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 46 毫秒
1.
A 3-hydroxypropionate/4-hydroxybutyrate cycle operates during autotrophic CO2 fixation in various members of the Crenarchaea. In this cycle, as determined using Metallosphaera sedula, malonyl-coenzyme A (malonyl-CoA) and succinyl-CoA are reductively converted via their semialdehydes to the corresponding alcohols 3-hydroxypropionate and 4-hydroxybutyrate. Here three missing oxidoreductases of this cycle were purified from M. sedula and studied. Malonic semialdehyde reductase, a member of the 3-hydroxyacyl-CoA dehydrogenase family, reduces malonic semialdehyde with NADPH to 3-hydroxypropionate. The latter compound is converted via propionyl-CoA to succinyl-CoA. Succinyl-CoA reduction to succinic semialdehyde is catalyzed by malonyl-CoA/succinyl-CoA reductase, a promiscuous NADPH-dependent enzyme that is a paralogue of aspartate semialdehyde dehydrogenase. Succinic semialdehyde is then reduced with NADPH to 4-hydroxybutyrate by succinic semialdehyde reductase, an enzyme belonging to the Zn-dependent alcohol dehydrogenase family. Genes highly similar to the Metallosphaera genes were found in other members of the Sulfolobales. Only distantly related genes were found in the genomes of autotrophic marine Crenarchaeota that may use a similar cycle in autotrophic carbon fixation.The thermoacidophilic autotrophic crenarchaeum Metallosphaera sedula uses a 3-hydroxypropionate/4-hydroxybutyrate cycle for CO2 fixation (9, 28, 29, 35) (Fig. (Fig.1).1). A similar cycle may operate in other autotrophic members of the Sulfolobales (31) and in mesophilic marine group I Crenarchaea (Cenarchaeum sp., Nitrosopumilus sp.). This cycle uses elements of the 3-hydroxypropionate cycle that was originally discovered in the phototrophic bacterium Chloroflexus aurantiacus (15, 22-25, 41, 42). It involves the carboxylation of acetyl coenzyme A (acetyl-CoA) to malonyl-CoA by a biotin-dependent acetyl-CoA carboxylase (12, 29). The carboxylation product is reduced to malonic semialdehyde by malonyl-CoA reductase (1). Malonic semialdehyde is further reduced to 3-hydroxypropionate, the characteristic intermediate of the pathway (9, 31, 35). 3-Hydroxypropionate is further reductively converted to propionyl-CoA (3), which is carboxylated to (S)-methylmalonyl-CoA by propionyl-CoA carboxylase. Only one copy of the genes encoding the acetyl-CoA/propionyl-CoA carboxylase subunits is present in most Archaea, indicating that this enzyme is a promiscuous enzyme that acts on both acetyl-CoA and propionyl-CoA (12, 29). (S)-Methylmalonyl-CoA is isomerized to (R)-methylmalonyl-CoA, which is followed by carbon rearrangement to succinyl-CoA catalyzed by coenzyme B12-dependent methylmalonyl-CoA mutase.Open in a separate windowFIG. 1.Proposed 3-hydroxypropionate/4-hydroxybutyrate cycle in M. sedula and other autotrophic Sulfolobales. Enzymes: 1, acetyl-CoA carboxylase; 2, malonyl-CoA reductase (NADPH); 3, malonate semialdehyde reductase (NADPH); 4, 3-hydroxypropionate-CoA ligase (AMP forming); 5, 3-hydroxypropionyl-CoA dehydratase; 6, acryloyl-CoA reductase (NADPH); 7, propionyl-CoA carboxylase, identical to acetyl-CoA carboxylase; 8, (S)-methylmalonyl-CoA epimerase; 9, methylmalonyl-CoA mutase; 10, succinyl-CoA reductase (NADPH), identical to malonyl-CoA reductase; 11, succinic semialdehyde reductase (NADPH); 12, 4-hydroxybutyrate-CoA ligase (AMP forming); 13, 4-hydroxybutyryl-CoA dehydratase; 14, crotonyl-CoA hydratase; 15, (S)-3-hydroxybutyryl-CoA dehydrogenase (NAD+); 16, acetoacetyl-CoA β-ketothiolase. The highlighted steps are catalyzed by the enzymes studied here.Succinyl-CoA is converted via succinic semialdehyde and 4-hydroxybutyrate to two molecules of acetyl-CoA (9), thus regenerating the starting CO2 acceptor molecule and releasing another acetyl-CoA molecule for biosynthesis. Hence, the 3-hydroxypropionate/4-hydroxybutyrate cycle (Fig. (Fig.1)1) can be divided into two parts. The first part transforms one acetyl-CoA molecule and two bicarbonate molecules into succinyl-CoA (Fig. (Fig.1,1, steps 1 to 9), and the second part converts succinyl-CoA to two acetyl-CoA molecules (Fig. (Fig.1,1, steps 10 to 16).The second part of the autotrophic cycle also occurs in the dicarboxylate/4-hydroxybutyrate cycle, which operates in autotrophic CO2 fixation in Desulfurococcales and Thermoproteales (Crenarchaea) (27, 37), raising the question of whether the enzymes in these two lineages have common roots (37). The first part of the cycle also occurs in the 3-hydroxypropionate cycle for autotrophic CO2 fixation in Chloroflexus aurantiacus and a few related green nonsulfur phototrophic bacteria (19, 22, 23, 32, 49).The two-step reduction of malonyl-CoA to 3-hydroxpropionate in Chloroflexus is catalyzed by a single bifunctional 300-kDa enzyme (30). The M. sedula malonyl-CoA reductase is completely unrelated and forms only malonic semialdehyde (1), and the enzyme catalyzing the second malonic semialdehyde reduction step that forms 3-hydroxypropionate is unknown. In the second part of the 3-hydroxypropionate/4-hydroxybutyrate cycle a similar reduction of succinyl-CoA via succinic semialdehyde to 4-hydroxybutyrate takes place. The enzymes responsible for these reactions also have not been characterized.In this work we purified the enzymes malonic semialdehyde reductase, succinyl-CoA reductase, and succinic semialdehyde reductase from M. sedula. The genes coding for these enzymes were identified in the genome, and recombinant proteins were studied in some detail. Interestingly, succinyl-CoA reductase turned out to be identical to malonyl-CoA reductase. We also show here that enzymes that are highly similar to succinyl-CoA reductase in Thermoproteus neutrophilus do not function as succinyl-CoA reductases in M. sedula.  相似文献   

2.
