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.
In silico analysis of group 4 [NiFe]-hydrogenases from a hyperthermophilic archaeon, Thermococcus onnurineus NA1, revealed a novel tripartite gene cluster consisting of dehydrogenase-hydrogenase-cation/proton antiporter subunits, which may be classified as the new subgroup 4b of [NiFe]-hydrogenases-based on sequence motifs.Hydrogenases are the key enzymes involved in the metabolism of H2, catalyzing the following chemical reaction: 2H+ + 2e ↔ H2. Hydrogenases can be classified into [NiFe]-hydrogenases, [FeFe]-hydrogenases, and [Fe]-hydrogenases, based on their distinctive functional core containing the catalytic metal center (11, 17).The genomic analysis of Thermococcus onnurineus NA1, a hyperthermophilic archaeon isolated from a deep-sea hydrothermal vent area, revealed the presence of several distinct gene clusters encoding seven [NiFe]-hydrogenases and one homolog similar to Mbx (membrane-bound oxidoreductase) from Pyrococcus furiosus (1, 6, 8, 12). According to the classification system of hydrogenases by Vignais et al. (17), three hydrogenases (one F420-reducing and two NADP-reducing hydrogenases) belong to group 3 [NiFe]-hydrogenases, and four hydrogenases belong to group 4 [NiFe]-hydrogenases. The group 4 hydrogenases are widely distributed among bacteria and archaea (17), with Hyc and Hyf (hydrogenase 3 and 4, respectively) from Escherichia coli (19), Coo (CO-induced hydrogenase) from Rhodospirillum rubrum (4), Ech (energy-converting hydrogenase) from Methanosarcina barkeri (7), and Mbh (membrane-bound hydrogenase) from P. furiosus (6, 10, 12) being relatively well-characterized hydrogenases in this group. One of the four group 4 hydrogenases from T. onnurineus NA1 was found to be similar in sequence to that of P. furiosus Mbh (10).  相似文献   

2.
Using a metagenomics approach, we have cloned a piece of environmental DNA from the Sargasso Sea that encodes an [NiFe] hydrogenase showing 60% identity to the large subunit and 64% to the small subunit of a Thiocapsa roseopersicina O2-tolerant [NiFe] hydrogenase. The DNA sequence of the hydrogenase identified by the metagenomic approach was subsequently found to be 99% identical to the hyaA and hyaB genes of an Alteromonas macleodii hydrogenase, indicating that it belongs to the Alteromonas clade. We were able to express our new Alteromonas hydrogenase in T. roseopersicina. Expression was accomplished by coexpressing only two accessory genes, hyaD and hupH, without the need to express any of the hyp accessory genes (hypABCDEF). These results suggest that the native accessory proteins in T. roseopersicina could substitute for the Alteromonas counterparts that are absent in the host to facilitate the assembly of a functional Alteromonas hydrogenase. To further compare the complex assembly machineries of these two [NiFe] hydrogenases, we performed complementation experiments by introducing the new Alteromonas hyaD gene into the T. roseopersicina hynD mutant. Interestingly, Alteromonas endopeptidase HyaD could complement T. roseopersicina HynD to cleave endoproteolytically the C-terminal end of the T. roseopersicina HynL hydrogenase large subunit and activate the enzyme. This study refines our knowledge on the selectivity and pleiotropy of the elements of the [NiFe] hydrogenase assembly machineries. It also provides a model for functionally analyzing novel enzymes from environmental microbes in a culture-independent manner.Hydrogen is a promising energy carrier for the future (10). Photosynthetic microbes such as cyanobacteria have attracted considerable attention, because they can split water photolytically to produce H2. However, one major drawback of the processes is that their H2-evolving hydrogenases are extremely sensitive to O2, which is an inherent by-product of oxygenic photosynthesis. Thus, transfer of O2-tolerant [NiFe] hydrogenases into cyanobacteria might be one approach to overcome this O2 sensitivity issue. A small number of O2-tolerant hydrogenases has been identified (9, 21, 47). However, they tend to favor H2 uptake over evolution. Searching for novel O2-tolerant [NiFe] hydrogenases from environmental microbes therefore becomes an important part of the effort to construct such biophotolytic systems.The oceans harbor an abundance of microorganisms with H2 production capability. Traditionally, new hydrogenases have been screened only from culturable organisms. However, since only a few microbes can be cultured (14), many of them have not been identified, and their functions remain unknown. Metagenomics is a rapidly growing field, which allows us to obtain information about uncultured microbes and to understand the true diversity of microbes in their natural environments. Metagenomics analysis provides a completely new approach for identifying novel [NiFe] hydrogenases from the oceans in a culture-independent manner. The Global Ocean Sampling (GOS) expedition has produced the largest metagenomic data set to date, providing a rich catalog of proteins and protein families, including those enzymes involved in hydrogen metabolism (45, 52, 56-58). Putative novel [NiFe] hydrogenase enzymes that were identified from marine microbial metagenomic data in these expeditions can be examined to find potentially important new hydrogenases. Because source organisms for metagenomic sequences are not typically known, these hydrogenases have to be heterologously expressed in culturable foreign hosts for protein and functional analyses.Unlike most proteins, hydrogenases have a complex architecture and must be assembled and matured through a multiple-step process (7, 11). Hydrogenases are divided into three distinct groups based on their metal contents (54): Fe-S cluster-free hydrogenases (22, 23, 48), [FeFe] hydrogenases (1, 12, 25), and [NiFe] hydrogenases (2, 3, 55). [NiFe] hydrogenases are heterodimers composed of a large subunit and a small subunit, and their NiFe catalytic centers are located in the large subunits (2, 15, 19, 40). A whole set of accessory proteins are required to properly assemble the catalytic centers (7). The accessory protein HypE first interacts with HypF to form a HypF-HypE complex, and the carbamyl group linked to HypF is then dehydrated by HypE in the presence of ATP to release the CN group that is transferred to iron through a HypC-HypD-HypE complex (6). The origin of the CO ligand that is also bound to the iron is not clear, and possibly it comes from formate, formyl-tetrahydrofolate, or acetate. The liganded Fe atom is inserted into the immature large subunit, in which HypC proteins function as chaperones to facilitate the metal insertion (5, 34, 36). Ni is delivered to the catalytic center by the zinc-metalloenzyme HypA that interacts with HypB, a nickel-binding and GTP-hydrolyzing protein. The final step in the maturation process is endoproteolytic cleavage. Once the nickel is transferred to the active site, the endopeptidase, such as HyaD or HynD, cleaves the C-terminal end of the large subunit (33, 43), which triggers a conformational change of the protein so that the Ni-Fe catalytic center can be internalized.Heterologous expression of functional [NiFe] hydrogenases has been demonstrated in several studies (4, 18, 31, 39, 44, 50), suggesting that it could be a feasible approach to express novel hydrogenases from the environment for functional analysis. In this study, we sought to prove the concept that metagenomically derived environmental DNA can give rise to a functional [NiFe] hydrogenase through expression in a foreign host and that novel [NiFe] hydrogenases from environmental microbes can be studied in a culture-independent manner. We cloned environmental DNA that harbors the genes of a putative novel hydrogenase that shows strong homology to a known O2-tolerant hydrogenase, HynSL, from the phototrophic purple sulfur bacterium Thiocapsa roseopersicina (21, 28, 41, 59). We heterologously expressed the two structural genes (hyaA and hyaB) and two accessory genes (hupH and hyaD) of this novel environmental hydrogenase in T. roseopersicina, a foreign host that may already have the necessary machinery required to process the environmental hydrogenase since it carries the homologous hydrogenase HynSL. We analyzed the new hydrogenase protein and its functions. In addition, we compared the maturation mechanisms between the two homolog hydrogenases by performing complementation experiments.  相似文献   

3.
