cultures in which a resuspended mass of cells was introduced into the medium. The results of this study suggest that this strain of K. fragilis may be useful commercially in the utilization of cheese whey lactose and the concomitant production of glycerol.  相似文献   

12.
MnO2 reduction by aerobic growing cultures of Bacillus 29 and coccus 32, isolated from ferromanganese nodules, was assessed for 7 days. A 1-day lag was observed before the onset of MnO2 reduction by either culture. Addition of HgCl2 to a final concentration of about 10-3 M caused a rapid cessation of MnO2 reduction by the growing cultures. Neither culture reduced MnO2 when grown under continued anaerobiosis from the start of an experiment. However, if conditions were made anaerobic after MnO2 reduction was initiated, reduction continued at a rate only slightly lower than that under aerobic conditions. Resting-cell cultures reduced MnO2 equally well aerobically and anaerobically, provided that ferricyanide was present to serve as electron carrier. These findings showed that oxygen is needed for culture adaptation to MnO2 reduction, and that oxygen does not interfere with microbial MnO2 reduction itself. Both cultures caused sharp drops in the pH of the medium during MnO2 reduction: with coccus 32, during the entire incubation time; with Bacillus 29, for the first 3 days. The Eh of the medium fluctuated with either culture and never fell below 469 mv with Bacillus 29 and below 394 mv with coccus 32. The rates of glucose consumption and Mn2+ release by Bacillus 29 and coccus 32 were fairly constant, but the rates of lactate and pyruvate production were not. Although acid production undoubtedly helped in the reduction of pyrolusite (MnO2) by the bacteria, it did not appear to be important in the reduction of manganese oxide in ferromanganese nodules, as shown by the results with a nodule enrichment.  相似文献   

13.
MnO2 reduction by aerobic growing cultures of Bacillus 29 and coccus 32, isolated from ferromanganese nodules, was assessed for 7 days. A 1-day lag was observed before the onset of MnO2 reduction by either culture. Addition of HgCl2 to a final concentration of about 10-3 M caused a rapid cessation of MnO2 reduction by the growing cultures. Neither culture reduced MnO2 when grown under continued anaerobiosis from the start of an experiment. However, if conditions were made anaerobic after MnO2 reduction was initiated, reduction continued at a rate only slightly lower than that under aerobic conditions. Resting-cell cultures reduced MnO2 equally well aerobically and anaerobically, provided that ferricyanide was present to serve as electron carrier. These findings showed that oxygen is needed for culture adaptation to MnO2 reduction, and that oxygen does not interfere with microbial MnO2 reduction itself. Both cultures caused sharp drops in the pH of the medium during MnO2 reduction: with coccus 32, during the entire incubation time; with Bacillus 29, for the first 3 days. The Eh of the medium fluctuated with either culture and never fell below 469 mv with Bacillus 29 and below 394 mv with coccus 32. The rates of glucose consumption and Mn2+ release by Bacillus 29 and coccus 32 were fairly constant, but the rates of lactate and pyruvate production were not. Although acid production undoubtedly helped in the reduction of pyrolusite (MnO2) by the bacteria, it did not appear to be important in the reduction of manganese oxide in ferromanganese nodules, as shown by the results with a nodule enrichment.  相似文献   