Metallosphaera sedula (Sulfolobales, Crenarchaeota) uses the 3-hydroxypropionate/4-hydroxybutyrate cycle for autotrophic carbon fixation. In this pathway, acetyl-coenzyme A (CoA) and succinyl-CoA are the only intermediates that can be considered common to the central carbon metabolism. We addressed the question of which intermediate of the cycle most biosynthetic routes branch off. We labeled autotrophically growing cells by using 4-hydroxy[1-14C]butyrate and [1,4-13C1]succinate, respectively, as precursors for biosynthesis. The labeling patterns of protein-derived amino acids verified the operation of the proposed carbon fixation cycle, in which 4-hydroxybutyrate is converted to two molecules of acetyl-CoA. The results also showed that major biosynthetic flux does not occur via acetyl-CoA, except for the formation of building blocks that are directly derived from acetyl-CoA. Notably, acetyl-CoA is not assimilated via reductive carboxylation to pyruvate. Rather, our data suggest that the majority of anabolic precursors are derived from succinyl-CoA, which is removed from the cycle via oxidation to malate and oxaloacetate. These C4 intermediates yield pyruvate and phosphoenolpyruvate (PEP). Enzyme activities that are required for forming intermediates from succinyl-CoA were detected, including enzymes catalyzing gluconeogenesis from PEP. This study completes the picture of the central carbon metabolism in autotrophic Sulfolobales by connecting the autotrophic carbon fixation cycle to the formation of central carbon precursor metabolites.Sulfolobales (Crenarchaeota) comprise extreme thermoacidophiles from volcanic areas that grow best at a pH of around 2 and a temperature of 60 to 90°C (32, 33). Most Sulfolobales can grow chemoautotrophically on sulfur, pyrite, or H2 under microaerobic conditions, which also applies to Metallosphaera sedula (31), the organism studied here. Its genome has been sequenced (2). Some species of the Sulfolobales secondarily returned to a facultative anaerobic or even strictly anaerobic life style (33), and some laboratory strains appear to have lost their ability to grow autotrophically (8). Autotrophic representatives of the Sulfolobales use a 3-hydroxypropionate/4-hydroxybutyrate cycle (in short, hydroxypropionate/hydroxybutyrate cycle) for autotrophic carbon fixation (Fig. (Fig.1)1) (6-8, 38). The enzymes of this cycle are oxygen tolerant, which predestines the cycle for the lifestyle of the aerobic Crenarchaeota (8). The presence of genes coding for key enzymes of the hydroxypropionate/hydroxybutyrate cycle in the mesophilic aerobic “marine group I” Crenarchaeota suggests that these abundant marine archaea use a similar autotrophic carbon fixation mechanism (6, 24, 68) (for a review of autotrophic carbon fixation in Archaea, see reference 7).Open in a separate windowFIG. 1.Proposed 3-hydroxypropionate/4-hydroxybutyrate cycle functioning in autotrophic carbon fixation in Sulfolobales and its relation to the central carbon metabolism, as studied in this work for Metallosphaera sedula. The situation may be similar in other Sulfolobales and possibly in autotrophic marine Crenarchaeota. Enzymes: 1, acetyl-CoA/propionyl-CoA carboxylase; 2, malonyl-CoA reductase (NADPH); 3, malonic semialdehyde reductase (NADPH); 4, 3-hydroxypropionate-CoA ligase (AMP forming); 5, 3-hydroxypropionyl-CoA dehydratase; 6, acryloyl-CoA reductase (NADPH); 7, acetyl-CoA/propionyl-CoA carboxylase; 8, methylmalonyl-CoA epimerase; 9, methylmalonyl-CoA mutase; 10, succinyl-CoA reductase (NADPH); 11, succinic semialdehyde reductase (NADPH); 12, 4-hydroxybutyrate-CoA ligase (AMP forming); 13, 4-hydroxybutyryl-CoA dehydratase; 14 and 15, crotonyl-CoA hydratase/(S)-3-hydroxybutyryl-CoA dehydrogenase (NAD+); 16, acetoacetyl-CoA β-ketothiolase; 17, succinyl-CoA synthetase (ADP forming); 18, succinic semialdehyde dehydrogenase; 19, succinate dehydrogenase (natural electron acceptor unknown); 20, fumarate hydratase; 21, malate dehydrogenase; 22, malic enzyme; 23, PEP carboxykinase (GTP); 24, pyruvate:water dikinase (ATP); 25, enolase; 26, phosphoglycerate mutase; 27, phosphoglycerate kinase; 28, glyceraldehyde 3-phosphate dehydrogenase; 29, triosephosphate isomerase; 30, fructose 1,6-bisphosphate aldolase/phosphatase; 31, (si)-citrate synthase; 32, aconitase; 33, isocitrate dehydrogenase.In the cycle, one molecule of acetyl-coenzyme A (CoA) is formed from two molecules of bicarbonate. The key carboxylating enzyme is a bifunctional biotin-dependent acetyl-CoA/propionyl-CoA carboxylase (10, 11, 36, 38, 48, 49). In Bacteria and Eukarya, acetyl-CoA carboxylase catalyzes the first step in fatty acid biosynthesis. However, archaea do not contain fatty acids, and therefore acetyl-CoA carboxylase obviously plays a different metabolic role. The hydroxypropionate/hydroxybutyrate cycle can be divided into two parts. The first transforms acetyl-CoA and two bicarbonate molecules via 3-hydroxypropionate to succinyl-CoA, and the second converts succinyl-CoA via 4-hydroxybutyrate to two acetyl-CoA molecules. In brief, the product of the acetyl-CoA carboxylase reaction, malonyl-CoA, is reduced via malonic semialdehyde to 3-hydroxypropionate, which is further reductively converted to propionyl-CoA. Propionyl-CoA is carboxylated to (S)-methylmalonyl-CoA by the same carboxylase as that that carboxylates acetyl-CoA (11, 36). (S)-Methylmalonyl-CoA is isomerized to (R)-methylmalonyl-CoA, followed by carbon rearrangement to succinyl-CoA catalyzed by coenzyme B12-dependent methylmalonyl-CoA mutase.Succinyl-CoA then is converted into two molecules of acetyl-CoA via succinic semialdehyde, 4-hydroxybutyrate, 4-hydroxybutyryl-CoA, crotonyl-CoA, 3-hydroxyacetyl-CoA, and acetoacetyl-CoA. This reaction sequence apparently is common to the autotrophic Crenarchaeota, as it also is used by autotrophic Crenarchaeota of the orders Thermoproteales and Desulfurococcales, which use a dicarboxylate/4-hydroxybutyrate cycle for autotrophic carbon fixation (8, 34, 55, 56) (also see the accompanying work [57]).From the list of intermediates of the hydroxypropionate/hydroxybutyrate cycle, acetyl-CoA and succinyl-CoA are the only intermediates considered common to the central carbon metabolism. In this work, we addressed the question of which intermediate of the cycle most biosynthetic routes branch off, and we came to the conclusion that succinyl-CoA serves as the main precursor for cellular carbon. This requires one turn of the cycle to regenerate the CO2 acceptor and to generate one extra molecule of acetyl-CoA from two molecules of bicarbonate. Acetyl-CoA plus another two bicarbonate molecules are converted by an additional half turn of the cycle to succinyl-CoA. This strategy differs from that of the anaerobic pathways, in which acetyl-CoA is reductively carboxylated to pyruvate, and from there the other precursors for building blocks ultimately are derived (discussed in reference 7).  相似文献   

3.