Newly designed primers for [Fe-Fe]-hydrogenases indicated that (i) fermenters, acetogens, and undefined species in a fen harbor hitherto unknown hydrogenases and (ii) Clostridium- and Thermosinus-related primary fermenters, as well as secondary fermenters related to sulfate or iron reducers might be responsible for hydrogen production in the fen. Comparative analysis of [Fe-Fe]-hydrogenase and 16S rRNA gene-based phylogenies indicated the presence of homologous multiple hydrogenases per organism and inconsistencies between 16S rRNA gene- and [Fe-Fe]-hydrogenase-based phylogenies, necessitating appropriate qualification of [Fe-Fe]-hydrogenase gene data for diversity analyses.Molecular hydrogen (H2) is important in intermediary ecosystem metabolism (i.e., processes that link input to output) in wetlands (7, 11, 12, 33) and other anoxic habitats like sewage sludges (34) and the intestinal tracts of animals (9, 37). H2-producing fermenters have been postulated to form trophic links to H2-consuming methanogens, acetogens (i.e., organisms capable of using the acetyl-coenzyme A [CoA] pathway for acetate synthesis) (7), Fe(III) reducers (17), and sulfate reducers in a well-studied moderately acidic fen in Germany (11, 12, 16, 18, 22, 33). 16S rRNA gene analysis revealed the presence of Clostridium spp. and Syntrophobacter spp., which represent possible primary and secondary fermenters, as well as H2 producers in this fen (11, 18, 33). However, H2-producing bacteria are polyphyletic (30, 31, 29). Thus, a structural marker gene is required to target this functional group by molecular methods. [Fe-Fe]-hydrogenases catalyze H2 production in fermenters (19, 25, 29, 30, 31), and genes encoding [Fe-Fe]-hydrogenases represent such a marker gene. The objectives of this study were to (i) develop primers specific for highly diverse [Fe-Fe]-hydrogenase genes, (ii) analyze [Fe-Fe]-hydrogenase genes in pure cultures of fermenters, acetogens, and a sulfate reducer, (iii) assess [Fe-Fe]-hydrogenase gene diversity in H2-producing fen soil enrichments, and (iv) evaluate the limitations of the amplified [Fe-Fe]-hydrogenase fragment as a phylogenetic marker.  相似文献   

4.
Three functional NiFe hydrogenases were previously characterized in Thiocapsa roseopersicina BBS: two of them are attached to the periplasmic membrane (HynSL and HupSL), and one is localized in the cytoplasm (HoxEFUYH). The ongoing genome sequencing project revealed the presence of genes coding for another soluble Hox-type hydrogenase enzyme (hox2FUYH). Hox2 is a heterotetrameric enzyme; no indication for an additional subunit was found. Detailed comparative in vivo and in vitro activity and expression analyses of HoxEFUYH (Hox1) and the newly discovered Hox2 enzyme were performed. Functional differences between the two soluble NiFe hydrogenases were disclosed. Hox1 seems to be connected to both sulfur metabolism and dark/photofermentative processes. The bidirectional Hox2 hydrogenase was shown to be metabolically active under specific conditions: it can evolve hydrogen in the presence of glucose at low sodium thiosulfate concentration. However, under nitrogen-fixing conditions, it can oxidize H2 but less than the other hydrogenases in the cell.Hydrogenases are metalloenzymes involved in microbial hydrogen metabolism. A great variety of them have been identified and studied in various microorganisms and grouped on the basis of their metal content as NiFe, FeFe, and iron-sulfur cluster free hydrogenases (10, 42, 43). The basic protein structure of NiFe hydrogenases is heterodimeric, while FeFe hydrogenases are mostly composed of a single amino acid chain with multiple iron-sulfur clusters (28, 43, 44). Well-defined maturation proteins assist for the assembly and activation of hydrogenase enzymes; NiFe hydrogenases require a more complex accessory machinery than FeFe enzymes (2, 3, 24).Thiocapsa roseopersicina BBS is a photosynthetic purple sulfur bacterium belonging to the Chromatiaceae family (4). It prefers to utilize reduced sulfur compounds for anaerobic photochemolithoautotrophic growth, but simple organic substrates such as glucose or acetate can be also used as extra carbon, energy, and electron sources. It can be cultivated under aerobic (nonphotosynthetic) conditions in the presence of organic compounds. In the absence of other nitrogen sources, it is able to fix molecular nitrogen; this process is accompanied by H2 production. T. roseopersicina was earlier shown to possess at least three NiFe hydrogenases varying in their in vivo functions, localizations, and compositions. Hyn and Hup hydrogenases are attached to the membrane facing the periplasmic side (6, 18, 30). Hyn is a bidirectional enzyme with extraordinary stability (17). Recent study has demonstrated that the HynSL subunits are physiologically connected to cellular redox processes via the Isp1 and Isp2 proteins, which play an essential role in electron transfer (27). The second membrane-associated enzyme, Hup, is involved in H2 oxidation and shows homology to uptake hydrogenases, which recycle H2 produced by the nitrogenase enzyme complex or present in the environment. Next to the hydrogenase small and large subunits (HupSL), a b-type cytochrome, HupC, was demonstrated to be part of the in vivo active enzyme as a transmitter of electrons to the quinone pool (27). In several bacteria, e.g., Rhodobacter capsulatus (7) and Ralstonia eutropha (15, 20), the expression of the hydrogenase(s) was shown to be regulated by the hydrogen level in the environment. The genes encoding the hydrogen-sensing system also exist in T. roseopersicina (hupUV, hupT, and hupR), but the hupTUV genes proved to be silent in the wild-type strain—only hupR is expressed—which is why expression of hupSL genes is constitutive (16).A Hox-type soluble hydrogenase was also identified in T. roseopersicina (31); it is a representative of the bidirectional heteromultimeric cytoplasmic NiFe hydrogenases (37, 39). Enzymes belonging to this group are basically composed of two moieties: hydrogenase (HoxYH) and diaphorase (HoxFU) heterodimers. Additional subunits were identified in few cases. In R. eutropha H16, two HoxI proteins completing the Hox complex were suggested to provide a binding domain for NADPH (5). HoxE has been identified as the fifth subunit of heteropentameric NAD+-reducing Hox hydrogenases in several cyanobacteria, Allochromatium vinosum and T. roseopersicina (21, 31, 37). In-frame deletion of the hoxE gene ceased both the H2-producing and -oxidizing activities of Hox in vivo, but these were not affected in vitro. Consequently, an electron transfer role of the HoxE subunit was suggested (31, 32).The possibility of the presence of further hydrogenases in T. roseopersicina was noted few years ago (31). In the hynSL hupSL hoxH triple-mutant strain (GB112131), a small in vivo and in vitro hydrogenase activity could be measured under photomixotrophic growth conditions (both CO2 and organic compounds are used for growth) at the late growth phase. This residual activity could not be detected in the hypF mutant strain (M539). Since HypF protein has an essential role in the maturation process of all NiFe hydrogenases (9), these results suggested the presence of a previously unknown hydrogenase. Here we describe the identification and characterization of the second Hox-type hydrogenase, emphasizing the functional similarities and differences between the two soluble enzymes of this bacterium. In order to distinguish between the two Hox-type enzymes unequivocally, the HoxEFUYH complex will be renamed Hox1 and the newly described Hox2FUYH enzyme is called Hox2.  相似文献   

5.
