首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
3.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

4.
5.
Extracellular DNA (eDNA) is produced by several bacterial species and appears to contribute to biofilm development and cell-cell adhesion. We present data showing that the oral commensals Streptococcus sanguinis and Streptococcus gordonii release DNA in a process induced by pyruvate oxidase-dependent production of hydrogen peroxide (H2O2). Surprisingly, S. sanguinis and S. gordonii cell integrity appears unaffected by conditions that cause autolysis in other eDNA-producing bacteria. Exogenous H2O2 causes release of DNA from S. sanguinis and S. gordonii but does not result in obvious lysis of cells. Under DNA-releasing conditions, cell walls appear functionally intact and ribosomes are retained over time. During DNA release, intracellular RNA and ATP are not coreleased. Hence, the release mechanism appears to be highly specific for DNA. Release of DNA without detectable autolysis is suggested to be an adaptation to the competitive oral biofilm environment, where autolysis could create open spaces for competitors to invade. Since eDNA promotes cell-to-cell adhesion, release appears to support oral biofilm formation and facilitates exchange of genetic material among competent strains.The release of bacterial DNA into the environment is of recent interest since this polymer is now recognized to stabilize cell-to-cell adherence and biofilm architecture (1, 35, 37). Treatment of extracellular DNA (eDNA) with DNase results in reduced intercellular stickiness, consistent with an adhesive function for eDNA. Furthermore, eDNA from Neisseria meningitis appears to have sufficient structural integrity to transform competent strains (11), indicating chromosomal origin. Since the abundance of eDNA is influenced by growth conditions, DNA release can also be regulated (40).DNA release is typically a consequence of cell lysis. Linked to DNA release, genetic transformation is the natural ability of competent bacterial species to take up DNA from the environment (13, 34, 42). During competence development, Streptococcus pneumoniae DNA is released by lysis of a subpopulation of cells (30, 42). Cell lysis and DNA release are controlled in a cell density-dependent signal transduction process. The S. pneumoniae comX regulon, carrying late competence genes, also includes the murein hydrolase genes lytA and cbpD (19, 42). Murein hydrolases digest structural components of the peptidoglycan, contributing to remodeling, recycling, and daughter cell separation. Furthermore, murein hydrolases trigger autolytic cell wall digestion, leading to release of DNA and other cellular content into the environment (36). The autolysis of bacterial cells as part of a regulated death program seems to be an important source for eDNA in diverse species, including Staphylococcus aureus (4, 36, 37), Staphylococcus epidermidis (35), Enterococcus faecalis (44), and Pseudomonas aeruginosa (1). In these species, the eDNA contributes to biofilm formation as a component of the extracellular biofilm matrix (35, 37, 44).Unlike for cell lysis-dependent release, the oral streptococci appear to induce eDNA release by a novel mechanism. In dual-species cultures, the oral commensals Streptococcus sanguinis and Streptococcus gordonii release eDNA in a manner dependent on pyruvate oxidase (Pox) generation of hydrogen peroxide (H2O2) under the control of ambient oxygen (23). In this report, we now provide direct evidence of selective H2O2-induced eDNA release by these oral commensal streptococci.  相似文献   