4.
A 3-hydroxypropionate/4-hydroxybutyrate cycle operates in autotrophic CO2 fixation in various Crenarchaea, as studied in some detail in Metallosphaera sedula. This cycle and the autotrophic 3-hydroxypropionate cycle in Chloroflexus aurantiacus have in common the conversion of acetyl-coenzyme A (CoA) and two bicarbonates via 3-hydroxypropionate to succinyl-CoA. Both cycles require the reductive conversion of 3-hydroxypropionate to propionyl-CoA. In M. sedula the reaction sequence is catalyzed by three enzymes. The first enzyme, 3-hydroxypropionyl-CoA synthetase, catalyzes the CoA- and MgATP-dependent formation of 3-hydroxypropionyl-CoA. The next two enzymes were purified from M. sedula or Sulfolobus tokodaii and studied. 3-Hydroxypropionyl-CoA dehydratase, a member of the enoyl-CoA hydratase family, eliminates water from 3-hydroxypropionyl-CoA to form acryloyl-CoA. Acryloyl-CoA reductase, a member of the zinc-containing alcohol dehydrogenase family, reduces acryloyl-CoA with NADPH to propionyl-CoA. Genes highly similar to the Metallosphaera CoA synthetase, dehydratase, and reductase genes were found in autotrophic members of the Sulfolobales. The encoded enzymes are only distantly related to the respective three enzyme domains of propionyl-CoA synthase from C. aurantiacus, where this trifunctional enzyme catalyzes all three reactions. This indicates that the autotrophic carbon fixation cycles in Chloroflexus and in the Sulfolobales evolved independently and that different genes/enzymes have been recruited in the two lineages that catalyze the same kinds of reactions.In the thermoacidophilic autotrophic crenarchaeum Metallosphaera sedula, CO2 fixation proceeds via a 3-hydroxypropionate/4-hydroxybutyrate cycle (8, 23, 24, 28) (Fig. (Fig.1).1). A similar cycle may operate in other autotrophic members of the Sulfolobales and in mesophilic Crenarchaea (Cenarchaeum sp. and Nitrosopumilus sp.) of marine group I. The cycle uses elements of the 3-hydroxypropionate cycle that was originally discovered in the phototrophic bacterium Chloroflexus aurantiacus (11, 16, 17, 19, 20, 32, 33). It involves the carboxylation of acetyl-coenzyme A (CoA) to malonyl-CoA by the biotin-dependent acetyl-CoA carboxylase. Malonyl-CoA is reduced via malonate semialdehyde to 3-hydroxypropionate (1), which is further reductively converted to propionyl-CoA (3). Propionyl-CoA is carboxylated to (S)-methylmalonyl-CoA by a propionyl-CoA carboxylase that is similar or identical to acetyl-CoA carboxylase. In fact, only one copy of the genes for the acetyl-CoA/propionyl-CoA carboxylase subunits is present in most Archaea, suggesting that this is a promiscuous enzyme that acts on both acetyl-CoA and propionyl-CoA (24). (S)-Methylmalonyl-CoA is epimerized to (R)-methylmalonyl-CoA, followed by carbon rearrangement to succinyl-CoA by coenzyme B12-dependent methylmalonyl-CoA mutase.Open in a separate windowFIG. 1.Proposed 3-hydroxypropionate/4-hydroxybutyrate cycle in M. sedula and other members of the Sulfolobales. Enzymes are the following: 1, acetyl-CoA carboxylase; 2, malonyl-CoA reductase (NADPH); 3, malonate semialdehyde reductase (NADPH); 4, 3-hydroxypropionyl-CoA synthetase (3-hydroxypropionate-CoA ligase, AMP forming); 5, 3-hydroxypropionyl-CoA dehydratase; 6, acryloyl-CoA reductase (NADPH); 7, propionyl-CoA carboxylase; 8, methylmalonyl-CoA epimerase; 9, methylmalonyl-CoA mutase; 10, succinyl-CoA reductase (NADPH); 11, succinate semialdehyde reductase (NADPH); 12, 4-hydroxybutyryl-CoA synthetase (4-hydroxybutyrate-CoA ligase, AMP-forming); 13, 4-hydroxybutyryl-CoA dehydratase; 14, crotonyl-CoA hydratase; 15, (S)-3-hydroxybutyryl-CoA dehydrogenase (NAD+); 16, acetoacetyl-CoA β-ketothiolase. The two steps of interest are highlighted.In Chloroflexus succinyl-CoA is converted to (S)-malyl-CoA, which is cleaved by (S)-malyl-CoA lyase to acetyl-CoA (thus regenerating the CO2 acceptor molecule) and glyoxylate (16). Glyoxylate is assimilated into cell material by a yet not completely resolved pathway (37). In Metallosphaera succinyl-CoA is converted via 4-hydroxybutyrate to two molecules of acetyl-CoA (8), thus regenerating the starting CO2 acceptor molecule and releasing another acetyl-CoA for biosynthesis. Hence, the 3-hydroxypropionate/4-hydroxybutyrate cycle (Fig. (Fig.1)1) can be divided into two parts. The first part transforms one acetyl-CoA and two bicarbonates into succinyl-CoA, and the second part converts succinyl-CoA to two acetyl-CoA molecules.The reductive conversion of 3-hydroxypropionate to propionyl-CoA requires three enzymatic steps: activation of 3-hydroxypropionate to its CoA ester, dehydration of 3-hydroxypropionyl-CoA to acryloyl-CoA, and reduction of acryloyl-CoA to propionyl-CoA. In C. aurantiacus these three steps are catalyzed by a single large trifunctional enzyme, propionyl-CoA synthase (2). This 200-kDa fusion protein consists of a CoA ligase, a dehydratase, and a reductase domain. Attempts to isolate a similar enzyme from M. sedula failed. Rather, a 3-hydroxypropionyl-CoA synthetase was found (3), suggesting that the other two reactions may also be catalyzed by individual enzymes.Here, we purified the missing enzymes 3-hydroxypropionyl-CoA dehydratase and acryloyl-CoA reductase from M. sedula, identified the coding genes in the genome of M. sedula and other members of the Sulfolobales, produced recombinant enzymes as proof of function, and studied the enzymes in some detail. A comparison with the respective domains of propionyl-CoA synthase from C. aurantiacus indicates that the conversion of 3-hydroxypropionate to propionyl-CoA via the 3-hydroxypropionate route has evolved independently in these two phyla.  相似文献   

5.