6.
The facultative intracellular pathogen Salmonella enterica serovar Typhimurium relies on its Salmonella pathogenicity island 2 (SPI2) type III secretion system (T3SS) for intracellular replication and virulence. We report that the oxidoreductase thioredoxin 1 (TrxA) and SPI2 are coinduced for expression under in vitro conditions that mimic an intravacuolar environment, that TrxA is needed for proper SPI2 activity under these conditions, and that TrxA is indispensable for SPI2 activity in both phagocytic and epithelial cells. Infection experiments in mice demonstrated that SPI2 strongly contributed to virulence in a TrxA-proficient background whereas SPI2 did not affect virulence in a trxA mutant. Complementation analyses using wild-type trxA or a genetically engineered trxA coding for noncatalytic TrxA showed that the catalytic activity of TrxA is essential for SPI2 activity in phagocytic cells whereas a noncatalytic variant of TrxA partially sustained SPI2 activity in epithelial cells and virulence in mice. These results show that TrxA is needed for the intracellular induction of SPI2 and provide new insights into the functional integration between catalytic and noncatalytic activities of TrxA and a bacterial T3SS in different settings of intracellular infections.In Escherichia coli, thioredoxin 1 (TrxA, encoded by trxA) is an evolutionary conserved 11-kDa cytosolic highly potent reductase that supports the activities of various oxidoreductases and ribonucleotide reductases (1, 29) and interacts with a number of additional cytoplasmic proteins through the formation of temporary covalent intermolecular disulphide bonds (32). Consequently, as trxA mutants of E. coli (51), Helicobacter pylori (13), and Rhodobacter sphaeroides (34) show increased sensitivity to hydrogen peroxide, TrxA has been defined as a significant oxidoprotectant. In addition, TrxA possess a protein chaperone function that is disconnected from cysteine interactions (30, 32).Salmonella enterica serovar Typhimurium is closely related to E. coli. During divergent evolution, the Salmonella genome acquired a number of virulence-associated genes (20). Many of these genes are clustered on genetic regions termed Salmonella pathogenicity islands (or SPIs). Of these, SPI1 and SPI2 code for separate type III secretion systems (T3SSs). T3SSs are supramolecular virulence-associated machineries that, in several pathogenic gram-negative bacterial species, enable injection of effector proteins from the bacteria into host cells (22, 57). The effector proteins, in turn, manipulate intrinsic host cell functions to facilitate the infection.The SPI1 T3SS of S. serovar Typhimurium is activated for expression in the intestine in response to increased osmolarity and decreased oxygen tension (22, 57). SPI1 effector proteins are primarily secreted into cells that constitute the epithelial layer and interfere with host cell Cdc42 and Rac-1 signaling and actin polymerization. This enables the bacteria to orchestrate their own actin-dependent uptake into nonphagocytic cells (57). SPI1 effector proteins also induce inflammatory signaling and release of interleukin-1β from infected cells (25, 26).Subsequent systemic progression of S. serovar Typhimurium from the intestinal tissue relies heavily on an ability to survive and replicate in phagocytic cells (18, 46, 53, 54). S. serovar Typhimurium uses an additional set of effector proteins secreted by the SPI2 T3SS for replication inside host cells and for coping with phagocyte innate responses to the infection (10, 11, 54). The functions of SPI2 effectors include diversion of vesicular trafficking, induction of apoptotic responses, and manipulation of ubiquitination of host proteins (28, 40, 45, 53). Hence, SPI2 effector proteins create a vacuolar environment that sustains intracellular replication of S. serovar Typhimurium (28).In addition to pathogenicity islands, the in vivo fitness of Salmonella spp. relies on selected functions shared with other enterobacteria. Thus, many virulence genes are integrated into “housekeeping” gene regulatory networks, coded for by a core genome, which steer bacterial stress responses (12, 17, 27, 55). Selected anabolic pathways also contribute to virulence of S. serovar Typhimurium (18, 27), evidently by providing biochemical building blocks for bacterial replication (36).In S. serovar Typhimurium, TrxA is a housekeeping protein that strongly contributes to virulence in cell culture and mouse infection models (8). However, the mechanism by which TrxA activity adds to virulence has not been defined. Here we show that the contribution of TrxA to virulence of S. serovar Typhimurium associates with its functional integration with the SPI2 T3SS under conditions that prevail in the intracellular vacuolar compartment of the host cell. These findings ascribe a novel role to TrxA in bridging environmental adaptations with virulence gene expression and illuminate a new aspect of the interaction between evolutionary conserved and horizontally acquired gene functions in bacteria.  相似文献   

7.
8.
Kinetic analyses of bacterial growth, carbohydrate consumption, and metabolite production of five butyrate-producing clostridial cluster XIVa colon bacteria grown on acetate plus fructose, oligofructose, inulin, or lactate were performed. A gas chromatography method was set up to assess H2 and CO2 production online and to ensure complete coverage of all metabolites produced. Method accuracy was confirmed through the calculation of electron and carbon recoveries. Fermentations with Anaerostipes caccae DSM 14662T, Roseburia faecis DSM 16840T, Roseburia hominis DSM 16839T, and Roseburia intestinalis DSM 14610T revealed similar patterns of metabolite production with butyrate, CO2, and H2 as the main metabolites. R. faecis DSM 16840T and R. intestinalis DSM 14610T were able to degrade oligofructose, displaying a nonpreferential breakdown mechanism. Lactate consumption was only observed with A. caccae DSM 14662T. Roseburia inulinivorans DSM 16841T was the only strain included in the present study that was able to grow on fructose, oligofructose, and inulin. The metabolites produced were lactate, butyrate, and CO2, without H2 production, indicating an energy metabolism distinct from that of other Roseburia species. Oligofructose degradation was nonpreferential. In a coculture of R. inulinivorans DSM 16841T with the highly competitive strain Bifidobacterium longum subsp. longum LMG 11047 on inulin, hardly any production of butyrate and CO2 was detected, indicating a lack of competitiveness of the butyrate producer. Complete recovery of metabolites during fermentations of clostridial cluster XIVa butyrate-producing colon bacteria allowed stoichiometric balancing of the metabolic pathway for butyrate production, including H2 formation.The implementation of 16S rRNA gene-based analytical techniques in the ongoing exploration of the microbial diversity of the human colon ecosystem has both broadened and sharpened the prevailing image of its population (17, 24, 32). While a rather conservative perception of the composition of the colon microbiota has dominated gut research for several decades (36), recent studies have revealed the importance of previously largely neglected bacterial groups and have reduced historically numerically overestimated subpopulations to their actual (marginal) size (8, 22, 52). The human colon has been shown to be a remarkably selective environment, which is reflected by a rather shallow microbial diversity (32). Species belonging to the bacterial divisions Firmicutes, Bacteroidetes, Proteobacteria, and Actinobacteria make up more than 98% of the bacterial population of the human colon (2, 17, 24). However, this superficial uniformity only covers an overwhelming diversity at the lower taxonomic levels; the human colon has been estimated to harbor between 500 and 1,000 species, representing over 7,000 strains, with up to 80% of them considered uncultivable using presently available methodologies (14, 28, 53).Assessing identity and abundance of the major microbial groups composing the colon microbiota is a first and indispensable step toward a better understanding of the ecosystem of the large intestine (48). However, defining a complex ecosystem such as the human colon requires more than the construction of a catalog of its members (32). A major challenge of gastrointestinal microbiology lies in linking phylogenetic subgroups with particular ecological habitats and niches (7, 8, 23). The latter requires further development of highly discriminating 16S rRNA gene-targeted probes to monitor spatial bacterial distribution, combined with renewed efforts toward species isolation through the application of innovative cultivation methods and media, and extensive metabolic characterization of representative strains (19, 35, 48).Recently, a global ecological approach, combining efforts in probe development (1, 27), species isolation (3), and metabolic characterization (4, 11, 15, 20), has led to the identification of a functional group of microorganisms, composed of species belonging to the clostridial clusters IV and XIVa, that are responsible for colon butyrate production. As butyrate is regarded as a key metabolite for the maintenance of colon health, this functional subunit of the colon microbiota could have a major influence on human well-being and might be considered as a target for