14.
15.
Vanillic acid (4-hydroxy-3-methoxybenzoic acid) supported the anaerobic (nitrate respiration) but not the aerobic growth of Pseudomonas sp. strain PN-1. Cells grown anaerobically on vanillate oxidized vanillate, p-hydroxybenzoate, and protocatechuic acid (3,4-dihydroxybenzoic acid) with O2 or nitrate. Veratric acid (3,4-dimethoxybenzoic acid) but not isovanillic acid (3-hydroxy-4-methoxybenzoic acid) induced cells for the oxic and anoxic utilization of vanillate, and protocatechuate was detected as an intermediate of vanillate breakdown under either condition. Aerobic catabolism of protocatechuate proceeded via 4,5-meta cleavage, whereas anaerobically it was probably dehydroxylated to benzoic acid. Formaldehyde was identified as a product of aerobic demethylation, indicating a monooxygenase mechanism, but was not detected during anaerobic demethylation. The aerobic and anaerobic systems had similar but not identical substrate specificities. Both utilized m-anisic acid (3-methoxybenzoic acid) and veratrate but not o- or p-anisate and isovanillate. Syringic acid (4-hydroxy-3,5-dimethoxybenzoic acid), 3-O-methylgallic acid (3-methoxy-4,5-dihydroxybenzoic acid), and 3,5-dimethoxybenzoic acid were attacked under either condition, and formaldehyde was liberated from these substrates in the presence of O2. The anaerobic demethylating system but not the aerobic enzyme was also active upon guaiacol (2-methoxyphenol), ferulic acid (3-[4-hydroxy-3-methoxyphenyl]-2-propenoic acid), 3,4,5-trimethoxycinnamic acid (3-[3,4,5-trimethoxyphenyl]-2-propenoic acid), and 3,4,5-trimethoxybenzoic acid. The broad specificity of the anaerobic demethylation system suggests that it probably is significant in the degradation of lignoaromatic molecules in anaerobic environments.  相似文献   

16.
Biogenic emissions of nitric and nitrous oxides have important impacts on the photochemistry and chemistry of the atmosphere. Although biogenic production appears to be the overwhelming source of N2O, the magnitude of the biogenic emission of NO is very uncertain. In soils, possible sources of NO and N2O include nitrification by autotrophic and heterotrophic nitrifiers, denitrification by nitrifiers and denitrifiers, nitrate respiration by fermenters, and chemodenitrification. The availability of oxygen determines to a large extent the relative activities of these various groups of organisms. To better understand this influence, we investigated the effect of the partial pressure of oxygen (pO2) on the production of NO and N2O by a wide variety of common soil nitrifying, denitrifying, and nitrate-respiring bacteria under laboratory conditions. The production of NO per cell was highest by autotrophic nitrifiers and was independent of pO2 in the range tested (0.5 to 10%), whereas N2O production was inversely proportional to pO2. Nitrous oxide production was highest in the denitrifier Pseudomonas fluorescens, but only under anaerobic conditions. The molar ratio of NO/N2O produced was usually greater than unity for nitrifiers and much less than unity for denitrifiers. Chemodenitrification was the major source of both the NO and N2O produced by the nitrate respirer Serratia marcescens. Chemodenitrification was also a possible source of NO and N2O in nitrifier cultures but only when high concentrations of nitrite had accumulated or were added to the medium. Although most of the denitrifiers produced NO and N2O only under anaerobic conditions, chemostat cultures of Alcaligenes faecalis continued to emit these gases even when the cultures were sparged with air. Based upon these results, we predict that aerobic soils are primary sources of NO and that N2O is produced only when there is sufficient soil moisture to provide the anaerobic microsites necessary for denitrification by either denitrifiers or nitrifiers.  相似文献   

17.
18.
Rhodospirillum rubrum (F1) maintained electron balance mainly by producing propionate, formate and H2 during fermentation metabolism. H2 formation was inversely correlated with the production of propionate.In diluted, growing cultures high amounts of H2 and only traces or no propionate were produced from pyruvate. In dense cultures or in resting cultures without (NH4)2SO4, however, propionate was formed from pyruvate in relatively high amounts Cultures always produced much more propionate than H2 from fructose in contrast to cells with pyruvate. Kinetic studies of growth and excretion of fermentation products indicated that the enzyme system for H2 formation is adaptive. Chloramphenicol (3 μg/ml) completely inhibited the formation of H2 if the cells were not adapted to fermentation metabolism. The production of propionate, on the other hand, was not prevented by chloramphenicol after shifting the cells from aerobic dark culture with malate to fermentation conditions with pyruvate.H2 formation was not influenced by sodium ascorbate but it was significantly decreased by K3[Fe(CN)6].Poly(β-hydroxybutyric acid) was also synthesized by the cells during anaerobic dark metabolism especially in dense cultures, probably favoured by the rapid acidification of the medium. Formate can also accumulate in the fermentation metabolism, especially in young growing cultures.These results give an explanation for the differing reports in the literature on the fermentation metabolism of R. rubrum.  相似文献   