16S rRNA gene libraries from the lithoautotrophic Fe(II)-oxidizing, nitrate-reducing enrichment culture described by Straub et al. (K. L. Straub, M. Benz, B. Schink, and F. Widdel, Appl. Environ. Microbiol. 62:1458-1460, 1996) were dominated by a phylotype related (95% 16S rRNA gene homology) to the autotrophic Fe(II) oxidizer Sideroxydans lithotrophicus. The libraries also contained phylotypes related to known heterotrophic nitrate reducers Comamonas badia, Parvibaculum lavamentivorans, and Rhodanobacter thiooxidans. The three heterotrophs were isolated and found to be capable of only partial (12 to 24%) Fe(II) oxidation, suggesting that the Sideroxydans species has primary responsibility for Fe(II) oxidation in the enrichment culture.A variety of microorganisms oxidize Fe(II) with nitrate under anaerobic, circumneutral pH conditions (29) and may contribute to an active microbially driven anoxic Fe redox cycle (1, 27-29, 31, 32). Straub et al. (28) obtained the first Fe(II)-oxidizing, nitrate-reducing (enrichment) culture capable of fully autotrophic growth by a reaction such as 5Fe2+ + NO3 + 12H2O → 5Fe(OH)3 + 0.5N2 + 9H+. This process has since been demonstrated in detail with the hyperthermophilic archaeon Ferroglobus placidus (9) and with the mesophilic Proteobacteria Chromobacterium violacens strain 2002 (34) and Paracoccus ferrooxidans strain BDN-1 (16). Nitrate-dependent Fe(II) oxidation in the presence of fixed carbon has been documented for Dechlorosoma suillum strain PS (4), Geobacter metallireducens (7), Desulfitobacterium frappieri (23), and Acidovorax strain BoFeN1 (15). In addition to oxidizing insoluble Fe(II)-bearing minerals (33), the enrichment culture described by Straub et al. (28) is the only autotrophic Fe(II)-oxidizing, nitrate-reducing culture capable of near-complete oxidation of uncomplexed Fe(II) with reduction of nitrate to N2. During Fe(II) oxidation, F. placidus reduces nitrate to nitrite, which may play a significant role in overall Fe(II) oxidation. Although both C. violacens and Paracoccus ferrooxidans reduce nitrate to N2, C. violacens oxidizes only 20 to 30% of the initial Fe(II), and P. ferrooxidans uses FeEDTA2− but not free (uncomplexed) Fe(II) in medium analogous to that used for cultivation of the enrichment culture described by Straub et al. (28). The enrichment culture described by Straub et al. (28) is thus the most robust culture capable of autotrophic growth coupled to nitrate-dependent Fe(II) oxidation available at present. The composition and activity of this culture was investigated with molecular and cultivation techniques. The culture examined is one provided by K. L. Straub to E. E. Roden in 1998 for use in studies of nitrate-dependent oxidation of solid-phase Fe(II) compounds (33) and has been maintained in our laboratory since that time.  相似文献   

6.
Four stratified basins in Lake Kivu (Rwanda-Democratic Republic of the Congo) were sampled in March 2007 to investigate the abundance, distribution, and potential biogeochemical role of planktonic archaea. We used fluorescence in situ hybridization with catalyzed-reported deposition microscopic counts (CARD-FISH), denaturing gradient gel electrophoresis (DGGE) fingerprinting, and quantitative PCR (qPCR) of signature genes for ammonia-oxidizing archaea (16S rRNA for marine Crenarchaeota group 1.1a [MCG1] and ammonia monooxygenase subunit A [amoA]). Abundance of archaea ranged from 1 to 4.5% of total DAPI (4′,6-diamidino-2-phenylindole) counts with maximal concentrations at the oxic-anoxic transition zone (∼50-m depth). Phylogenetic analysis of the archaeal planktonic community revealed a higher level of richness of crenarchaeal 16S rRNA gene sequences (21 of the 28 operational taxonomic units [OTUs] identified [75%]) over euryarchaeotal ones (7 OTUs). Sequences affiliated with the kingdom Euryarchaeota were mainly recovered from the anoxic water compartment and mostly grouped into methanogenic lineages (Methanosarcinales and Methanocellales). In turn, crenarchaeal phylotypes were recovered throughout the sampled epipelagic waters (0- to 100-m depth), with clear phylogenetic segregation along the transition from oxic to anoxic water masses. Thus, whereas in the anoxic hypolimnion crenarchaeotal OTUs were mainly assigned to the miscellaneous crenarchaeotic group, the OTUs from the oxic-anoxic transition and above belonged to Crenarchaeota groups 1.1a and 1.1b, two lineages containing most of the ammonia-oxidizing representatives known so far. The concomitant vertical distribution of both nitrite and nitrate maxima and the copy numbers of both MCG1 16S rRNA and amoA genes suggest the potential implication of Crenarchaeota in nitrification processes occurring in the epilimnetic waters of the lake.Lake Kivu is a meromictic lake located in the volcanic region between Rwanda and the Democratic Republic of the Congo and is the smallest of the African Great Rift Lakes. The monimolimnion of the lake contains a large amount of dissolved CO2 and methane (300 km3 and 60 km3, respectively) as a result of geological and biological activity (24, 73, 85). This massive accumulation converts Lake Kivu into one of the largest methane reservoirs in the world and into a unique ecosystem for geomicrobiologists interested in the methane cycle and in risk assessment and management (34, 71, 72, 85). Comprehensive studies on the diversity and activity of planktonic populations of both large and small eukaryotes and their trophic interplay operating in the epilimnetic waters of the lake are available (33, 39, 49). Recent surveys have also provided a deeper insight into the seasonal variations of photosynthetic and heterotrophic picoplankton (67, 68), although very few data exist on the composition, diversity, and spatial distribution of bacterial and archaeal communities. In this regard, the studies conducted so far of the bacterial/archaeal ecology in Lake Kivu have been mostly focused on the implications on the methane cycle (34, 73), but none have addressed the presence and distribution of additional archaeal populations in the lake.During the last few years, microbial ecology studies carried out in a wide variety of habitats have provided compelling evidence of the ubiquity and abundance of mesophilic archaea (4, 10, 13, 19). Moreover, the discovery of genes encoding enzymes related to nitrification and denitrification in archaeal metagenomes from soil and marine waters (29, 86, 88) and the isolation of the first autotrophic archaeal nitrifier (40) demonstrated that some archaeal groups actively participate in the carbon and nitrogen cycles (56, 64, 69). In relation to aquatic environments, genetic markers of ammonia-oxidizing archaea (AOA) of the marine Crenarchaeota group 1.1a (MCG1) have consistently been found in water masses of several oceanic regions (6, 14, 17, 26, 28, 30, 37, 42, 51, 52, 89), estuaries (5, 9, 26, 53), coastal aquifers (26, 66), and stratified marine basins (15, 41, 44). Although less information is available for freshwater habitats, recent studies carried out in oligotrophic high-mountain and arctic lakes showed an important contribution of AOA in both the planktonic and the neustonic microbial assemblages (4, 61, 89).The oligotrophic nature of Lake Kivu and the presence of a well-defined redoxcline may provide an optimal niche for the development of autotrophic AOA populations. Unfortunately, no studies of the involvement of microbial planktonic populations in cycling nitrogen in the lake exist, and only data on the distribution of dissolved inorganic nitrogen species in relation to phytoplankton ecology (67, 68) and nutrient loading are available (54, 58). Our goals here were to ascertain whether or not archaeal populations other than methane-related lineages were relevant components of the planktonic microbial community and to determine whether the redox gradient imposed by the oxic-anoxic interphase acts as a threshold for their vertical distribution in epipelagic waters (0- to 100-m depth). To further explore the presence and potential activity of nitrifying archaeal populations in Lake Kivu, samples were analyzed for the abundance and vertical distribution of signature genes for these microorganisms, i.e., the 16S rRNA of MCG1 and the ammonia monooxygenase subunit A (amoA) gene by quantitative PCR.  相似文献   

7.