prebiotic dietary interventions (25, 35, 45). Some recently described lactate- and/or acetate-converting colon butyrate producers have been reported to be able to degrade prebiotic inulin-type fructans, although the kinetics of their respective breakdown mechanisms have hardly been investigated (10, 20). The enhancement of colon butyrate production observed after consumption of oligofructose or inulin (6, 31, 40)—the so-called butyrogenic effect—as well as the limited stimulatory effect of these prebiotics on the clostridial cluster IV and XIVa colon populations (16, 30) have been attributed to cross-feeding with bifidobacteria, which are still considered the primary fructan degraders (5, 38). Anaerostipes caccae as well as Roseburia spp. have been shown to be able to (co)metabolize end products of bifidobacterial fructan fermentation (lactate and/or acetate) or to grow on short oligosaccharides and monosaccharides released by Bifidobacterium spp. during fructan degradation (4, 20).Recently, many clostridial cluster IV and XIVa butyrate producers characterized in detail have been shown to produce gases, mainly CO2 and H2 (12, 15, 20, 46). Consequently, they might be responsible for an enhancement of gas production as a result of fructan fermentation, through either cross-feeding or direct degradation of inulin-type fructans (15, 16). Indeed, inulin-type fructan consumption has been reported to cause some gastrointestinal discomfort related to gas production—essentially, flatulence and bloating (43)—while bifidobacteria, the main beneficiaries of dietary fructan intake, do not produce gases (19, 49). Although CO2 and H2 production by colon butyrate producers could have implications for human intestinal well-being, (in vitro) production has not been satisfactorily monitored up to now, probably due to limited availability of a performant apparatus for (online) gas analysis (15, 20). Moreover, the currently proposed pathway for colon butyrate production does not provide a conclusive quantitative link between bacterial (co)substrate metabolism and H2 formation (11).This study investigated the kinetics of inulin-type fructan degradation by representatives of the genera Anaerostipes and Roseburia. A method based on online gas chromatography (GC) was developed to assess gas production qualitatively and quantitatively in a continuously sparged fermentation vessel for complete coverage of metabolite production. The competitiveness of inulin-degrading butyrate producers was investigated through coculture fermentations with Bifidobacterium longum subsp. longum LMG 11047, a strain representing a highly competitive cluster of bifidobacteria that share both high fructose consumption and oligofructose degradation rates and are able to perform partial breakdown of inulin (18, 20). A stoichiometrically balanced pathway for butyrate production, including H2 production, is proposed.  相似文献   

9.
The molecular complexes involved in the nonhomologous end-joining process that resolves recombination-activating gene (RAG)-induced double-strand breaks and results in V(D)J gene rearrangements vary during mammalian ontogeny. In the mouse, the first immunoglobulin gene rearrangements emerge during midgestation periods, but their repertoires have not been analyzed in detail. We decided to study the postgastrulation DJH joints and compare them with those present in later life. The embryo DJH joints differed from those observed in perinatal life by the presence of short stretches of nontemplated (N) nucleotides. Whereas most adult N nucleotides are introduced by terminal deoxynucleotidyl transferase (TdT), the embryo N nucleotides were due to the activity of the homologous DNA polymerase μ (Polμ), which was widely expressed in the early ontogeny, as shown by analysis of Polμ−/− embryos. Based on its DNA-dependent polymerization ability, which TdT lacks, Polμ also filled in small sequence gaps at the coding ends and contributed to the ligation of highly processed ends, frequently found in the embryo, by pairing to internal microhomology sites. These findings show that Polμ participates in the repair of early-embryo, RAG-induced double-strand breaks and subsequently may contribute to preserve the genomic stability and cellular homeostasis of lymphohematopoietic precursors during development.The adaptive immune system is characterized by the great diversity of its antigen receptors, which result from the activities of enzymatic complexes that cut and paste the genomic DNA of antigen receptor loci. The nonhomologous end-joining (NHEJ) machinery is then recruited to repair the double-strand DNA breaks (DSBs) inflicted by the products of the recombination-activating genes (RAGs) (45, 65). Within B cells, each immunoglobulin (Ig) receptor represents a singular shuffling of two heavy (H) and two light (L) chains, which are derived from the recombination of V, D, and J gene segments of the IgH locus and of V and J for IgL (71). Besides these combinatorial possibilities, most Ig variability derives from extensive processing of the coding ends, including exonucleolytic trimming of DNA ends, together with the addition of palindromic (P) nucleotides templated by the adjacent germ line sequence and of nontemplated (N) nucleotides secondary to the activity of the terminal deoxynucleotidyl transferase (TdT), a lymphoid-specific member of family X of DNA polymerases (reviewed in reference 56). During B-lineage differentiation, IgH rearrangements occur before those of the IgL locus, and D-to-JH rearrangements precede V-to-DJH rearrangements (62). DJH joints are formed in any of the three open reading frames (ORFs). ORF1 is predominantly used in mature Igs, ORF2 is transcribed as a Dμ protein that provides negative signals to the B-cell precursors, and ORF3 frequently leads to stop codons (32, 33, 37). Germ line V, D, and J gene segments display short stretches of mutually homologous nucleotides (SSH), which are frequently used in gene rearrangements during perinatal periods, when N additions are absent (27, 32, 55, 57). The actual Ig V-region repertoires represent both the results of the NHEJ process associated with genomic VDJ recombination and those of antigen-independent and -dependent selection events. Although the core NHEJ components (Ku-Artemis-DNA-PK and XLF-XRCC4-DNA ligase IV) are by themselves able to join RAG-induced, incompatible DNA ends, family X DNA polymerases can be recruited to fill gaps created by imprecise coding ends with 3′ overhangs (DNA polymerase μ [Polμ] and Polλ) and/or to promote diversity through the addition of N nucleotides (TdT) (34, 56).The lymphoid differentiation pathways and clonotypic repertoires are developmentally regulated and differ between the embryo-fetal and adult periods (2, 44, 68). The perinatal B cells result from a wave of B lymphopoiesis occurring during the last third of mouse gestation (13, 14, 21, 70). Perinatal VH gene usage differs from that predominating in the adult (1, 69), and the former VDJ joints rarely display N additions, leading to V-region repertoires enriched in multi- and self-reactive specificities (36, 40). The program of B-cell differentiation starts at embryonic days 10 to 11 (E10 to E11) in embryo hematopoietic sites, after the emergence of multipotent progenitors (at E8.5 to E9.5) (18, 19, 23, 31, 51, 73). DJH rearrangements were detected in these early embryos, whereas full VDJH sequences were not observed before E14 (14, 18, 51, 66), when VJκ rearrangements were also found (63). The earliest mouse DJH/VDJH Ig sequences analyzed to date corresponded to late fetuses (E16) (14, 53). We reasoned that the true baseline of the Ig rearrangement process occurs in midgestation embryos, when the first DJHs are not yet transcribed and, consequently, not subjected to selection and are conditioned only for the evolutionarily established and developmentally regulated usage of distinct NHEJ machineries.We report here the sequence profiles of the earliest embryo E10 to E12 DJH joints. Unexpected frequencies of embryonic DJH joints bearing N nucleotides, in the absence of detectable TdT expression, were found. Moreover, the embryo DJH joints lacking N nucleotides (N) used fewer SSH to recombine than newborn DJHs, and these SSH were widely dispersed along the embryo D sequences, in contrast to the most joint-proximal ones, which predominated in newborn DJHs. Considering that Polμ is the closest relative of TdT (42% amino acid identity) (22), which is able to introduce N nucleotides in vitro (4, 22, 34, 39, 49) and to join DNA ends with minimal or even null complementarity (17, 58), and that it is expressed in early-embryo organs, we decided to investigate its putative contribution to the first embryo DJH joints. The DJH joints obtained from Polμ−/− embryos (48) showed a significant reduction of N nucleotides compared to wild-type (WT) embryos. Moreover, highly preserved DJH joints (with <3 deleted nucleotides) were selectively depleted in the Polμ−/− mouse embryos, while the remaining DJHs preferentially relied upon longer stretches of homology for end ligation. These findings support the idea that Polμ is active during early-embryo DJH rearrangements and that both its template-dependent and -independent ambivalent functions may be used to fill in small nucleotide gaps generated after asymmetric hairpin nicking and also to extend coding ends via a limited TdT-like activity.  相似文献   

10.