6.
7.
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   

8.
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   

9.
10.
Hantaviruses infect endothelial cells and cause 2 vascular permeability-based diseases. Pathogenic hantaviruses enhance the permeability of endothelial cells in response to vascular endothelial growth factor (VEGF). However, the mechanism by which hantaviruses hyperpermeabilize endothelial cells has not been defined. The paracellular permeability of endothelial cells is uniquely determined by the homophilic assembly of vascular endothelial cadherin (VE-cadherin) within adherens junctions, which is regulated by VEGF receptor-2 (VEGFR2) responses. Here, we investigated VEGFR2 phosphorylation and the internalization of VE-cadherin within endothelial cells infected by pathogenic Andes virus (ANDV) and Hantaan virus (HTNV) and nonpathogenic Tula virus (TULV) hantaviruses. We found that VEGF addition to ANDV- and HTNV-infected endothelial cells results in the hyperphosphorylation of VEGFR2, while TULV infection failed to increase VEGFR2 phosphorylation. Concomitant with the VEGFR2 hyperphosphorylation, VE-cadherin was internalized to intracellular vesicles within ANDV- or HTNV-, but not TULV-, infected endothelial cells. Addition of angiopoietin-1 (Ang-1) or sphingosine-1-phosphate (S1P) to ANDV- or HTNV-infected cells blocked VE-cadherin internalization in response to VEGF. These findings are consistent with the ability of Ang-1 and S1P to inhibit hantavirus-induced endothelial cell permeability. Our results suggest that pathogenic hantaviruses disrupt fluid barrier properties of endothelial cell adherens junctions by enhancing VEGFR2-VE-cadherin pathway responses which increase paracellular permeability. These results provide a pathway-specific mechanism for the enhanced permeability of hantavirus-infected endothelial cells and suggest that stabilizing VE-cadherin within adherens junctions is a primary target for regulating endothelial cell permeability during pathogenic hantavirus infection.Hantaviruses cause 2 human diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (50). HPS and HFRS are multifactorial in nature and cause thrombocytopenia, immune and endothelial cell responses, and hypoxia, which contribute to disease (7, 11, 31, 42, 62). Although these syndromes sound quite different, they share common components which involve the ability of hantaviruses to infect endothelial cells and induce capillary permeability. Edema, which results from capillary leakage of fluid into tissues and organs, is a common finding in both HPS and HFRS patients (4, 7, 11, 31, 42, 62). In fact, both diseases can present with renal or pulmonary sequelae, and the renal or pulmonary focus of hantavirus diseases is likely to result from hantavirus infection of endothelial cells within vast glomerular and pulmonary capillary beds (4, 7, 11, 31, 42, 62). All hantaviruses predominantly infect endothelial cells which line capillaries (31, 42, 44, 61, 62), and endothelial cells have a primary role in maintaining fluid barrier functions of the vasculature (1, 12, 55). Although hantaviruses do not lyse endothelial cells (44, 61), this primary cellular target underlies hantavirus-induced changes in capillary integrity. As a result, understanding altered endothelial cell responses following hantavirus infection is fundamental to defining the mechanism of permeability induced by pathogenic hantaviruses (1, 12, 55).Pathogenic, but not nonpathogenic, hantaviruses use β3 integrins on the surface of endothelial cells and platelets for attachment (19, 21, 23, 39, 46), and β3 integrins play prominent roles in regulating vascular integrity (3, 6, 8, 24, 48). Pathogenic hantaviruses bind to basal, inactive conformations of β3 integrins (35, 46, 53) and days after infection inhibit β3 integrin-directed endothelial cell migration (20, 46). This may be the result of cell-associated virus (19, 20, 22) which keeps β3 in an inactive state but could also occur through additional regulatory processes that have yet to be defined. Interestingly, the nonpathogenic hantaviruses Prospect Hill virus (PHV) and Tula virus (TULV) fail to alter β3 integrin functions, and their entry is consistent with the use of discrete α5β1 integrins (21, 23, 36).On endothelial cells, αvβ3 integrins normally regulate permeabilizing effects of vascular endothelial growth factor receptor-2 (VEGFR2) (3, 24, 48, 51). VEGF was initially identified as an edema-causing vascular permeability factor (VPF) that is 50,000 times more potent than histamine in directing fluid across capillaries (12, 14). VEGF is responsible for disassembling adherens junctions between endothelial cells to permit cellular movement, wound repair, and angiogenesis (8, 10, 12, 13, 17, 26, 57). Extracellular domains of β3 integrins and VEGFR2 reportedly form a coprecipitable complex (3), and knocking out β3 causes capillary permeability that is augmented by VEGF addition (24, 47, 48). Pathogenic hantaviruses inhibit β3 integrin functions days after infection and similarly enhance the permeability of endothelial cells in response to VEGF (22).Adherens junctions form the primary fluid barrier of endothelial cells, and VEGFR2 responses control adherens junction disassembly (10, 17, 34, 57, 63). Vascular endothelial cadherin (VE-cadherin) is an endothelial cell-specific adherens junction protein and the primary determinant of paracellular permeability within the vascular endothelium (30, 33, 34). Activation of VEGFR2, another endothelial cell-specific protein, triggers signaling responses resulting in VE-cadherin disassembly and endocytosis, which increases the permeability of endothelial cell junctions (10, 12, 17, 34). VEGF is induced by hypoxic conditions and released by endothelial cells, platelets, and immune cells (2, 15, 38, 52). VEGF acts locally on endothelial cells through the autocrine or paracrine activation of VEGFR2, and the disassembly of endothelial cell adherens junctions increases the availability of nutrients to tissues and facilitates leukocyte trafficking and diapedesis (10, 12, 17, 55). The importance of endothelial cell barrier integrity is often in conflict with requirements for endothelial cells to move in order to permit angiogenesis and repair or cell and fluid egress, and as a result, VEGF-induced VE-cadherin responses are tightly controlled (10, 17, 18, 32, 33, 59). This limits capillary permeability while dynamically responding to a variety of endothelial cell-specific factors and conditions. However, if unregulated, this process can result in localized capillary permeability and edema (2, 9, 10, 12, 14, 17, 29, 60).Interestingly, tissue edema and hypoxia are common findings in both HPS and HFRS patients (11, 31, 62), and the ability of pathogenic hantaviruses to infect human endothelial cells provides a means for hantaviruses to directly alter normal VEGF-VE-cadherin regulation. In fact, the permeability of endothelial cells infected by pathogenic Andes virus (ANDV) or Hantaan virus (HTNV) is dramatically enhanced in response to VEGF addition (22). This response is absent from endothelial cells comparably infected with the nonpathogenic TULV and suggests that enhanced VEGF-induced endothelial cell permeability is a common underlying response of both HPS- and HFRS-causing hantaviruses (22). In these studies, we comparatively investigate responses of human endothelial cells infected with pathogenic ANDV and HTNV, as well as nonpathogenic TULV.  相似文献   