Giardia intestinalis is a ubiquitous intestinal protozoan parasite and has been proposed to represent the earliest diverging lineage of extant eukaryotes. Despite the importance of Giardia as a model organism, research on Giardia has been hampered by an inability to achieve cell cycle synchrony for in vitro cultures. This report details successful methods for attaining cell cycle synchrony in Giardia cultures. The research presented here demonstrates reversible cell cycle arrest in G1/S and G2/M with aphidicolin and nocodazole, respectively. Following synchronization, cells were able to recover completely from drug treatment and remained viable and maintained synchronous growth for 6 h. These techniques were used to synchronize Giardia cultures to increase the percentages of mitotic spindles in the cultures. This method of synchronization will enhance our ability to study cell cycle-dependent processes in G. intestinalis.Giardia intestinalis is a ubiquitous intestinal protozoan parasite causing disease in humans and animals worldwide (1, 11). In developing countries, diarrheal disease is responsible for 80% of the deaths of children under 2 years of age (21), and Giardia is one of the major causes of this condition. As a diplomonad, Giardia has been proposed to represent the earliest diverging lineage of extant eukaryotes, based on single rRNA and single and/or concatenated protein phylogenies developed by considering an archaeal out-group (2, 3, 5, 15, 23), making it a valuable organism for studying the evolution of biological processes in all eukaryotes. Characteristic of the order Diplomonadida, Giardia trophozoites contain two nuclei that remain separate during mitosis, with each daughter cell inheriting one copy of each parental nucleus (19). The trophozoite form, which attaches to the small intestine of the host, has a tetraploid (4N) DNA content in G1 since each nucleus is 2N (4). Following a round of DNA synthesis, each G2 nucleus is 4N, making the cell 8N. According to previous flow cytometry results, actively growing Giardia cultures spend the majority of the cell cycle in the G2/M phase and significantly less time in the G1 and S phases (4); in contrast, many tissue culture cells display a lengthy G1 phase. Until recently, an inability to synchronize in vitro Giardia cultures to any degree has severely hampered the ability of researchers to study cell cycle-dependent processes (16, 20).This work demonstrates successful cell cycle arrest by using nocodazole, a microtubule-destabilizing drug that leads to the depolymerization of spindle microtubules in Giardia (6, 20). A brief nocodazole treatment resulted in cells arrested early in mitosis or at the end of G2, presumably by the activation of a mitotic spindle checkpoint (22). G2 arrest using nocodazole was combined with G1 arrest using aphidicolin, a drug that presumably acts through the inhibition of polymerase-dependent DNA synthesis (8, 12, 14, 25). By combining these two treatments, we were able to effectively synchronize Giardia cultures while maintaining cell viability. These synchronization methods were used to enrich cultures with mitotic spindles at the M phase. Moreover, these methods will be a valuable tool for studying other aspects of Giardia biology such as encystation, the time in the life cycle when the trophozoite transforms into an infectious cyst.  相似文献   

8.
Bocavirus is a newly classified genus of the family Parvovirinae. Infection with Bocavirus minute virus of canines (MVC) produces a strong cytopathic effect in permissive Walter Reed/3873D (WRD) canine cells. We have systematically characterized the MVC infection-produced cytopathic effect in WRD cells, namely, the cell death and cell cycle arrest, and carefully examined how MVC infection induces the cytopathic effect. We found that MVC infection induces an apoptotic cell death characterized by Bax translocalization to the mitochondrial outer membrane, disruption of the mitochondrial outer membrane potential, and caspase activation. Moreover, we observed that the activation of caspases occurred only when the MVC genome was replicating, suggesting that replication of the MVC genome induces apoptosis. MVC infection also induced a gradual cell cycle arrest from the S phase in early infection to the G2/M phase at a later stage, which was confirmed by the upregulation of cyclin B1 and phosphorylation of cdc2. Cell cycle arrest at the G2/M phase was reproduced by transfection of a nonreplicative NS1 knockout mutant of the MVC infectious clone, as well as by inoculation of UV-irradiated MVC. In contrast with other parvoviruses, only expression of the MVC proteins by transfection did not induce apoptosis or cell cycle arrest. Taken together, our results demonstrate that MVC infection induces a mitochondrion-mediated apoptosis that is dependent on the replication of the viral genome, and the MVC genome per se is able to arrest the cell cycle at the G2/M phase. Our results may shed light on the molecular pathogenesis of Bocavirus infection in general.The Bocavirus genus is newly classified within the subfamily Parvovirinae of the family Parvoviridae (21). The currently known members of the Bocavirus genus include bovine parvovirus type 1 (BPV1) (17), minute virus of canines (MVC) (57), and the recently identified human bocaviruses (HBoV, HBoV2, and HBoV3) (4, 7, 36).MVC was first recovered from canine fecal samples in 1970 (10). The virus causes respiratory disease with breathing difficulty (14, 32, 49) and enteritis with severe diarrhea (11, 39), which often occurs with coinfection with other viruses (39), spontaneous abortion of fetuses, and death of newborn puppies (14, 29). Pathological lesions in fetuses in experimental infections were found in the lymphoid tissue of the lung and small intestine (14). MVC was isolated and grown in the Walter Reed/3873D (WRD) canine cell line (10), which is derived from a subdermoid cyst of an irradiated male dog (10). The full-length 5.4-kb genome of MVC was recently mapped with palindromic termini (60). Under the control of a single P6 promoter, through the mechanism of