19.
The purple nonsulfur bacterium Rhodospirillum rubrum has been employed to study physiological adaptation to limiting oxygen tensions (microaerophilic conditions). R. rubrum produces maximal levels of photosynthetic membranes when grown with both succinate and fructose as carbon sources under microaerophilic conditions in comparison to the level (only about 20% of the maximum) seen in the absence of fructose. Employing a unique partial O2 pressure (pO2) control strategy to reliably adjust the oxygen tension to values below 0.5%, we have used bioreactor cultures to investigate the metabolic rationale for this effect. A metabolic profile of the central carbon metabolism of these cultures was obtained by determination of key enzyme activities under microaerophilic as well as aerobic and anaerobic phototrophic conditions. Under aerobic conditions succinate and fructose were consumed simultaneously, whereas oxygen-limiting conditions provoked the preferential breakdown of fructose. Fructose was utilized via the Embden-Meyerhof-Parnas pathway. High levels of pyrophosphate-dependent phosphofructokinase activity were found to be specific for oxygen-limited cultures. No glucose-6-phosphate dehydrogenase activity was detected under any conditions. We demonstrate that NADPH is supplied mainly by the pyridine-nucleotide transhydrogenase under oxygen-limiting conditions. The tricarboxylic acid cycle enzymes are present at significant levels during microaerophilic growth, albeit at lower levels than those seen under fully aerobic growth conditions. Levels of the reductive tricarboxylic acid cycle marker enzyme fumarate reductase were also high under microaerophilic conditions. We propose a model by which the primary “switching” of oxidative and reductive metabolism is performed at the level of the tricarboxylic acid cycle and suggest how this might affect redox signaling and gene expression in R. rubrum.  相似文献   

20.
Most heterotrophic bacteria assimilate CO2 in various carboxylation reactions during biosynthesis. In this study, assimilation of 14CO2 by heterotrophic bacteria was used for isotope labeling of active microorganisms in pure cultures and environmental samples. Labeled cells were visualized by microautoradiography (MAR) combined with fluorescence in situ hybridization (FISH) to obtain simultaneous information about activity and identity. Cultures of Escherichia coli and Pseudomonas putida assimilated sufficient 14CO2 during growth on various organic substrates to obtain positive MAR signals. The MAR signals were comparable with the traditional MAR approach based on uptake of 14C-labeled organic substrates. Experiments with E. coli showed that 14CO2 was assimilated during both fermentation and aerobic and anaerobic respiration. The new MAR approach, HetCO2-MAR, was evaluated by targeting metabolic active filamentous bacteria, including “Candidatus Microthrix parvicella” in activated sludge. “Ca. Microthrix parvicella” was able to take up oleic acid under anaerobic conditions, as shown by the traditional MAR approach with [14C]oleic acid. However, the new HetCO2-MAR approach indicated that “Ca. Microthrix parvicella,” did not significantly grow on oleic acid under anaerobic conditions with or without addition of NO2, whereas the addition of O2 or NO3 initiated growth, as indicated by detectable 14CO2 assimilation. This is a metabolic feature that has not been described previously for filamentous bacteria. Such information could not have been derived by using the traditional MAR procedure, whereas the new HetCO2-MAR approach differentiates better between substrate uptake and substrate metabolism that result in growth. The HetCO2-MAR results were supported by stable isotope analysis of 13C-labeled phospholipid fatty acids from activated sludge incubated under aerobic and anaerobic conditions in the presence of 13CO2. In conclusion, the novel HetCO2-MAR approach expands the possibility for studies of the ecophysiology of uncultivated microorganisms.  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