11.
12.
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   

13.
Claudin-1, a component of tight junctions between liver hepatocytes, is a hepatitis C virus (HCV) late-stage entry cofactor. To investigate the structural and functional roles of various claudin-1 domains in HCV entry, we applied a mutagenesis strategy. Putative functional intracellular claudin-1 domains were not important. However, we identified seven novel residues in the first extracellular loop that are critical for entry of HCV isolates drawn from six different subtypes. Most of the critical residues belong to the highly conserved claudin motif W30-GLW51-C54-C64. Alanine substitutions of these residues did not impair claudin-1 cell surface expression or lateral protein interactions within the plasma membrane, including claudin-1-claudin-1 and claudin-1-CD81 interactions. However, these mutants no longer localized to cell-cell contacts. Based on our observations, we propose that cell-cell contacts formed by claudin-1 may generate specialized membrane domains that are amenable to HCV entry.Hepatitis C virus (HCV) is a major human pathogen that affects approximately 3% of the global population, leading to cirrhosis and hepatocellular carcinoma in chronically infected individuals (5, 23, 42). Hepatocytes are the major target cells of HCV (11), and entry follows a complex cascade of interactions with several cellular factors (6, 8, 12, 17). Infectious viral particles are associated with lipoproteins and initially attach to target cells via glycosaminoglycans and the low-density lipoprotein receptor (1, 7, 31). These interactions are followed by direct binding of the E2 envelope glycoprotein to the scavenger receptor class B type I (SR-B1) and then to the CD81 tetraspanin (14, 15, 33, 36). Early studies showed that CD81 and SR-B1 were necessary but not sufficient for HCV entry, and claudin-1 was discovered to be a requisite HCV entry cofactor that appears to act at a very late stage of the process (18).Claudin-1 is a member of the claudin protein family that participates in the formation of tight junctions between adjacent cells (25, 30, 37). Tight junctions regulate the paracellular transport of solutes, water, and ions and also generate apical-basal cell polarity (25, 37). In the liver, the apical surfaces of hepatocytes form bile canaliculi, whereas the basolateral surfaces face the underside of the endothelial layer that lines liver sinusoids. Claudin-1 is highly expressed in tight junctions formed by liver hepatocytes as well as on all hepatoma cell lines that are permissive to HCV entry (18, 24, 28). Importantly, nonhepatic cell lines that are engineered to express claudin-1 become permissive to HCV entry (18). Claudin-6 and -9 are two other members of the human claudin family that enable HCV entry into nonpermissive cells (28, 43).The precise role of claudin-1 in HCV entry remains to be determined. A direct interaction between claudins and HCV particles or soluble E2 envelope glycoprotein has not been demonstrated (18; T. Dragic, unpublished data). It is possible that claudin-1 interacts with HCV entry receptors SR-B1 or CD81, thereby modulating their ability to bind to E2. Alternatively, claudin-1 may ferry the receptor-virus complex to fusion-permissive intracellular compartments. Recent studies show that claudin-1 colocalizes with the CD81 tetraspanin at the cell surface of permissive cell lines (22, 34, 41). With respect to nonpermissive cells, one group observed that claudin-1 was predominantly intracellular (41), whereas another reported associations of claudin-1 and CD81 at the cell surface, similar to what is observed in permissive cells (22).Claudins comprise four transmembrane domains along with two extracellular loops and two cytoplasmic domains (19, 20, 25, 30, 37). The first extracellular loop (ECL1) participates in pore formation and influences paracellular charge selectivity (25, 37). It has been shown that the ECL1 of claudin-1 is required for HCV entry (18). All human claudins comprise a highly conserved motif, W30-GLW51-C54-C64, in the crown of ECL1 (25, 37). The exact function of this domain is unknown, and we hypothesized that it is important for HCV entry. The second extracellular loop is required for the holding function and oligomerization of the protein (25). Claudin-1 also comprises various signaling domains and a PDZ binding motif in the intracellular C terminus that binds ZO-1, another major component of tight junctions (30, 32, 37). We further hypothesized that some of these domains may play a role in HCV entry.To understand the role of claudin-1 in HCV infection, we developed a mutagenesis strategy targeting the putative sites for internalization, glycosylation, palmitoylation, and phosphorylation. The functionality of these domains has been described by others (4, 16, 25, 35, 37, 40). We also mutagenized charged and bulky residues in ECL1, including all six residues within the highly conserved motif W30-GLW51-C54-C64. None of the intracellular domains were found to affect HCV entry. However, we identified seven residues in ECL1 that are critical for entry mediated by envelope glycoproteins derived from several HCV subtypes, including all six residues of the conserved motif. These mutants were still expressed at the cell surface and able to form lateral homophilic interactions within the plasma membrane as well as to engage in lateral interactions with CD81. In contrast, they no longer engaged in homophilic trans interactions at cell-cell contacts. We conclude that the highly conserved motif W30-GLW51-C54-C64 of claudin-1 is important for HCV entry into target cells and participates in the formation of cell-cell contacts.  相似文献   