alternative splicing and alternative polyadenylation, MVC expresses two nonstructural proteins (NS1 and NP1) and two capsid proteins (VP1 and VP2). Like the NS1 proteins of other parvoviruses, the NS1 of MVC is indispensable for genome replication. The NP1 protein, which is unique to the Bocavirus genus, appears to be critical for optimal viral replication, as the NP1 knockout mutant of MVC suffers from severe impairment of replication (60). A severe cytopathic effect during MVC infection of WRD cells has been documented (10, 60).The HBoV genome has been frequently detected worldwide in respiratory specimens from children under 2 years old with acute respiratory illnesses (2, 34, 55). HBoV is associated with acute expiratory wheezing and pneumonia (3, 34, 55) and is commonly detected in association with other respiratory viruses (34, 55). Further studies are necessary, however, to identify potential associations of HBoV infection with clinical symptoms or disease of acute gastroenteritis (7, 36). The full-length sequence of infectious MVC DNA (GenBank accession no. FJ214110) that we have reported shows 52.6% identity to HBoV, while the NS1, NP1, and VP1 proteins are 38.5%, 39.9%, and 43.7% identical to those of HBoV, respectively (60).The cytopathic effect induced during parvovirus infection has been widely documented, e.g., in infections with minute virus of mice (MVM) (13), human parvovirus B19 (B19V) (58), parvovirus H-1 (25, 52), and BPV1 (1). In Bocavirus, cell death during BPV1 infection of embryonic bovine tracheal cells has been shown to be achieved through necrosis, independent of apoptosis (1). B19V-induced cell death of primary erythroid progenitor cells has been shown to be mainly mediated by an apoptotic pathway (58) in which the nonstructural protein 11kDa plays a key role (16). In contrast, the MVM-induced cytopathic effect has been revealed to be mediated by NS1 interference with intracellular casein kinase II (CKII) signaling (22, 44, 45), a nonapoptotic cell death. Oncolytic parvovirus H-1 infections can induce either apoptosis or nonapoptotic cell death, depending on the cell type (25, 40). Therefore, the mechanisms underlying parvovirus infection-induced cell death vary, although NS1 has been widely shown to be involved in both apoptotic and nonapoptotic cell death. The nature of the cytopathic effect during Bocavirus MVC infection has not been studied.Parvovirus replication requires infected cells at the S phase. Infection with parvovirus has been revealed to accompany a cell cycle perturbation that mostly leads to an arrest in the S/G2 phase or the G2/M phase during infection (30, 33, 42, 47, 65). MVM NS1 expression induces an accumulation of sensitive cells in the S/G2 phase (6, 46, 47). Whether MVC infection-induced cell death is accompanied by an alternation of cell cycle progression and whether the viral nonstructural protein is involved in these processes have not been addressed.In this study, we found, in contrast with other members of the family Parvoviridae, expression of both the nonstructural and structural proteins of MVC by transfection did not induce cell death or cell cycle arrest. However, the cytopathic effect induced during MVC infection is a replication-coupled, mitochondrion-mediated and caspase-dependent apoptosis, accompanied with a gradual cell cycle arrest from the S phase to the G2/M phase, which is facilitated by the MVC genome.  相似文献   

9.
10.
Bioreactor cultures of Escherichia coli recombinants carrying phaBAC and phaP of Azotobacter sp. FA8 grown on glycerol under low-agitation conditions accumulated more poly(3-hydroxybutyrate) (PHB) and ethanol than at high agitation, while in glucose cultures, low agitation led to a decrease in PHB formation. Cells produced smaller amounts of acids from glycerol than from glucose. Glycerol batch cultures stirred at 125 rpm accumulated, in 24 h, 30.1% (wt/wt) PHB with a relative molecular mass of 1.9 MDa, close to that of PHB obtained using glucose.Polyhydroxyalkanoates (PHAs), accumulated as intracellular granules by many bacteria under unfavorable conditions (5, 8), are carbon and energy reserves and also act as electron sinks, enhancing the fitness of bacteria and contributing to redox balance (9, 11, 19). PHAs have thermoplastic properties, are totally biodegradable by microorganisms present in most environments, and can be produced from different renewable carbon sources (8).Poly(3-hydroxybutyrate) (PHB) is the best known PHA, and its accumulation in recombinant Escherichia coli from several carbon sources has been studied (1, 13). In the last few years, increasing production of biodiesel has caused a sharp fall in the cost of its main by-product, glycerol (22). Its use for microbial PHA synthesis has been analyzed for natural PHA producers, such as Methylobacterium rhodesianum, Cupriavidus necator (formerly called Ralstonia eutropha) (3), several Pseudomonas strains (22), the recently described bacterium Zobellella denitrificans (7), and a Bacillus sp. (18), among others. Glycerol has also been used for PHB synthesis in recombinant E. coli (12, 15). PHAs obtained from glycerol were reported to have a significantly lower molecular weight than polymer synthesized from other substrates, such as glucose or lactose (10, 23).Apart from the genes that catalyze polymer biosynthesis, natural PHA producers have several genes that are involved in granule formation and/or have regulatory functions, such as phasins, granule-associated proteins that have been shown to enhance polymer synthesis and the number and size of PHA granules (17, 24). The phasin PhaP has been shown to exert a beneficial effect on bacterial growth and PHB accumulation from glycerol in bioreactor cultures of strain K24KP, a recombinant E. coli that carries phaBAC and phaP of Azotobacter sp. FA8 (6).Because the redox state of the cells is known to affect the synthesis of PHB (1, 4, 14), the present study investigates the behavior of this recombinant strain under different aeration conditions, by using two substrates, glucose and glycerol, with different oxidation states.  相似文献   

11.
12.
13.
14.