11.
12.
13.
We examined the diversity of a marker gene for homoacetogens in two cockroach gut microbial communities. Formyltetrahydrofolate synthetase (FTHFS or fhs) libraries prepared from a wood-feeding cockroach, Cryptocercus punctulatus, were dominated by sequences that affiliated with termite gut treponemes. No spirochete-like sequences were recovered from the omnivorous roach Periplaneta americana, which was dominated by Firmicutes-like sequences.The guts of wood-feeding termites and Cryptocercus punctulatus cockroaches share an unusual pattern of electron flow, as high rates of CO2-reductive acetogenesis typically supplant methanogenesis as the terminal electron sink (2, 3). Past studies have shown that from 10 to 30% of gut acetate produced in environments of termites and wood-feeding cockroaches is microbially generated from CO2 (3, 28), ultimately powering 18 to 26% of the host insect''s own respiratory energy metabolism (25). Nevertheless, most termites emit methane (2), and termite emissions constitute approximately 4% of the global methane budget (27). Cockroaches have been proposed to represent an additional source of note (9). Interestingly, methanogenic termites and cockroaches exhibit increased acetogenesis following addition of exogenous H2 (3, 29). This suggests that these insects are host to a robust population of bacteria that are capable of homoacetogenesis but may be primarily using alternative electron donors (and other substrates and pathways) in vivo.Acetogenic bacteria belonging to two bacterial phyla, Firmicutes and Spirochaetes, have been isolated from the guts of termites (1, 4, 11, 12, 14). Several surveys have targeted and used the gene for formyltetrahydrofolate synthetase (FTHFS), a key gene in the Wood-Ljungdahl pathway of acetogenesis (16), as a potential marker for the pathway (15, 18). For the wood-feeding termites that have been examined, the studies have revealed an abundance of FTHFS sequences that form a coherent phylogenetic cluster, together with genes from homoacetogenic termite gut spirochetes belonging to the genus Treponema (24, 26, 30). This suggests that treponemes may be among the more abundant of the homoacetogens active in these environments.Little is known about the population structure and biology of CO2-reducing, acetogenic bacteria in the guts of either omnivorous or wood-feeding cockroaches. The wood-feeding cockroach Cryptocercus hosts an abundance of flagellate protozoa closely related to those believed to dominate polysaccharide fermentation in the guts of termites (5, 6, 22), suggesting that at least one key environmental niche is filled by similar microbes in both termites and Cryptocercidae. Additionally, Cryptocercidae cockroaches, like termites, house diverse spirochetes and are the site of intense CO2 reduction into acetate (3, 7). Possibly, spirochetes capable of CO2 reduction into acetate are present in the hindguts of cockroaches. However, no evidence has yet been presented for the existence of homoacetogenic treponemes in environments other than the guts of termites, and FTHFS surveys of human (21) or horse (15) fecal matter and bovine rumen samples (20) revealed only Firmicutes-like and other FTHFS alleles that are distinct from those comprising the termite treponeme cluster.Here, by examining FTHFS gene diversity in Cryptocercus punctulatus and Periplaneta americana guts, we endeavored to learn more about the distribution and origins of homoacetogenic treponemes (and their genes) that are found in wood-feeding termites. In particular, we wished to ascertain whether FTHFS genes present in either of the two cockroaches are termite treponeme-like and, if so, whether analysis reveals any obvious signal indicating recent or ancient lateral community transfer events between insect lineages.  相似文献   

14.
15.
Homoacetogens produce acetate from H2 and CO2 via the Wood-Ljungdahl pathway. Some homoacetogens have been isolated from the rumen, but these organisms are expected to be only part of the full diversity present. To survey the presence of rumen homoacetogens, we analyzed sequences of formyltetrahydrofolate synthetase (FTHFS), a key enzyme of the Wood-Ljungdahl pathway. A total of 275 partial sequences of genes encoding FTHFS were PCR amplified from rumen contents of a cow, two sheep, and a deer. Phylogenetic trees were constructed using these FTHFS gene sequences and the translated amino acid sequences, together with other sequences from public databases and from novel nonhomoacetogenic bacteria isolated from the rumen. Over 90% of the FTHFS sequences fell into 34 clusters defined with good bootstrap support. Few rumen-derived FTHFS sequences clustered with sequences of known homoacetogens. Conserved residues were identified in the deduced FTHFS amino acid sequences from known homoacetogens, and their presence in the other sequences was used to determine a “homoacetogen similarity” (HS) score. A homoacetogen FTHFS profile hidden Markov model (HoF-HMM) was used to assess the homology of rumen and homoacetogen FTHFS sequences. Many clusters had low HS scores and HoF-HMM matches, raising doubts about whether the sequences originated from homoacetogens. In keeping with these findings, FTHFS sequences from nonhomoacetogenic bacterial isolates grouped in these clusters with low scores. However, sequences that formed 10 clusters containing no known isolates but representing 15% of our FTHFS sequences from rumen samples had high HS scores and HoF-HMM matches and so could represent novel homoacetogens.Feed ingested by ruminant animals is fermented in the rumen by a complex community of microbes. This community produces, among other products, the volatile fatty acids acetate, propionate, and butyrate, which are absorbed across the rumen wall and satisfy a large part of the animals'' carbon and energy requirements. Hydrogen gas (H2) is also formed and is the major precursor of the methane (CH4) formed in ruminant animals. This ruminant-derived CH4 is a contributor to global greenhouse gas emissions (46) and also represents an energy loss for the animals (34). Proposed ruminant greenhouse gas mitigation strategies include using feeds that produce less CH4 and more volatile fatty acids (31). Alternative strategies include interventions that slow or halt methanogenesis by vaccination, using natural inhibitors found in plants, and supplementing feed with fats and oils or small-molecule inhibitors (31, 32). In the absence of methanogenesis, accumulation of H2 could lead to a decrease in the rate of feed fermentation (31, 53) and hence a decrease in animal productivity. Other microbes that use H2 without producing methane could be valuable in conjunction with intervention strategies that inhibit methanogens. This possibility has sparked interest in possible inoculation of ruminants with alternative H2 users.Bacteria that use the Wood-Ljungdahl pathway to produce acetate from CO2 are metabolically (6) and phylogenetically (48) diverse and are designated “homoacetogens.” Homoacetogens grow with H2 or other suitable electron donors, such as formate or sugars, plus CO2 as a terminal electron acceptor, heterotrophically with organic substrates such as sugars and methoxylated compounds, or mixotrophically with, e.g., H2 and organic substrates. Homoacetogens have been reported to occur in a normally functioning rumen, but they are unlikely to compete