14.
We present a simple strategy for isolating and accurately enumerating target DNA from high-clay-content soils: desorption with buffers, an optional magnetic capture hybridization step, and quantitation via real-time PCR. With the developed technique, μg quantities of DNA were extracted from mg samples of pure kaolinite and a field clay soil.Isolating and characterizing DNA sequences for use in molecular methods are integral to evaluating microbial community diversity in soil (6, 21, 22, 24, 37). Any isolation protocol should maximize nucleic acid isolation while minimizing copurification of enzymatic inhibitors. Although several methods that focus on extraction of total community DNA from environmental soil and water samples have been published (7, 21, 26, 34), the lack of a standard nucleic acid isolation protocol (32) reflects the difficulty in accomplishing these goals, most likely due to the complex nature of the soil environment.DNA extraction is especially difficult for soils containing clay (3, 5), given the tight binding of DNA strands to clay soil particles (7, 10, 20). Additionally, extracellular DNA binds to and is copurified with soil humic substances (10), which inhibit the activity of enzymes such as restriction endonucleases and DNA polymerase (6, 13, 23). Although clay-bound DNA can be PCR amplified in the absence of inhibitors (1), it is often the case that inhibitors are present in the soil environment, among them bilirubin, bile salts, urobilinogens, and polysaccharides (40). Of these inhibitors, humic substances have been found to be the most recalcitrant (36).A promising technique for isolating specific target sequences from soil particles and enzymatic inhibitors is the magnetic capture hybridization-PCR technique (MCH-PCR) presented by Jacobsen (19) and used to obtain high detection sensitivities (11, 38).We have found no evidence in the published literature of the use of MCH-PCR on soils that have high clay contents and here present a three-step strategy for isolating specific DNA sequences from the most difficult soil environment—clay that contains humic substances—and enumerating a specific target sequence from the crude extract.  相似文献   