The bivalve Codakia orbicularis, hosting sulfur-oxidizing gill endosymbionts, was starved (in artificial seawater filtered through a 0.22-μm-pore-size membrane) for a long-term experiment (4 months). The effects of starvation were observed using transmission electron microscopy, fluorescence in situ hybridization and catalyzed reporter deposition (CARD-FISH), and flow cytometry to monitor the anatomical and physiological modifications in the gill organization of the host and in the symbiotic population housed in bacteriocytes. The abundance of the symbiotic population decreased through starvation, with a loss of one-third of the bacterial population each month, as shown by CARD-FISH. At the same time, flow cytometry revealed significant changes in the physiology of symbiotic cells, with a decrease in cell size and modifications to the nucleic acid content, while most of the symbionts maintained a high respiratory activity (measured using the 5-cyano-2,3-ditolyl tetrazolium chloride method). Progressively, the number of symbiont subpopulations was reduced, and the subsequent multigenomic state, characteristic of this symbiont in freshly collected clams, turned into one and five equivalent genome copies for the two remaining subpopulations after 3 months. Concomitant structural modifications appeared in the gill organization. Lysosymes became visible in the bacteriocytes, while large symbionts disappeared, and bacteriocytes were gradually replaced by granule cells throughout the entire lateral zone. Those data suggested that host survival under these starvation conditions was linked to symbiont digestion as the main nutritional source.The entire marine Lucinidae family, found in a wide range of sulfidic habitats, lives in association with chemoautotrophic sulfide-oxidizing bacterial symbionts, generally hosted in the gills of the bivalve. Lucinids are usually found in shallow water, such as intertidal mud or seagrasses (4, 53), in deeper water, e.g., Bathyaustriella thionipta (30), and in deep oceans at a 2,000-m depth, i.e., Lucinoma kazani (21, 55). The chemoautotrophic endosymbionts involved in such relationships are always localized inside specialized cells called bacteriocytes, and they have been found in several genera of the Lucinidae family, such as Codakia (4, 28), Loripes (39, 43), Lucina, and Lucinoma (17). Sulfur granules inside the symbiont cytoplasm have been demonstrated in most of the investigated species. The intracellular symbionts take energy from the oxidation of reduced sulfur compounds (27, 56, 59) and synthesize organic molecules by CO2 fixation in a Calvin-Benson cycle, translocated to the host (18). This relationship between the host and its symbionts represents the autotrophic pathway for host nutrition (27). It has also been suggested that in symbiotic bivalves, intracellular digestion of the symbionts may be a nutrient source for the host, based on studies of hydrothermal vent and shallow water bivalves (6, 26, 39).The relative importance of the autotrophic versus heterotrophic nutritional pathway can be estimated by measuring the carbon isotope (δ-13C) ratios in the host tissue. Measuring this δ-13C ratio on a wide range of invertebrates suggested that bivalves, including members of the Lucinidae, that live in reduced sediment may obtain a significant proportion of their organic carbon from chemoautotrophic endosymbionts (4, 10, 51, 52, 57). This suggestion was in agreement with the reduced functional digestive system previously described for the Lucinidae family (2, 53). Structural and morphological studies of gills of a few lucinids (belonging to the genera Lucina and Lucinoma) strongly suggested that symbionts play an important role in host nutrition since they occupy about 30% of the gill tissue and produce most of the host energy (17). Nevertheless, an alternative pathway for feeding, i.e., heterotrophic particulate feeding, could occur in some of the lucinid bivalves, since diatoms were found in the stomach of some lucinids (57). Duplessis et al. (22) showed that particulate feeding could be an important part of the nutritional strategy in symbiont-bearing Lucinoma, as opposed to the anatomical features that gave the impression that this bivalve relied only on symbiont nutrition.In natural habitats, chemoautotrophic bivalves live at the interface between an anoxic sulfide-generating zone and water column oxygenated sediment. However, even if they are not close to a vent, these symbiotic organisms often have to deal with environments that are periodically depleted of oxygen (5, 12) and with extremely low sulfide concentration (13, 15, 16). These natural environmental variations lead to annual and seasonal changes in the δ-13C ratio, as observed for some thyasirid species (13, 14). This δ-13C ratio variation may be assumed to correspond to the capability of the host to rely on both autotrophic and heterotrophic pathways, and the preponderance of one pathway versus the other in the mixotrophic diet has been considered to be the way in which these organisms deal with changes in the chemical composition of their environment.Apart from the decrease in symbiont abundance suggested by transmission electron microscopy (TEM) analysis and a decrease in sulfur and protein content in the gill tissue of thyasirids (20, 38, 40), little is known about the physiological status of these symbionts and the changes undergone by the symbiotic population of starved bivalves. A previous study of the population was carried out under natural conditions with Codakia orbicularis, a chemoautotrophic bivalve. This tropical bivalve lives in shallow-water sediment among the roots of seagrasses (Thalassia testudinum) (1). Like all lucinids studied so far, it is associated with sulfur-oxidizing symbionts (4, 27, 28) containing elemental sulfur in their cytoplasm (42). The bacterial symbiont of C. orbicularis, environmentally transmitted to the host (31), belongs to a single taxonomic group (Gammaproteobacteria) (25) and is shared by several other tropical lucinids (24, 25, 34, 35). Only a few data are available on the physiology of this symbiont. It was characterized by the presence of Rubisco and ATP sulfurylase enzymes and a δ-13C ratio typical of chemoautotrophic bivalves (4). Unlike other related thyasirids tested for nitrate respiration even under oxygenated conditions (37), the symbionts of C. orbicularis use oxygen as the primary electron acceptor (23). Initial investigations of the population of Codakia orbicularis'' symbiont revealed that the symbiotic population hosted by freshly collected individuals contained a high proportion of large bacterial cells containing multiple copies of their genome, typical of actively growing cells, despite the absence of dividing cells (11). It was assumed that the host could maintain a pure culture of the symbiont inside the bacteriocytes by regulating the entry and growth of newly recruited symbionts from sediment and probably regulating symbiont densities by host digestion (11).This study was undertaken to investigate the dynamics of the symbiotic population hosted by C. orbicularis under experimental conditions based on long-term starvation of bivalves, i.e., incubated without planktonic food. We set out the ultrastructural, structural, and physiological changes that occurred in the symbiont population by examining the host gill sections using TEM and fluorescence in situ hybridization and catalyzed reporter deposition (CARD-FISH). A purified fraction of gill endosymbionts was analyzed by flow cytometry (FCM) to investigate the nucleic acid content and cell size of symbionts and by using the 5-cyano-2,3-ditolyl tetrazolium chloride (CTC) method and epifluorescence microscopy to detect the respiratory activity of symbionts. The modifications induced by host starvation in the symbiotic population are described for the period of long-term starvation.  相似文献   

15.
16.
17.