with methanogens for H2 (24, 25, 34). However, homoacetogens could play an important role in the disposal of H2 if methanogens are not established in or are eliminated from the rumen (11, 17). At present, it is not clear whether resident rumen homoacetogens could fulfill the H2 disposal role or whether homoacetogens would have to be added to the rumen to take over this role from the methanogens.Cultivation-based enumeration techniques have shown that the sizes of rumen acetogen populations range from undetectable to 1.2 × 109 per g of rumen contents and that the prevalence of these acetogens depends on diet, animal age, and time of sampling (5, 7, 23, 24). Several homoacetogens, including Acetitomaculum ruminis (15), Eubacterium limosum (14, 17), Blautia schinkii, and Blautia producta (11), have been isolated from ruminants. Homoacetogens have also been isolated from the kangaroo forestomach, whose function is analogous to that of the rumen, which suggests that homoacetogenesis may play a role in hydrogen removal in the low-methane-emission forestomach (37).Because homoacetogens occur in different lineages of bacteria (48), traditional 16S rRNA gene-based surveys provide little information on their prevalence. The formyltetrahydrofolate synthetase (FTHFS) gene (fhs) has been used as a functional marker for homoacetogens, as the enzyme that it encodes catalyzes a key step in the reductive acetogenesis pathway (26). The structure of the enzyme of the homoacetogen Moorella thermoacetica has been reported, and putative functional features have been identified (27, 41, 42). FTHFS sequences from true homoacetogens differ from their homologs in sulfate-reducing bacteria and in other bacteria that degrade purines and amino acids via the glycine synthase-glycine reductase pathway (12, 21, 22, 26). At present, only a limited number of FTHFS sequences have been deposited in databases, and the vast majority of them are partial sequences retrieved from complex microbial communities. FTHFS sequences have been surveyed in sludge (39, 43, 54), termites (40, 44), salt marsh plant roots (21), horse manure (22), cow manure, freshwater sediment, rice field soil, and sewage (54), but so far only one study has investigated bovine ruminal FTHFS sequences (30). The rumen FTHFS sequences had low levels of similarity to the FTHFS sequences of known homoacetogens and could be sequences of novel homoacetogens. To our knowledge, no bacteria with these unique FTHFS sequences have been identified.The aims of this study were to assess the diversity of FTHFS gene sequences retrieved from rumen samples and to screen novel rumen isolates for the presence of FTHFS genes and test their ability to grow as homoacetogens. We used alignments of FTHFS sequences to define a homoacetogen similarity score based on the presence of diagnostic amino acids and developed a hidden Markov model to assess the likelihood that FTHFS sequences of unknown origin are sequences from true homoacetogens that are able to use H2 or alternative electron donors for reductive acetogenesis.  相似文献   

16.
An intracellular multiplication F (IcmF) family protein is a conserved component of a newly identified type VI secretion system (T6SS) encoded in many animal and plant-associated Proteobacteria. We have previously identified ImpLM, an IcmF family protein that is required for the secretion of the T6SS substrate hemolysin-coregulated protein (Hcp) from the plant-pathogenic bacterium Agrobacterium tumefaciens. In this study, we characterized the topology of ImpLM and the importance of its nucleotide-binding Walker A motif involved in Hcp secretion from A. tumefaciens. A combination of β-lactamase-green fluorescent protein fusion and biochemical fractionation analyses revealed that ImpLM is an integral polytopic inner membrane protein comprising three transmembrane domains bordered by an N-terminal domain facing the cytoplasm and a C-terminal domain exposed to the periplasm. impLM mutants with substitutions or deletions in the Walker A motif failed to complement the impLM deletion mutant for Hcp secretion, which provided evidence that ImpLM may bind and/or hydrolyze nucleoside triphosphates to mediate T6SS machine assembly and/or substrate secretion. Protein-protein interaction and protein stability analyses indicated that there is a physical interaction between ImpLM and another essential T6SS component, ImpKL. Topology and biochemical fractionation analyses suggested that ImpKL is an integral bitopic inner membrane protein with an N-terminal domain facing the cytoplasm and a C-terminal OmpA-like domain exposed to the periplasm. Further comprehensive yeast two-hybrid assays dissecting ImpLM-ImpKL interaction domains suggested that ImpLM interacts with ImpKL via the N-terminal cytoplasmic domains of the proteins. In conclusion, ImpLM interacts with ImpKL, and its Walker A motif is required for its function in mediation of Hcp secretion from A. tumefaciens.Many pathogenic gram-negative bacteria employ protein secretion systems formed by macromolecular complexes to deliver proteins or protein-DNA complexes across the bacterial membrane. In addition to the general secretory (Sec) pathway (18, 52) and twin-arginine translocation (Tat) pathway (7, 34), which transport proteins across the inner membrane into the periplasm, at least six distinct protein secretion systems occur in gram-negative bacteria (28, 46, 66). These systems are able to secrete proteins from the cytoplasm or periplasm to the external environment or the host cell and include the well-documented type I to type V secretion systems (T1SS to T5SS) (10, 15, 23, 26, 30) and a recently discovered type VI secretion system (T6SS) (4, 8, 22, 41, 48, 49). These systems use ATPase or a proton motive force to energize assembly of the protein secretion machinery and/or substrate translocation (2, 6, 41, 44, 60).Agrobacterium tumefaciens is a soilborne pathogenic gram-negative bacterium that causes crown gall disease in a wide range of plants. Using an archetypal T4SS (9), A. tumefaciens translocates oncogenic transferred DNA and effector proteins to the host and ultimately integrates transferred DNA into the host genome. Because of its unique interkingdom DNA transfer, this bacterium has been extensively studied and used to transform foreign DNA into plants and fungi (11, 24, 40, 67). In addition to the T4SS, A. tumefaciens encodes several other secretion systems, including the Sec pathway, the Tat pathway, T1SS, T5SS, and the recently identified T6SS (72). T6SS is highly conserved and widely distributed in animal- and plant-associated Proteobacteria and plays an important role in the virulence of several human and animal pathogens (14, 19, 41, 48, 56, 63, 74). However, T6SS seems to play only a minor role or even a negative role in infection or virulence of the plant-associated pathogens or symbionts studied to date (5, 37-39, 72).T6SS was initially designated IAHP (IcmF-associated homologous protein) clusters (13). Before T6SS was documented by Pukatzki et al. in Vibrio cholerae (48), mutations in this gene cluster in the plant symbiont Rhizobium leguminosarum (5) and the fish pathogen Edwardsiella tarda (51) caused defects in protein secretion. In V. cholerae, T6SS was responsible for the loss of cytotoxicity for amoebae and for secretion of two proteins lacking a signal peptide, hemolysin-coregulated protein (Hcp) and valine-glycine repeat protein (VgrG). Secretion of Hcp is the hallmark of T6SS. Interestingly, mutation of hcp blocks the secretion of VgrG proteins (VgrG-1, VgrG-2, and VgrG-3), and, conversely, vgrG-1 and vgrG-2 are both required for secretion of the Hcp and VgrG proteins from V. cholerae (47, 48). Similarly, a requirement of Hcp for VgrG secretion and a requirement of VgrG for Hcp secretion have also been shown for E. tarda (74). Because Hcp forms a hexameric ring (41) stacked in a tube-like structure in vitro (3, 35) and VgrG has a predicted trimeric phage tail spike-like structure similar to that of the T4 phage gp5-gp27 complex (47), Hcp and VgrG have been postulated to form an extracellular translocon. This model is further supported by two recent crystallography studies showing that Hcp, VgrG, and a T4 phage gp25-like protein resembled membrane penetration tails of bacteriophages (35, 45).Little is known about the topology and structure of T6SS machinery subunits and the distinction between genes encoding machinery subunits and genes encoding regulatory proteins. Posttranslational regulation via the phosphorylation of Fha1 by a serine-threonine kinase (PpkA) is required for Hcp secretion from Pseudomonas aeruginosa (42). Genetic evidence for P. aeruginosa suggested that the T6SS may utilize a ClpV-like AAA+ ATPase to provide the energy for machinery assembly or substrate translocation (41). A recent study of V. cholerae suggested that ClpV ATPase activity is responsible for remodeling the VipA/VipB tubules which are crucial for type VI substrate secretion (6). An outer membrane lipoprotein, SciN, is an essential T6SS component for mediating Hcp secretion from enteroaggregative Escherichia coli (1). A systematic study of the T6SS machinery in E. tarda revealed that 13 of 16 genes in the evp gene cluster are essential for secretion of T6S substrates (74), which suggests the core components of the T6SS. Interestingly, most of the core components conserved in T6SS are predicted soluble proteins without recognizable signal peptide and transmembrane (TM) domains.The intracellular multiplication F (IcmF) and H (IcmH) proteins are among the few core components with obvious TM domains (8). In Legionella pneumophila Dot/Icm T4SSb, IcmF and IcmH are both membrane localized and partially required for L. pneumophila replication in macrophages (58, 70, 75). IcmF and IcmH are thought to interact with each other in stabilizing the T4SS complex in L. pneumophila (58). In T6SS, IcmF is one of the essential components required for secretion of Hcp from several animal pathogens, including V. cholerae (48), Aeromonas hydrophila (63), E. tarda (74), and P. aeruginosa (41), as well as the plant pathogens A. tumefaciens (72) and Pectobacterium atrosepticum (39). In E. tarda, IcmF (EvpO) interacted with IcmH (EvpN), EvpL, and EvpA in a yeast two-hybrid assay, and its putative nucleotide-binding site (Walker A motif) was not essential for secretion of T6SS substrates (74).In this study, we characterized the topology and interactions of the IcmF and IcmH family proteins ImpLM and ImpKL, which are two essential components of the T6SS of A. tumefaciens. We adapted the nomenclature proposed by Cascales (8), using the annotated gene designation followed by the letter indicated by Shalom et al. (59). Our data indicate that ImpLM and ImpKL are both integral inner membrane proteins and interact with each other via their N-terminal domains residing in the cytoplasm. We also provide genetic evidence showing that ImpLM may function as a nucleoside triphosphate (NTP)-binding protein or nucleoside triphosphatase to mediate T6S machinery assembly and/or substrate secretion.  相似文献   

17.
Nitrate-reducing enrichments, amended with n-hexadecane, were established with petroleum-contaminated sediment from Onondaga Lake. Cultures were serially diluted to yield a sediment-free consortium. Clone libraries and denaturing gradient gel electrophoresis analysis of 16S rRNA gene community PCR products indicated the presence of uncultured alpha- and betaproteobacteria similar to those detected in contaminated, denitrifying environments. Cultures were incubated with H34-hexadecane, fully deuterated hexadecane (d34-hexadecane), or H34-hexadecane and NaH13CO3. Gas chromatography-mass spectrometry analysis of silylated metabolites resulted in the identification of [H29]pentadecanoic acid, [H25]tridecanoic acid, [1-13C]pentadecanoic acid, [3-13C]heptadecanoic acid, [3-13C]10-methylheptadecanoic acid, and d27-pentadecanoic, d25-, and d24-tridecanoic acids. The identification of these metabolites suggests a carbon addition at the C-3 position of hexadecane, with subsequent β-oxidation and transformation reactions (chain elongation and C-10 methylation) that predominantly produce fatty acids with odd numbers of carbons. Mineralization of [1-14C]hexadecane was demonstrated based on the recovery of 14CO2 in active cultures.Linear alkanes account for a large component of crude and refined petroleum products and, therefore, are of environmental significance with respect to their fate and transport (38). The aerobic activation of alkanes is well documented and involves monooxygenase and dioxygenase enzymes in which not only is oxygen required as an electron acceptor but it also serves as a reactant in hydroxylation (2, 16, 17, 32, 34). Alkanes are also degraded under anoxic conditions via novel degradation strategies (34). To date, there are two known pathways of anaerobic n-alkane degradation: (i) alkane addition to fumarate, commonly referred to as fumarate addition, and (ii) a putative pathway, proposed by So et al. (25), involving carboxylation of the alkane. Fumarate addition proceeds via terminal or subterminal addition (C-2 position) of the alkane to the double bond of fumarate, resulting in the formation of an alkylsuccinate. The alkylsuccinate is further degraded via carbon skeleton rearrangement and β-oxidation (4, 6, 8, 12, 13, 21, 37). Alkane addition to fumarate has been documented for a denitrifying isolate (21, 37), sulfate-reducing consortia (4, 8, 12, 13), and five sulfate-reducing isolates (4, 6-8, 12). In addition to being demonstrated in these studies, fumarate addition in a sulfate-reducing enrichment growing on the alicyclic alkane 2-ethylcyclopentane has also been demonstrated (23). In contrast to fumarate addition, which has been shown for both sulfate-reducers and denitrifiers, the putative carboxylation of n-alkanes has been proposed only for the sulfate-reducing isolate strain Hxd3 (25) and for a sulfate-reducing consortium (4). Experiments using NaH13CO3 demonstrated that bicarbonate serves as the source of inorganic carbon for the putative carboxylation reaction (25). Subterminal carboxylation of the alkane at the C-3 position is followed by elimination of the two terminal carbons, to yield a fatty acid that is one carbon shorter than the parent alkane (4, 25). The fatty acids are subject to β-oxidation, chain elongation, and/or C-10 methylation (25).In this study, we characterized an alkane-degrading, nitrate-reducing consortium and surveyed the metabolites of the consortium incubated with either unlabeled or labeled hexadecane in order to elucidate the pathway of n-alkane degradation. We present evidence of a pathway analogous to the proposed carboxylation pathway under nitrate-reducing conditions.  相似文献   

18.
19.