15.
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   

16.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

17.
Double and triple uptake-type hydrogenase mutants were used to determine which hydrogenase recycles fermentatively produced hydrogen. The Δhyb Δhya and Δhyd Δhya double mutants evolved H2 at rates similar to that of the triple mutant strain, so Hya alone oxidizes the bulk of H2 produced during fermentation. When only Hya was present, no hydrogen production was observed in nutrient-limited medium. H2 uptake assays showed that Hya can oxidize both exogenously added H2 and formate hydrogen lyase-evolved H2 anaerobically. Even after anaerobic growth, all three uptake-type hydrogenases could function in the presence of oxygen, including using O2 as a terminal acceptor.Due to the anticipated scarcity of fossil fuels, there has been a surge of interest in H2 production for alternative energy means. Numerous studies have attempted to engineer H2-producing organisms, such as photosynthetic bacteria, cyanobacteria, and Escherichia coli, to produce maximal amounts of H2 while minimizing the H2-oxidizing capability of the organism (4, 5, 9, 10, 19). Hydrogenase expression and activity are controlled by multiple regulatory pathways and respond to fluctuations in pH, oxygen levels, and availability of metabolites and metal cofactors (17). In addition, the presence of hydrogen uptake hydrogenases decreases the net H2 yield even under conditions that favor H2 production. It is therefore important to understand the interactions between H2-oxidizing enzymes (i.e., respiratory hydrogenases) and H2-producing enzymes.Gene sequence analysis has revealed that many enteric bacteria contain the genes necessary for hydrogen production and oxidation. The E. coli hydrogenases have been studied extensively, while Salmonella enterica serovar Typhimurium hydrogenases have been studied to a lesser extent. Both E. coli and Salmonella serovar Typhimurium contain the hydrogen-oxidizing hydrogenases Hya and Hyb. Salmonella serovar Typhimurium also contains Hyd, which is another hydrogen-oxidizing hydrogenase (2, 13, 15). Hyc and Hyf are hydrogen-evolving hydrogenases that are present in both E. coli and Salmonella serovar Typhimurium, although it is unknown whether Hyf is functional (1).The Salmonella serovar Typhimurium hydrogenases are important for cellular metabolism. Hyc produces H2 in order to remove excess reductant generated during mixed-acid fermentation. Hyc and formate dehydrogenase constitute the formate hydrogen lyase (FHL) complex (16), which oxidizes formate to produce CO2 and H2 (12). The hyb genes in E. coli and Salmonella serovar Typhimurium are expressed at high levels under anaerobic respiration conditions, and Hyb probably contributes to energy conservation (11, 15, 20). Hyb oxidizes H2 and generates electrons, which are passed through the electron transport chain to terminal acceptors such as fumarate. The protons generated contribute to the proton-motive force. The role of Hya is not as well characterized. Hya may be used to recycle Hyc-produced H2, since the hya operon is expressed at high levels during fermentative growth, or it may play a role in acid stress resistance (6, 14, 20, 21). The hyb genes are expressed at high levels under aerobic conditions in Salmonella serovar Typhimurium, and Hyb may couple H2 oxidation to O2 reduction (20).Redwood et al. recently examined the roles of uptake-type hydrogenases on net hydrogen production in E. coli (10). Cells were pregrown aerobically or anaerobically with formate and then allowed to ferment in anaerobic bottles. H2 gas was collected, and other fermentation products were measured. They found that H2 production increased by 37% in an hya hyb double mutant (compared to that in the wild type) that was grown overnight aerobically with formate. This increase in production was associated with the loss of hyb and not hya. Therefore, in E. coli, Hyb may be responsible for recycling fermentatively produced H2.In this study, we measured the effect of uptake-type hydrogenase mutations on H2 production in Salmonella serovar Typhimurium. We found that the majority of H2-recycling activity in fermenting cells was dependent on the presence of hya, and having only Hya was sufficient to prevent any detectable H2 evolution. These results demonstrate yet another difference between H2 metabolism in E. coli and H2 metabolism in Salmonella serovar Typhimurium.  相似文献   