Iron oxidation at neutral pH by the phototrophic anaerobic iron-oxidizing bacterium Rhodobacter sp. strain SW2 leads to the formation of iron-rich minerals. These minerals consist mainly of nano-goethite (α-FeOOH), which precipitates exclusively outside cells, mostly on polymer fibers emerging from the cells. Scanning transmission X-ray microscopy analyses performed at the C K-edge suggest that these fibers are composed of a mixture of lipids and polysaccharides or of lipopolysaccharides. The iron and the organic carbon contents of these fibers are linearly correlated at the 25-nm scale, which in addition to their texture suggests that these fibers act as a template for mineral precipitation, followed by limited crystal growth. Moreover, we evidence a gradient of the iron oxidation state along the mineralized fibers at the submicrometer scale. Fe minerals on these fibers contain a higher proportion of Fe(III) at cell contact, and the proportion of Fe(II) increases at a distance from the cells. All together, these results demonstrate the primordial role of organic polymers in iron biomineralization and provide first evidence for the existence of a redox gradient around these nonencrusting, Fe-oxidizing bacteria.Fe(II) can serve as a source of electrons for phylogenetically diverse microorganisms that precipitate iron minerals as products of their metabolism (see, e.g., references 3, 5, 25, and 30). For example, mixotrophic or autotrophic bacteria can couple the oxidation of Fe(II) to the reduction of nitrate in anoxic and neutral-pH environments. With Fe(III) being highly insoluble at neutral pH, this metabolism leads to the formation of poorly to well-crystallized iron minerals (3, 18, 26, 27) that precipitate partly within the cell periplasm for some strains (22). Similar Fe minerals are also synthesized by autotrophic bacteria that perform anoxygenic photosynthesis, using Fe(II) as an electron donor and light as a source of energy for CO2 fixation (8, 12, 30), according to the equation HCO3 + 4 Fe2+ + 10 H2O ⇆ <CH2O> + 4 Fe(OH)3 + 7 H+.However, the biological mechanisms of iron oxidation in these bacteria and in particular the way they cope with the formation of minerals within their ultrastructures are still not fully understood. Indeed, iron minerals are potentially lethal since their precipitation may alter cellular ultrastructures but also catalyze the production of free radicals (2). Recent genetic studies of the phototrophic, iron-oxidizing bacteria Rhodobacter sp. strain SW2 (6) and Rhodopseudomonas palustris strain TIE-1 (16) have identified genes (fox and pio operons, respectively) encoding proteins specific for iron oxidation. Interestingly, Jiao and Newman (16) suggested that one of these proteins could have a periplasmic localization. However, in contrast to what has been observed in some other phototrophic iron oxidizers (25) and in some nitrate-reducing, iron-oxidizing bacteria (22), no iron-mineral precipitation occurs within the periplasm of the purple nonsulfur iron-oxidizing bacterium Rhodobacter sp. strain SW2 (3). Similarly to some other anaerobic neutrophilic (22, 25) and microaerobic iron-oxidizing bacteria (5, 10), this strain seems indeed to have the ability to localize iron biomineralization at a distance from the cells, leaving large areas of the cells free of precipitates (17, 25). While it has been shown that the Gallionella and Leptothrix genera, for example, produce extracellular polymers that facilitate the nucleation of iron minerals outside cells (see, e.g., references 5 and 9), only a little is known about the existence and function of such polymers in anaerobic, neutrophilic iron-oxidizing bacteria and particularly in the phototrophic strain SW2. In the present study, we investigate iron biomineralization by the photoautotrophic iron-oxidizing bacterium Rhodobacter sp. strain SW2. We use scanning transmission X-ray microscopy (STXM) to map and identify organic polymers produced by the cells as well as the redox state of iron at the 25-nanometer scale regularly during a 2 week-period. These results demonstrate the primordial role of organic polymers in iron biomineralization and provide the first evidence for the existence of a redox gradient around SW2 cells.  相似文献   

18.
19.
Previous studies have documented the capacity of European earthworms belonging to the family Lumbricidae to emit the greenhouse gas nitrous oxide (N2O), an activity attributed primarily to the activation of ingested soil denitrifiers. To extend the information base to earthworms in the Southern Hemisphere, four species of earthworms in New Zealand were examined for gut-associated denitrification. Lumbricus rubellus and Aporrectodea rosea (introduced species of Lumbricidae) emitted N2O, whereas emission of N2O by Octolasion cyaneum (an introduced species of Lumbricidae) and emission of N2O by Octochaetus multiporus (a native species of Megascolecidae) were variable and negligible, respectively. Exposing earthworms to nitrite or nitrate and acetylene significantly increased the amount of N2O emitted, implicating denitrification as the primary source of N2O and indicating that earthworms emitted dinitrogen (N2) in addition to N2O. The alimentary canal displayed a high capacity to produce N2O when it was supplemented with nitrite, and alimentary canal contents contained large amounts of carbohydrates and organic acids indicative of fermentation (e.g., succinate, acetate, and formate) that could serve as sources of reductant for denitrification. nosZ encodes a portion of the terminal oxidoreductase used in denitrification. The nosZ sequences detected in the alimentary canals of L. rubellus and O. multiporus were similar to those retrieved from soil and were distantly related to sequences of uncultured soil bacteria and genera common in soils (i.e., Bradyrhizobium, Azospirillum, Rhodopseudomonas, Rhodospirillum, Pseudomonas, Oligotropha, and Sinorhizobium). These findings (i) suggest that the capacity to emit N2O and N2 is a general trait of earthworms and not geographically restricted, (ii) indicate that species belonging to different earthworm families (i.e., Megascolecidae and Lumbricidae) may not have equal capacities to emit N2O, and (iii) also corroborate previous findings that link this capacity to denitrification in the alimentary canal.Earthworms are dominant members of the soil fauna and affect the structure and fertility of soils (5, 20, 22, 23). Various species of European earthworms belonging to the family Lumbricidae (e.g., Aporrectodea caliginosa, Lumbricus rubellus, and Octolasion lacteum) emit dinitrogen (N2) and the greenhouse gas nitrous oxide (N2O), and their burrowing activities and feeding habits in combination with in situ conditions can influence the emission of nitrogenous gases from soils that they inhabit (1, 2, 13, 17, 25, 27, 39).The microbiology of the earthworm alimentary canal has been addressed in numerous studies (3, 4, 6, 9, 14, 16, 32). The alimentary canal of the earthworm is anoxic, in marked contrast to the aerated material that earthworms ingest (14, 39). Anoxia and other in situ conditions of the alimentary canal appear to stimulate soil microbes capable of surviving under anaerobic conditions during passage through the gut (3, 4). Soils are rich in denitrifying bacteria (37), and the capacity of European earthworms to emit nitrogenous gases has been attributed primarily to the in situ activity of ingested denitrifying bacteria that appear to be highly active under the anoxic conditions of the earthworm alimentary canal (12, 15, 17, 25, 39). However, it is not known if the capacity to emit nitrogenous gases is a general trait of earthworms independent of their taxonomic family or geographic location. The main objectives of this study were to examine the capacity of Southern Hemisphere earthworms in New Zealand to emit N2O and to determine if this capacity was linked to denitrifying bacteria in the alimentary canal.  相似文献   

20.
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