2G12 is a broadly neutralizing anti-HIV-1 monoclonal human IgG1 antibody reactive with a high-mannose glycan cluster on the surface of glycoprotein gp120. A key feature of this very highly mutated antibody is domain exchange of the heavy-chain variable region (VH) with the VH of the adjacent Fab of the same immunoglobulin, which assembles a multivalent binding interface composed of two primary binding sites in close proximity. A non-germ line-encoded proline in the elbow between VH and CH1 and an extensive network of hydrophobic interactions in the VH/VH′ interface have been proposed to be crucial for domain exchange. To investigate the origins of domain exchange, a germ line version of 2G12 that behaves as a conventional antibody was engineered. Substitution of 5 to 7 residues for those of the wild type produced a significant fraction of domain-exchanged molecules, with no evidence of equilibrium between domain-exchanged and conventional forms. Two substitutions not previously implicated, AH14 and EH75, are the most crucial for domain exchange, together with IH19 at the VH/VH′ interface and PH113 in the elbow region. Structural modeling gave clues as to why these residues are essential for domain exchange. The demonstration that domain exchange can be initiated by a small number of substitutions in a germ line antibody suggests that the evolution of a domain-exchanged antibody response in vivo may be more readily achieved than considered to date.Protein oligomers are able to exchange or swap an element of their secondary structure or an entire protein domain. The functional unit in domain-exchanged proteins thereby stays preserved, as only the linking hinge loop changes conformation significantly (4, 17, 27). Analogous to other domain-swapped proteins, antibodies can exchange an entire domain, in this case the heavy-chain variable region (VH), with an equivalent heavy-chain variable region of an adjacent Fab (VH′) within the same immunoglobulin (Ig) molecule (11). The advantages of domain-exchanged proteins, including antibodies, are higher local concentrations of active sites, a larger binding surface, and a potential secondary active site at the new subunit interface (27, 45). The one and only antibody shown to be domain exchanged to date is 2G12 (7, 11), but this arrangement is potentially possible for any Ig and could have been overlooked at least in some instances.2G12 is one of only a few high-affinity monoclonal antibodies with broad neutralizing activity against different subtypes of HIV-1 (5, 30, 40, 43). The antibody binds a dense cluster of N-linked high-mannose glycans (Man8-9GlcNAc2) on the envelope surface glycoprotein gp120 (10, 35, 36, 41). The domain-exchanged arrangement forms a multivalent binding site composed of two primary binding sites in close proximity and a proposed secondary binding site formed by the novel VH/VH′ interface (11). 2G12 provides protection against infection in animal models (19, 31) and has been shown to induce neutralization escape following passive immunization in humans (39).Consensus has grown that a successful HIV-1 vaccine will need to include a component that elicits broadly neutralizing antibodies (8, 18, 21, 26, 32, 42). All attempts to elicit 2G12-like antibodies with the desired specificity and neutralization activity have failed to date (22, 29, 44), conceivably due to difficulties in generating adequate mimicry of the glycan cluster and tolerance mechanisms or, very likely, the inability to induce domain exchange (1). Unraveling the mechanism of domain exchange and how this conformation might have evolved is highly desirable to direct future HIV-1 vaccine design to elicit 2G12-like antibodies.By comparison with other domain-exchanged proteins (27), the following three mechanisms have been proposed to contribute to the unique structure of 2G12 compared to the structure of a conventional antibody: destabilization of the “closed” VH/VL interface, conformational change in the elbow between VH and CH1, and an energetically favorable “open” VH/VH′ interface (11). Key residues involved in promoting domain exchange were predicted based on examination of interacting residues at the two interfaces and by the effects of alanine substitutions on the binding of wild-type 2G12 to gp120. However, the importance of these key residues for domain exchange was not directly demonstrated experimentally (11).Here, we explored the minimal requirements for domain exchange of 2G12, starting with a germ line version of the antibody that adopts a conventional antibody structure. Although wild-type 2G12 is heavily somatically mutated, only five to seven substitutions in the germ line version of the antibody were shown to produce a significant fraction of domain-exchanged molecules. The results suggest the evolution of domain-exchanged antibody responses may be more facile than considered to date.  相似文献   

20.
Halogenases have been shown to play a significant role in biosynthesis and introducing the bioactivity of many halogenated secondary metabolites. In this study, 54 reduced flavin adenine dinucleotide (FADH2)-dependent halogenase gene-positive strains were identified after the PCR screening of a large collection of 228 reference strains encompassing all major families and genera of filamentous actinomycetes. The wide distribution of this gene was observed to extend to some rare lineages with higher occurrences and large sequence diversity. Subsequent phylogenetic analyses revealed that strains containing highly homologous halogenases tended to produce halometabolites with similar structures, and halogenase genes are likely to propagate by horizontal gene transfer as well as vertical inheritance within actinomycetes. Higher percentages of halogenase gene-positive strains than those of halogenase gene-negative ones contained polyketide synthase genes and/or nonribosomal peptide synthetase genes or displayed antimicrobial activities in the tests applied, indicating their genetic and physiological potentials for producing secondary metabolites. The robustness of this halogenase gene screening strategy for the discovery of particular biosynthetic gene clusters in rare actinomycetes besides streptomycetes was further supported by genome-walking analysis. The described distribution and phylogenetic implications of the FADH2-dependent halogenase gene present a guide for strain selection in the search for novel organohalogen compounds from actinomycetes.It is well known that actinomycetes, notably filamentous actinomycetes, have a remarkable capacity to produce bioactive molecules for drug development (4, 6). However, novel technologies are demanded for the discovery of new bioactive secondary metabolites from these microbes to meet the urgent medical need for drug candidates (5, 9, 31).Genome mining recently has been used to search for new drug leads (7, 20, 42, 51). Based on the hypothesis that secondary metabolites with similar structures are biosynthesized by gene clusters that harbor certain homologous genes, such homologous genes could serve as suitable markers for distinct natural-product gene clusters (26, 51). A wide range of structurally diverse bioactive compounds are synthesized by polyketide synthase (PKS) and nonribosomal peptide synthetase (NRPS) systems in actinomycetes, therefore much attention has been given to revealing a previously unrecognized biosynthetic potential of actinomycetes through the genome mining of these genes (2, 3, 22). However, the broad distribution of PKS and NRPS genes and their high numbers even in a single actinomycete complicate their use (2, 3). To rationally exploit the genetic potential of actinomycetes, more and more special genes, such as tailoring enzyme genes, are being utilized for this sequence-guided genetic screening strategy (20, 38).Tailoring enzymes, which are responsible for the introduction and generation of diversity and bioactivity in several structural classes during or after NRPS, PKS, or NRPS/PKS assembly lines, usually include acyltransferases, aminotransferases, cyclases, glycosyltransferases, halogenases, ketoreductases, methyltransferases, and oxygenases (36, 45). Halogenation, an important feature for the bioactivity of a large number of distinct natural products (16, 18, 30), frequently is introduced by one type of halogenase, called reduced flavin adenine dinucleotide (FADH2)-dependent (or flavin-dependent) halogenase (10, 12, 35). More than 4,000 halometabolites have been discovered (15), including commercially important antibiotics such as chloramphenicol, vancomycin, and teicoplanin (43).Previous investigations of FADH2-dependent halogenase genes were focused largely on related gene clusters in the genera Amycolatopsis (33, 44, 53) and Streptomyces (8, 10, 21, 27, 32, 34, 47-49) and also on those in the genera Actinoplanes (25), Actinosynnema (50), Micromonospora (1), and Nonomuraea (39); however, none of these studies has led to the rest of the major families and genera of actinomycetes. In addition, there is evidence that FADH2-dependent halogenase genes of streptomycetes usually exist in halometabolite biosynthetic gene clusters (20), but we lack knowledge of such genes and clusters in other actinomycetes.In the present study, we show that the distribution of the FADH2-dependent halogenase gene in filamentous actinomycetes does indeed correlate with the potential for halometabolite production based on other genetic or physiological factors. We also showed that genome walking near the halogenase gene locus could be employed to identify closely linked gene clusters that likely encode pathways for organohalogen compound production in actinomycetes other than streptomycetes.  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