18.
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   

19.
The human immunodeficiency virus type 1 (HIV-1) matrix (MA) protein targets HIV-1 precursor Gag (PrGag) proteins to assembly sites at plasma membrane (PM) sites that are enriched in cholesterol and phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2]. MA is myristoylated, which enhances membrane binding, and specifically binds PI(4,5)P2 through headgroup and 2′ acyl chain contacts. MA also binds nucleic acids, although the significance of this association with regard to the viral life cycle is unclear. We have devised a novel MA binding assay and used it to examine MA interactions with membranes and nucleic acids. Our results indicate that cholesterol increases the selectivity of MA for PI(4,5)P2-containing membranes, that PI(4,5)P2 binding tolerates 2′ acyl chain variation, and that the MA myristate enhances membrane binding efficiency but not selectivity. We also observed that soluble PI(4,5)P2 analogues do not compete effectively with PI(4,5)P2-containing liposomes for MA binding but surprisingly do increase nonspecific binding to liposomes. Finally, we have demonstrated that PI(4,5)P2-containing liposomes successfully outcompete nucleic acids for MA binding, whereas other liposomes do not. These results support a model in which RNA binding protects MA from associating with inappropriate cellular membranes prior to PrGag delivery to PM assembly sites.The matrix (MA) domain of the human immunodeficiency virus type 1 (HIV-1) precursor Gag (PrGag) protein serves several functions in the viral replication cycle. One essential function is to target PrGag proteins to their assembly sites at the plasma membranes (PMs) of infected cells (4, 5, 11, 16, 25, 29, 30, 33, 35, 39, 43-45, 47, 50, 54, 56, 57). A second function is the recruitment of the viral surface/transmembrane (SU/TM; also referred to as gp120/gp41) envelope (Env) protein complex into virions (14, 15, 18, 19, 27, 51-53). In addition to these activities, numerous reports have attributed nucleic acid binding properties to retroviral MAs (24, 38, 47), and with some viruses MA appears to serve in an encapsidation capacity (24). While no encapsidation role has been assigned for HIV-1 MA, experiments have shown that MA can substitute for the HIV-1 nucleocapsid (NC) protein assembly function (38) under some circumstances, presumably by virtue of its facility to concentrate PrGag proteins by binding them to RNAs (38).A number of structural studies have been conducted on HIV-1 MA (1, 22, 41, 42, 49). The protein is N terminally myristoylated and composed of six α helices, capped by a three-strand β sheet (7, 22, 41, 42, 49). The protein trimerizes in solution and in crystals (22, 28, 49) and recently has been shown to organize as hexamers of trimers on lipid membranes (1). The membrane binding face of HIV-1 MA is basic, fostering its ability to associate with negatively charged phospholipid headgroups (1, 22, 30, 41, 42, 49). The importance of such an interaction has been underscored in molecular genetic experiments which demonstrated that depletion of PM phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2] reduced the assembly efficiency of HIV-1 (9, 36). Consistent with these observations, HIV-1 MA preferentially binds to soluble PI(4,5)P2 mimics through contacts with the headgroup and 2′ acyl chain, and binding promotes exposure of the MA myristate group and protein oligomerization (17, 21, 40-43, 46). However, PI(4,5)P2 is not the only lipid to demonstrate an association with HIV-1. In particular, HIV-1 appears to assemble at cholesterol-rich PM sites, cholesterol is highly enriched in HIV-1 virions, and cholesterol depletion reduces viral infectivity (2, 6, 8, 20, 23, 26, 31, 34, 37). The HIV-1 lipidome shows additional differences from the PM lipids of infected cells (2, 5, 8), suggesting that other lipids could affect PrGag-membrane binding or virus assembly site selection.To gain a better understanding of the functions and interactions of HIV-1 MA, we have examined the liposome and nucleic acid binding properties of purified myristoylated MA. Using liposome flotation assays and a novel liposome bead binding assay, we have demonstrated that the PI(4,5)P2 binding specificity of MA is enhanced by cholesterol, that protein myristoylation increases membrane binding efficiency but not specificity, and that 2′ acyl chain variation is compatible with PI(4,5)P2 binding. We also examined whether soluble PI(4,5)P2 mimics could compete with liposomes for MA binding. Surprisingly, we found that soluble mimics not only failed to compete with PI(4,5)P2 liposomes but also increased MA binding to membranes that do not contain acidic phospholipids. Finally, we have observed that while MA does bind nucleic acids, nucleic acid binding is outcompeted by PI(4,5)P2-containing liposomes. Our results suggest models for PrGag-membrane and RNA association and the HIV-1 assembly pathway.  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